Skip to main content

Full text of "Organic Chemistry: Methane to Macromolecules"

See other formats


Organic chemistry 



ORGANIC 
CHEMISTRY 

methane to macromolecules 



JOHN D. ROBERTS 

California Institute of Technology 

ROSS STEWART 

University of British Columbia 

MARJORIE C. CASERIO 

University of California, Irvine 



$% 



«o7 i x** 



W. A. BENJAMIN, INC. 
New York 1971 



Organic chemistry: Methane to macromolecules 

Copyright © 1971 by W. A. Benjamin, Inc. 

All rights reserved 

Standard Book Number 8053-8332-8 

Library of Congress Catalog Card Number 71-130356 

Manufactured in the United States of America 

12345K54321 

Portions of this book appeared previously in 

Modern Organic Chemistry 

by John D. Roberts and Marjorie C. Caserio, 

published by W. A. Benjamin, Inc. 

1967, New York 

W. A. BENJAMIN, INC. 

New York, New York 10016 



preface 



The success achieved by this book's forerunners, Basic Principles of Organic 
Chemistry and Modern Organic Chemistry, was to a considerable extent due 
to the rigor with which the subject of organic chemistry was presented. In the 
present work we have tried to paint an interesting, relevant, and up-to-date 
picture of organic chemistry while retaining the rigorous approach of the 
earlier books. 

Organic chemistry sometimes appears to be enormously complex to the 
beginning student, particularly if he must immediately grapple with the 
subjects of structural isomerism and nomenclature. We have attempted to 
avoid this difficulty in the following way. Chapter 1 briefly relates carbon to 
its neighbors in the Periodic Table and reviews some fundamental concepts. 
Chapter 2 deals with the four C t and C 2 hydrocarbons — methane, ethane, 
ethene, and ethyne — and discusses their conformational and configurational 
properties and some of their chemical reactions. The reader thus makes an 
acquaintance with the properties of some important organic compounds before 
dealing in an open-ended way with families of compounds — alkanes, alcohols, 
etc. 

A heavy emphasis on spectroscopy is retained but the subject is introduced 
somewhat later than in the earlier books. Important additions are chapters 
dealing with enzymic processes and metabolism and with cyclization reac- 
tions. Many of the exercises of the earlier books have been retained and have 
been supplemented with drill-type problems. 

It seems a shame to burden the mind of the beginning student with trivial 
names, some of them quite illogical, and throughout we have stressed IUPAC 
nomenclature, which is both logical and easy to learn. The instructor, who 
may well carry lightly the excess baggage of redundant names, may occasion- 
ally find this irritating but we ask him to consider the larger good. As a further 
aid to the student, each chapter concludes with a summary of important 
points. 

The simple introduction to the subject and the emphasis on relevance, 
particularly to living systems, should make the book appealing to the general 
student. At the same time we hope that the up-to-date and more advanced 
topics that are included — the effect of orbital symmetry on cyclization reac- 
tions, for example — will also appeal to the chemistry specialist. 

We should like to acknowledge the help of many persons who read all or 
parts of the manuscript and offered sound advice. Professor George E. Hall 
read the manuscript at several stages of revision and we are particularly 



preface vi 

grateful to him. Others who helped us were Drs. E. Caress, L. D. Hall, D. N. 
Harpp, J. P. Kutney, T. Money, M. Smith, T. Spencer, and L. S. Weiler. 

We conclude this preface on a mildly philosophical note. The world of to- 
morrow will result from the interplay of powerful forces — some social, some 
technological. Responsible public action requires public knowledge and there 
are few areas of science that impinge more on the life around us than does 
organic chemistry. We hope that those who study this book will utilize their 
knowledge responsibly for the benefit of all who come after. 

JOHN D. ROBERTS 
ROSS STEWART 
MARJORIE C. CASERIO 

Pasadena, California 
Vancouver, British Columbia 
Irvine, California 



contents 



Chapter i Introduction I 

1-1 Bonding in Organic Compounds 5 

1-2 Methane, Ammonia, Water, and Hydrogen Fluoride 6 

Summary 13 

Exercises ^ 

Chapter 2 The C t and C 2 hydrocarbons 17 

2-1 Molecular Shape of CH 4 , C 2 H 6 , C 2 H 4 , and C 2 H 2 20 

2-2 Rotational Conformations of Ethane 21 

2-3 Space-Filling Models 22 

Chemical Reactions of the C t and C 2 Hydrocarbons 23 

2-4 Combustion 23 

2-5 Substitution Reactions of Saturated Hydrocarbons 26 

2-6 Addition Reactions of Unsaturated Hydrocarbons 34 

Summary 40 

Exercises 41 

Chapter 3 Alkanes 45 

3-1 Nomenclature 47 

3-2 Physical Properties of Alkanes— Concept of Homology 53 

3-3 Alkanes and Their Chemical Reactions 56 

3-4 Cycloalkanes 62 

Summary ^2 

Exercises ^4 

Chapter 4 Alkenes 79 

4-1 Nomenclature 81 

4-2 Isomerism in C 4 H 8 Compounds 83 

4-3 Cis and Trans Isomers 84 

4-4 Chemical Reactions of Alkenes 86 

Summary 103 

Exercises 105 

Chapter 5 Alkynes 109 

5-1 Nomenclature HI 

5-2 Physical Properties of Alkynes 112 



contents viii 



5-3 Ethyne 2J2 

5-4 Addition Reactions of Alkynes 113 

5-5 Alkynes as Acids U6 

5-6 Synthesis of Organic Compounds 117 

Summary 120 

Exercises 121 

Chapter 6 Bonding in conjugated unsaturated 

systems 1 25 

6-1 Bonding in Benzene 129 

6-2 Conjugate Addition 133 

6-3 Stabilization of Conjugated Dienes 135 

6-4 Stabilization of Cations and Anions 137 

6-5 Vinyl Halides and Ethers 139 

6-6 Rules for the Resonance Method 140 

6-7 Molecular Orbital Method of Htickel 140 

Summary 145 

Exercises 147 

Chapter 7 Isolation and identification of 

organic compounds 151 

7-1 Isolation and Purification 153 

7-2 Identification of Organic Compounds 156 

Spectroscopy 159 

7-3 Absorption of Electromagnetic Radiation 159 

7-4 Infrared Spectroscopy 161 

7-5 Ultraviolet and Visible Spectroscopy (Electronic Spectroscopy) 165 

7-6 Nuclear Magnetic Resonance Spectroscopy 168 

Summary 179 

Exercises 180 

Chapter 8 Nucleophilic displacement and 

elimination reactions 185 

8-1 Organic Derivatives of Inorganic Compounds 187 

8-2 Alcohol Nomenclature 187 

8-3 Ether Nomenclature 190 

8-4 Carboxylic Acid Nomenclature 190 

8-5 The Use of Greek Letters to Denote Substituent Positions 191 

8-6 Single- or Multiple-Word Names 191 

Nucleophilic Displacement Reactions 192 

8-7 General Considerations 192 

8-8 Mechanisms of S N Displacements 195 

8-9 Energetics of S N 1 and S N 2 Reactions 197 

8-10 Stereochemistry of S N 2 Displacements 200 

8-11 Structural and Solvent Effects in S N Reactions 201 



contents ix 

Elimination Reactions 205 

8-12 The E2 Reaction 206 

8-13 The El Reaction 207 

Summary 209 

Exercises 210 

Chapter 9 Alkyl halides and organometallic 

compounds 215 

9T Physical Properties 217 

9-2 Spectra 217 

9-3 Preparation of Alkyl Halides 218 

9-4 Reaction of Alkyl Halides 219 

9-5 Vinyl Halides 219 

9-6 Allyl Halides ^ 220 

9-7 Polyhalogen Compounds 222 

9-8 Fluorinated Alkanes 224 

9-9 Organometallic Compounds 226 

Summary 236 

Exercises 237 

Chapter 10 Alcohols and ethers 243 

10T Physical Properties of Alcohols 245 

10-2 Spectroscopic Properties of Alcohols — Hydrogen Bonding 247 

10-3 Preparation of Alcohols 249 

Chemical Reactions of Alcohols 251 

10-4 Reactions Involving the O— H Bond 251 

10-5 Reactions Involving the C—O Bond of Alcohols 255 

10-6 Oxidation of Alcohols 259 

10-7 Polyhydroxy Alcohols 260 

10-8 Unsaturated Alcohols 262 

Ethers 262 

10-9 Preparation of Ethers 263 

10-10 Reactions of Ethers 264 

10-11 Cyclic Ethers 265 

Summary 266 

Exercises 267 
Chapter 11 Aldehydes and ketones I. 

Reactions at the carbonyl group 273 

11-1 Nomenclature of Aldehydes and Ketones 275 

1 1 -2 Carbonyl Groups of Aldehydes and Ketones 275 

11-3 Preparation of Aldehydes and Ketones 278 

11-4 Reactions of Aldehydes and Ketones 281 

Summary 294 

Exercises 296 



Chapter 12 Aldehydes and ketones II. 

Reactions involving substituent 

groups. Polycarbonyl compounds 301 

12-1 Halogenation of Aldehydes and Ketones 303 

12-2 Reactions of Enolate Anions 306 

Unsaturated Carbonyl Compounds 310 

12-3 a,j6-Unsaturated Aldehydes and Ketones 311 

12-4 Ketenes 312 

Polycarbonyl Compounds 313 

12-5 1 ,2-Dicarbonyl Compounds 313 

12-6 1,3-Dicarbonyl Compounds 314 

Summary 316 

Exercises 318 

Chapter 13 Carboxylic acids and derivatives 327 

13-1 Physical Properties of Carboxylic Acids 330 

13-2 Spectra of Carboxylic Acids 334 

13-3 Preparation of Carboxylic Acids 336 

13-4 Dissociation of Carboxylic Acids 336 

13-5 Reactions at the Carbonyl Carbon of Carboxylic Acids 339 

13-6 Decarboxylation of Carboxylic Acids 341 

13-7 Reactions at the 2 Position* of Carboxylic Acids 342 

Functional Derivatives of Carboxylic Acids 344 

13-8 Displacement Reactions of Acid Derivatives 344 
13-9 Reactions at the 2 Position (oc position) of Carboxylic Acid 

Derivatives 350 
13-10 Reactions of Unsaturated Carboxylic Acids and Their 

Derivatives 355 

13-11 Dicarboxylic Acids 356 

Summary 359 

Exercises 361 
Chapter 14 Optical isomerism. Enantiomers 

and diastereomers 367 

14-1 Plane-Polarized Light and the Origin of Optical Rotation 369 

14-2 Specific Rotation 371 
14-3 Optically Active Compounds with Asymmetric Carbon Atoms 372 
14-4 Optically Active Compounds Having No Asymmetric Carbon 

Atoms 379 

14-5 Absolute and Relative Configuration 381 

14-6 Separation or Resolution of Enantiomers 384 

14-7 Asymmetric Synthesis and Asymmetric Induction 386 



contents xi 

14-8 Racemization 388 

14-9 Inversion of Configuration 389 

14-10 Optical Rotatory Dispersion 389 

Summary 391 

Exercises 393 

Chapter 15 Carbohydrates 397 

15-1 Classification of Carbohydrates 400 

15-2 Glucose 402 

15-3 Cyclic Structures 405 

15-4 Mutarotation 4° 6 

15-5 Glycosides 407 

15-6 Disaccharides 408 

15-7 Polysaccharides 410 

15-8 Vitamin C 413 

15-9 Immunologically Important Carbohydrates 413 

Summary 4 ^ 4 

Exercises 4 *° 

Chapter 16 Organic nitrogen compounds 419 

16-1 Amines 421 

16-2 Amides 434 

16-3 Nitriles 440 

16-4 Nitroso Compounds 440 

16-5 Nitro Compounds 441 

16-6 Some Compounds with Nitrogen-Nitrogen Bonds 442 

Summary 

Exercises 

Chapter 17 Amino acids, proteins, and nucleic 

acids 455 

17-1 Amino Acids 457 

17-2 Lactams 467 

17-3 Peptides 468 

17-4 Protein Structures 474 

17-5 Biosynthesis of Proteins 477 

17-6 The Structure of DNA 477 

17-7 Genetic Control and the Replication of DNA 483 

17-8 Chemical Evolution 486 

Summary 487 

Exercises 489 

Chapter 18 Enzymic processes and metabolism 493 

18-1 Catalysis in Organic Systems 495 

18-2 Enzymes and Coenzymes 503 

18-3 Hydrolytic Enzymes 504 



444 
446 



contents xii 

18-4 Oxidative Enzymes 506 

18-5 The Energetics of Metabolic Processes 509 

Summary 511 

Exercises 513 

Chapter 19 Organic compounds of sulfur, 

phosphorus, silicon and boron 515 

19-1 d Orbitals and Chemical Bonds 517 

19-2 Types and Nomenclature of Organic Compounds of Sulfur 520 

19-3 Phosphorus Compounds 528 

19-4 Organosilicon Compounds 531 

19-5 Organoboron Compounds 536 

Summary 540 

Exercises 542 

Chapter 20 Arenes. Electrophilic aromatic 

substitution 547 

20 T Nomenclature of Arenes 549 

20-2 Physical Properties of Arenes 553 

20-3 Spectroscopic Properties of Arenes 554 

20-4 Reactions of Aromatic Hydrocarbons 559 
20-5 Effect of Substituents on Reactivity and Orientation 

in Electrophilic Aromatic Substitution 567 
20-6 Substitution Reactions of Polynuclear Aromatic 

Hydrocarbons 574 

20-7 Nonbenzenoid Conjugated Cyclic Compounds 578 

Summary 579 

Exercises - 580 

Chapter 21 Aryl halogen compounds. Nucleo- 

philic aromatic substitution 587 

21 T Physical Properties of Aryl Halogen Compounds 590 

21-2 Preparation of Aryl Halides 590 

21-3 Reactions of Aryl Halides 592 

21-4 Organochlorine Pesticides 596 

Summary 598 

Exercises 599 

Chapter 22 Aryl nitrogen compounds 603 

Aromatic Nitro Compounds 606 

22-1 Synthesis of Nitro Compounds 606 

22-2 Reduction of Aromatic Nitro Compounds 608 

22-3 Polynitro Compounds 611 

22-4 Charge-Transfer and n Complexes 612 

Aromatic Amines $14 

22-5 General Properties 614 



contents xiii 

22-6 Aromatic Amines with Nitrous Acid 616 

Diazonium Salts 617 

22-7 Preparation and General Properties 617 

22-8 Replacement Reactions of Diazonium Salts 618 
22-9 Reactions of Diazonium Compounds that Occur Without 

Loss of Nitrogen 619 

Summary 620 

Exercises 621 

Chapter 23 Aryl oxygen compounds 625 

23-1 Synthesis and Physical Properties of Phenols 627 

23-2 Some Chemical Properties of Phenols 630 

23-3 Polyhydric Phenols 635 

23-4 Quinones 637 

23-5 Tropolones and Related Compounds 641 

Summary 642 

Exercises 644 

Chapter 24 Aromatic side-chain derivatives 647 

Preparation of Aromatic Side-Chain Compounds 649 

24- 1 Aromatic Carboxylic Acids 649 

24-2 Preparation of Side-Chain Aromatic Halogen Compounds 650 

24-3 Side-Chain Compounds Derived from Arylmethyl Halides 650 
24-4 Preparation of Aromatic Side-Chain Compounds by Ring 

Substitution 651 

Properties of Aromatic Side-Chain Derivatives 653 

24-5 Arylmethyl Halides. Stable Carbonium Ions, Carbanions, 

and Radicals 653 

24-6 Aromatic Aldehydes 656 
24-7 Natural Occurrence and Uses of Aromatic Side-Chain 

Derivatives 657 

24-8 Electron Paramagnetic Resonance (epr) Spectroscopy 659 

24-9 Linear Free-Energy Relations 661 

Summary 665 

Exercises 666 

Chapter 25 Heterocyclic compounds 669 

25-1 Aromatic Character of Pyrrole, Furan, and Thiophene 672 
25-2 Chemical Properties of Pyrrole, Furan, Thiophene, and 

Pyridine 673 

25-3 Polycyclic and Polyhetero Systems 679 

Heterocyclic Natural Products 680 

25-4 Natural Products Related to Pyrrole 680 

25-5 Natural Products Related to Indole 682 
25-6 Natural Products Related to Pyridine, Quinoline, and 

Isoquinoline 684 

25-7 Natural Products Related to Pyrimidine 685 



contents xiv 

25-8 Natural Products Related to Purine and Pteridine 686 

25-9 Natural Products Related to Pyran 686 

25-10 Polyhetero Natural Products 688 

Summary 688 

Exercises 689 

Chapter 26 Photochemistry 693 

26-1 Light Absorption, Fluorescence, and Phosphorescence 696 

26-2 Light Absorption and Structure 699 

26-3 Photodissociation Reactions 703 

26-4 Photochemical Reduction 704 

26-5 Photochemical Oxidation 705 
26-6 Photochemical Isomerization of Cis- and Trans- 

Unsaturated Compounds 707 

26-7 Photochemical Cycloadditions 707 

Summary 708 

Exercises 709 

Chapter 27 Cyclization reactions 713 

27-1 Cyclization Reactions of Carbonyl Compounds 715 

27-2 Cycloaddition Reactions of Carbon-Carbon Multiple Bonds 718 

27-3 Fluxional Systems 723 

27-4 Annulenes 725 

27-5 Orbital Symmetry and Cycloaddition 726 

Chapter 28 Polymers 735 

28-1 Types of Polymers 737 

Physical Properties of Polymers 739 

28-2 Forces Between Polymer Chains 739 

28-3 Correlation of Polymer Properties with Structure 743 

Preparation of Synthetic Polymers 748 

28-4 Condensation Polymers 748 

28-5 Addition Polymers 753 

28-6 Naturally Occurring Polymers 756 

28-7 Dyeing of Fibrous Polymers 758 
Chapter 29 Some aspects of the chemistry 

of natural products 767 

29-1 Civetone 769 
29-2 Spectroscopic Methods in the Determination of the 

Structures of Natural Products 772 

29-3 Terpenes 778 

29-4 Steroids 782 

29-5 Biogenesis of Terpenes and Steroids 789 

Index 801 




immm ®m&® m*%m ip§is «M ^ Witt 

\K£?#$M& .::\&.&M f>& .>% -v^ ^#i?S*S| £MD%&*S$M #ftiii*ll 






K-viffc;^ 



$'$ €'#'- 






s^i*? : 'v.U :€'■":.!•*'.!/£ iJ^^L"^ SSSi 







>Mi 


: $ ->^* ; V'/ 


iHi 


**.£ 


■ft !'**-*r\t i ' 


H 


:¥'<? 




■ 


£§ 

"*£** 


2t' ***'«'■ 


BH 




I^rTi*'' 1 *"?' 


H 






Sf 


i*SS* 


•V- $£ 


I 






■ 


•'**-••; 



fr. i 



i5" TCi*3jii*« vjra^^iiiN is 



ft^'i, 










it& 



•.V.Y.A 



: **V •''■:• >\ 


^S.t ..*. ^?'"- 


T ':'r£~ r S 


.v fc r. ^ 


J"i-V r> » ***'Ci 




ii't'r !w§>* 


'•**<-K %*.^ 


l'£- ■' 




i'VtV^ 1V-' 






;*;3*y 2 r ■;. 


••.^V^/M 


. ". . J 


W'li 




^*5ft 


isiis 


T-iiA? : 















, ipiift s&^ss ^w« ite#» israw ins ^^fe 









chap 1 introduction 3 

Twenty-two centuries ago the Greek mathematician Euclid wrote a textbook 
on geometry that is still in use today. Some 300 years ago Isaac Newton 
discovered the principles of mechanics that can still be applied with great 
precision to most macroscopic systems. By contrast, chemistry, and in 
particular organic chemistry, is in its infancy as a precise science. The study of 
molecular science — because this is what chemistry essentially is — depends on 
inferences about submicroscopic bodies drawn from observations of macro- 
scopic behavior. The speculations of the early chemists are best described as 
"haywire " rather than as " incorrect " and it was only in the last century that 
the foundations of chemical theory became firmly established. Since that time 
enormous strides have been made in extending chemical knowledge. And 
today, some 1,100,000 organic compounds alone (compounds containing 
carbon) have been prepared, their structures elucidated, and their properties 
examined. 

The flowering of organic chemistry in the past hundred years followed two 
events during the last century. The first occurred in 1828, when the German 
chemist Wohler discovered that ammonium cyanate (NH 4 ®CNO e ) could be 

O 

II 
converted to the compound urea (NH 2 CNH 2 ). The former was a typical salt 

and is considered part of the mineral world, whereas urea was a product of 
animal metabolism and therefore part of the living or organic world. The 
realization gradually followed that the boundary between living and nonliving 
systems could be crossed, and this provided the impetus for intensive investi- 
gation of the substances found in nature: the term " organic " was thus applied 
to all compounds of carbon whether they were found in nature or were pre- 
pared in the laboratory. 

The second important occurrence in the last century as far as organic 
chemistry was concerned was the recognition, achieved between 1858 and 
1872, that unique, three-dimensional structures could be drawn for the mole- 
cules of every known compound of carbon. This realization followed essenti- 
ally the work of six men : Avogadro, Cannizzaro, Kekule, Couper, LeBel, and 
van't Hoff. Avogadro and Cannizzaro distinguished between what we would 
now call empirical and molecular formulas. Avogadro's hypothesis that equal 
volumes of gases contained equal numbers of molecules was actually made in 
1811 but it was not until 1860 that the Italian chemist Cannizzaro made use of 
this idea to distinguish between different compounds having the same 
composition. Thus, the distinction between the molecules C 2 H 4 and C 4 H 8 
became clear, and it was realized that neither had the molecular formula 
CH 2 , though both had this empirical formula. At about the same time Kekule, 
a German, and Couper, a Scot, suggested that carbon was tetravalent, always 
forming four bonds. Two young chemists, LeBel and van't Hoff, then in- 
dependently provided convincing proof that these four bonds are tetrahedrally 
arranged around the carbon atom (Section 14-6). 

Part of the fascination of organic chemistry comes from the knowledge that 
the whole complex edifice has been built up from indirect study of molecular 
behavior, that is, microscopic understanding from macroscopic observation. 
The advent in recent years of sophisticated instrumental techniques for ex- 



chap 1 introduction 4 

amining structure has confirmed the vast majority of the structures assigned 
to organic compounds in the late nineteenth century. Spectroscopy and X-ray 
crystallography, in particular, have become powerful tools for checking 
previously assigned structures and for elucidating structures of compounds 
newly prepared in the laboratory or found in nature. 

Considerable use will be made in this book of spectroscopy — the study of 
how "light" (to be more exact, electromagnetic radiation) is absorbed by 
matter. The infrared, ultraviolet, and nuclear magnetic resonance spectra may 
permit the correct structure of a fairly complex compound to be assigned in a 
matter of hours. If, at this point, a search of the chemical literature reveals 
that the compound with the suspected structure has been previously prepared, 
it is only necessary to compare the reported physical or spectroscopic pro- 
perties with those of the substance being examined. However, if the compound 
has not been previously reported, it must be synthesized from starting materials 
of known structure before its structure is really considered to be proven. 

Organic chemistry occupies a central position in the undergraduate science 
curriculum. It is, of course, an important branch of knowledge in its own 
right, but in addition it is the foundation for basic studies in botany, zoology, 
microbiology, nutrition, forestry, agricultural sciences, dentistry, and medicine. 
Antibiotics, vitamins, hormones, sugars, and proteins are only a few of the 
important classes of chemical substances that are organic. In addition, many 
industrially important products are organic — plastics, rubber, petroleum 
products, most explosives, perfumes, flavors, and synthetic fibers. It is little 
wonder that there are more organic chemists than chemists of any other kind. 
Of the 2000 or so Ph.D.'s awarded each year in chemistry in North America, 
about half go to those whose research has been in organic chemistry. The 
research may have involved the synthesis of a new compound of unusual 
structure, the elucidation of the structure of a new compound extracted from 
a plant, or the discovery of the reaction path that is followed when one 
compound is converted to another. 

Only a few years ago most people regarded the effects of chemical tech- 
nology as being wholly beneficial. Pesticides, herbicides, plastics, and synthetic 
drugs all seemed to contribute in large or small measure to human welfare. 
Recently we have come to realize, however, that our environment cannot 
indefinitely accommodate all the products that are being added to it without 
being damaged. A pesticide may be extremely effective at eradicating some 
of man's enemies but may seriously endanger some of man's friends. A 
cogent example is the way the potent insecticide DDT (Section 24-7) causes 
bird egg shells to become thin. Further, a synthetic plastic may be immune to 
sunlight, rain, and bacterial decay but this is a doubtful benefit after the object 
made from it has been discarded in a park. 

Many of the substances that have been added in large amounts to our 
environment in recent years are synthetic organic chemicals. Some are harmful 
to man and to life in general; some are not. An understanding of the proper- 
ties and reactions of organic compounds will help us assess the possible perils 
associated with new processes or products and will help us develop suitable 
control procedures. The grave consequences of a polluted environment can 
only be avoided by the wise application of chemical knowledge. 



sec 1.1 bonding in organic compounds 5 



1 • 1 bonding in organic compounds 



Why is carbon unique ? What accounts for the apparently limitless number of 
carbon compounds that can be prepared? The answer is that bonds between 
carbon atoms are stable, allowing chains of carbon atoms to be formed, with 
each carbon atom of a chain being capable of joining to other atoms such as 
hydrogen, oxygen, sulfur, nitrogen, and the halogens. Neighboring atoms in 
the periodic table, such as boron, silicon, sulfur, and phosphorus, can also 
bond to themselves to form chains in the elemental state, but the resulting 
compounds are generally quite unstable and highly reactive when atoms of 
hydrogen or halogen, for example, are attached to them. The elements at the 
right or left of the periodic table do not form chains at all — their electron- 
attracting or electron-repelling properties are too great. 

The forces that hold atoms and groups of atoms together are the electrostatic 
forces of attraction between positively charged nuclei and negatively charged 
electrons on different atoms. We usually recognize two kinds of binding. 
The first is the familiar ionic bond that holds a crystal of sodium chloride 
together. Each Na® in the crystal feels a force of attraction to each Cl e , the 
force decreasing as the distance increases. (Repulsion between ions of the same 
sign of charge is also present, of course, but the stable crystal arrangement has 
more attraction than repulsion.) Thus, you cannot identify a sodium chloride 
pair as being a molecule of sodium chloride. Similarly, in an aqueous solution 
of sodium chloride, each sodium ion and chloride ion move in the resultant 
electric field of all the other ions in the solution. Sodium chloride, like other 
salts, can be vaporized at high temperatures. The boiling point of sodium 
chloride is 1400 . 1 For sodium chloride vapor, you can at last speak of sodium 
chloride molecules which, in fact, are pairs of ions, Na e Cl e . Enormous energy 
is required to vaporize the salt because in the vapor state each ion interacts 
with just one partner instead of many. 

The second kind of bonding referred to above results from the simul- 
taneous interaction of a pair of electrons (or, less frequently, just one 
electron) with two nuclei, and is called the covalent bond. Whereas metallic 
sodium reacts with chlorine by completely transferring an electron to it to 
form Na® and Cl e , the elements toward the middle of the rows of the periodic 
table tend to react with each other by sharing electrons. 

Transfer of an electron from a sodium atom to a chlorine atom produces 
two ions, each of which now possesses an octet of electrons. This means of 
achieving an octet of electrons is not open to an element such as carbon, which 
has two electrons in a filled inner K shell and four valence electrons in the 
outer L shell. A quadrinegative ion C 4e with an octet of electrons in the 
valence shell would have an enormous concentration of charge and be of 
very high energy. Similarly, the quadripositive ion C 4 ®, which would have a 
filled K shell like helium, would be equally unstable. Carbon (and to a great 
extent boron, nitrogen, oxygen, and the halogens) completes its valence-shell 
octet by sharing electrons with other atoms. 

"In this book all temperatures are given in degrees centigrade unless otherwise noted. 



chap 1 introduction 6 



In compounds with shared electron bonds (or covalent bonds) such as 
methane (CH 4 ) or tetrafluoromethane (CF 4 ), carbon has its valence shell 
filled, as shown in these Lewis structures : 

H :F: 

H:CH :F:C:F: 

H :F:" 

methane tetrafluoromethane 

(carbon tetrafluoride) 

For convenience, these molecules are usually written with each bonding 
pair of electrons represented by a dash: 

H F 

I I 

H— C-H F-C-F 

I I 

H F 



1-2 methane, ammonia, water, and hydrogen fluoride 

The elements in the first main row of the periodic table are 





Li 


Be 


B 


C N 





F Ne 


Atomic number 


3 


4 


5 


6 7 


8 


9 10 


Number of valence 














electrons 


1 


2 


3 


4 5 


6 


7 (8) 



Lithium and beryllium are able to form positive ions by loss of one or two 
electrons, respectively. Boron is in an intermediate position and its somewhat 
unusual bonding properties are considered later in the book (Section 19-5). 
Carbon, nitrogen, oxygen, and fluorine all have the ability to form covalent 
bonds because each can complete its octet by sharing electrons with other 
atoms. (Fluorine or oxygen can also exist as stable anions in compounds such 
as Na e F e or Na ffi OH e .) 

The degree of sharing of electrons in a covalent bond will not be exactly 
equal if the elements being linked are different. The relative attractive power 
exerted by an element on the electrons in a covalent bond can be expressed 
by its electronegativity. In one quantitative definition of electronegativity we 
have an increase in electronegativity along the series toward fluorine as 
follows : 

C N O F 

2.5 3.0 3.5 4.0 

The electronegativity of hydrogen is 2.0, close to that for carbon. Each 
covalent bond between elements with different electronegativities will have 
the bonding electrons unequally shared between them, which leads to what is 
called polar character. In a carbon-fluorine bond the pair of electrons are 



sec 1.2 methane, ammonia, water, and hydrogen fluoride 7 

attracted more to the fluorine nucleus than to the carbon nucleus. The regions 
of space occupied by electrons are called orbitals and, in a molecule such as 
CF 4 , the pair of electrons in the orbital that represents each covalent bond 
will not be divided equally between the carbon and fluorine but will be polar- 
ised towards fluorine. 



We say that such a bond is dipolar and it can be represented, when necessary, 
by the symbols 



\S© Se 




\ 




-C-F 


or 


— C- 


-»F 


/ 




/ 





28© 


Se 28© Se 





F— — F 


e / \ So 




F F 





The electronegativity of oxygen is less than that of fluorine and closer to that 
of carbon; therefore the polarity of a C — O bond will be less than that of a 
C— F bond. Clearly the polarity of a C — N bond will be smaller still. 

Even though a molecule contains polar bonds, the molecule itself may be 
nonpolar, that is, not possess a dipole moment. This will occur when the 
molecule has a shape (or symmetry) such that the dipoles of the individual 
bonds cancel each other. Thus, molecules such as F— O— F and H— O— H 
will have dipole moments (as they do) if the angle between the two F— O (or 
H— O) bonds is different from 180°, and zero dipole moment if the angle is 
180°. To predict whether a molecule has a dipole moment, it is therefore 



28e S© 28e S© 

A ~ H-O-H 

B / \ 8© 

H H 



necessary to know its shape, and in the next section the principles governing 
the shapes of covalently bound molecules are considered, with special refer- 
ence to the series CH 4 , NH 3 , H 2 0, and HF. 

If a substance is a liquid, it is an easy matter to show experimentally 
whether its molecules are polar and have a dipole moment. All you have to do 
is to hold an object carrying an electrostatic charge near a fine stream of the 
falling liquid and note whether the stream is deflected. The charged object 
can be as simple as a glass rod rubbed on silk or an amber rod rubbed on 
cat's fur, the charge being positive in the first case and negative in the second. 
A fine stream of water is sharply deflected by such an object and this shows 
that the individual molecules in the liquid have positive and negative ends. 
The molecules tend to orient themselves so that the appropriately charged 
end is directed towards the charged object (for example, the negative end, 
oxygen, toward the positively charged rod) and then the electrostatic attrac- 
tion draws the molecules toward the rod. A fine stream of carbon tetra- 



chap 1 introduction 8 

chloride (tetrachloromethane), CC1 4 , cannot be deflected at all. This shows 
that the CC1 4 molecule is sufficiently symmetrical in its arrangement of the 
four carbon-chlorine bonds so that the polarities of these bonds cancel each 
other. 



A. MOLECULAR SHAPES 

It is important to recognize that an understanding of the shapes of organic 
compounds is absolutely vital to understanding the physical, chemical, and 
biochemical properties of organic compounds. We present in this section a 
few simple concepts which will turn out later to be of great utility in predicting 
and correlating the shapes of complex organic molecules. 

The compounds CH 4 , NH 3 , H 2 0, and HF are all isoelectronic : they have 
the same number of electrons, 10. Two are in the inner K shell of the central 
atom and eight are in the valence, or bonding, shell. The bonding arrange- 
ments can be indicated by Lewis structures : 

H 
H:C=H H:N:H H:6:H H=F: 

H H 

Carbon, nitrogen, oxygen, and fluorine have, respectively, contributed four, 
five, six, and seven of the electrons that make up the octet. Because no more 
than two electrons can occupy an orbital, we will expect that the electrons in 
the octet can be treated as four distinct pairs. The electron pairs repel one 
another and, if the four pairs are to get as far away from each other as 
possible, we will expect to find the four orbitals directed toward the corners 
of a tetrahedron, because this provides the maximum separation between the 



^T 



tetrahedral tetrahedral angle 

arrangement of 
electron pairs 

electrons. Methane, CH 4 , is in fact tetrahedral, as is tetrafluoromethane, 
CF 4 . The three bonds in ammonia and the two bonds in water are directed at 
slightly different angles, 106.6° and 104.5°, respectively. This is reasonable 
because the repulsions between the four pairs of electrons in each of these 

ft 

H H 

bond angle 106.6° bond angle 104.5° 

molecules will not be the same. Thus, for water, two of the four orbitals 



H 


H 


1 


\ 


c— 


.C 


1 


H'7 


H 


H 



sec 1.2 methane, ammonia, water, and hydrogen fluoride 9 

contain protons and two do not. We expect somewhat greater repulsions 
between the nonbonding pairs than between bonding pairs, and this results 
in the angle of the bonding pairs being somewhat less than the tetrahedral 
value. 

Replacement of any of the hydrogen atoms in the three molecules CH 4 , 
NH 3 , and H 2 with another kind of group will alter the bond angles to some 
extent. Replacement of one or more such hydrogens by the methyl group 
(methane minus a hydrogen atom), CH 3 — , gives the structures shown in Table 
1-1. The methyl group is an especially important substituent group and can be 
conveniently represented in three ways, the last being a three-dimensional 
representation. 



CH, 



The four derivatives of methane shown in Table IT are all hydrocarbons — 
that is, they contain only carbon and hydrogen. Hence their physical and 
chemical properties will resemble those of methane itself. (The molecular 
shapes are not well represented by the structures in the table because each of 
the carbon atoms in these molecules will have a tetrahedral arrangement of 
bonds connected to it. The three-dimensional shapes of such hydrocarbons 
are considered in more detail in Chapter 3.) The three derivatives of ammonia 
are called amines and share many of the properties of ammonia; for example, 
like ammonia, they have dipole moments and are weak bases. A different 
situation exists with the derivatives of water; each of them is representative 
of a class of compounds The structure CH 3 — OH is an alcohol while CH 3 — 
O— CH 3 is an ether. The reason that water is considered to give two classes 
of compounds on methyl substitution can be traced to the great importance 
of the hydroxyl (OH) group in chemistry. Alcohols, like water, contain a 



Table 1-1 Some simple derivatives of methane, ammonia, and water 



CH 4 


NH 3 


H 2 


CH 3 CH 3 


CH 3 -NH 2 


CH 3 -OH 


CH 3 -CH 2 -CH 3 


CH 3 -NH 
1 
CH 3 


CH 3 -0-CH 3 


CH3-CH-CH3 


CH 3 -N-CH 3 

1 




CH 3 


CH 3 




CH 3 
1 






CH 3 -C-CH 3 
1 






CH 3 







chap 1 introduction 10 

hydroxyl group whereas ethers do not. Hydroxyl groups have a great in- 
fluence on molecular properties (see next section), and the properties of 
alcohols (ROH) and ethers (R— O— R) are quite different. (The symbol R is 
usually used in organic chemistry for an alkyl group, a connected group of 
atoms formed by removing a hydrogen atom from a hydrocarbon ; a methyl 
group is one kind of R group.) 

The bond angles at oxygen in the two compounds CH 3 — O— H and 
CH 3 — O— CH 3 are somewhat greater than those found in water. This is 
expected because the CH 3 group is larger than hydrogen, and interference 
between the CH 3 groups in CH 3 — O— CH 3 is lessened by opening the 
C— O— C bond angle in the ether. A compromise angle is allowed in which 
the interference between the CH 3 groups is reduced at the expense of moving 
the pairs of electrons on oxygen to less favorable arrangements with respect 
to one another. A common description of the overall change is " relief of 
steric hindrance between the CH 3 groups by opening the C — O— C bond 
angle." 

A A A 

H H CH 3 H CH 3 CH 3 

bond angle i04.5° bond angle 106° bond angle 1 12° 



B. PHYSICAL PROPERTIES 

The four compounds methane, ammonia, water, and hydrogen fluoride 
have the physical constants shown in Table 1 -2. In each of these compounds 
the atoms are held together to form molecules by strong covalent bonds. The 
melting and boiling points are governed not by these powerful forces but 
rather by the weaker interactions that exist between molecules — intermolec- 
ular forces. Everything else being the same, the weaker such intermolecular 
interactions, the lower the temperature which will usually be required, first 
to break down the crystal lattice of the solid by melting, and then to separate 
the molecules to relatively large distances by boiling. 

What is the origin of these weak, secondary forces that exist between 
neutral molecules? We shall consider two here: van der Waals forces and 
hydrogen bonding. Van der Waals forces, sometimes called London forces, 
depend in an important way on the numbers of electrons in a molecule. This 
means that, in general, the bigger the molecule the greater will be the various 



Table t-Z Physical properties of methane, ammonia, water, and hydrogen 
fluoride 



CH, NH, H,0 HF 



boiling point 


-161.5° 


-33° 


100° 


20° 


melting point 


-183° 


-78° 


0° 


-84° 


solubility in water 


very low 


high 


00 


00 


solubility in CC1 4 


high 


very low 


very low 


very low 



sec 1.2 methane, ammonia, water, and hydrogen fluoride 11 
Table 1-3 Boiling and melting points of some methane derivatives 



CH 3 
I 
CH^ CH3 — CH3 CH3 — CH 2 — CH3 Co 3 — CH — CH3 CM 3 — C — CM 3 

CHa CM* 



boiling point -161.5° -88.6° -42.1° -10.2° 9.5° 

melting point -183° -172° -188° -145° -20° 



possible intermolecular attractions and the higher the melting and boiling 
points will tend to be. Boiling points tend to increase regularly within a series 
of compounds as the molecular weight increases. Melting points, however, 
usually show much less regularity. This is because the stability of a crystal 
lattice depends so much on molecular symmetry, which largely determines the 
ability of the molecules to pack well in the lattice. Thus, the five hydrocarbons 
shown in Table IT have the boiling and melting points shown in Table 1-3. 
In addition to experiencing van der Waals forces (dispersion forces), 
molecules containing certain groups are attracted to one another by hydrogen 
bonding. To be effective, hydrogen bonding requires the presence of an 

—OH, — N— H, or F— H group; in other words, a hydrogen atom joined to a 

I 
small electronegative atom. The covalent bonds to such hydrogen atoms are 

8e se 
strongly polarized toward the electronegative atom, for example, R— O— H, 
and the partially positive hydrogen will be attracted toward the partially 
negative oxygen atom in a neighboring molecule. In the liquid state a number 



/ 






H 


I 




ae 


he/ 




aeO- 


-H- 


--<> 




/ 




\ 




R 




R 



of molecules may be linked together this way at any given time. These liaisons 
are not permanent because thermal energies of the molecules are sufficient 
to cause these bonds to break very rapidly (usually within milliseconds or 
less). Such bonds are continually being formed and broken and this leads to 
the description of such temporary aggregates in a hydrogen-bonded liquid as 
" flickering clusters." 

By far the most important of the groups responsible for hydrogen bonding 
is the hydroxyl group, —OH. The strength of O— H • • • O hydrogen bonds 
may be as much as one-tenth that of an ordinary carbon-carbon covalent 
bond. 2 (See Section 2-4 on bond strengths.) The highest boiling point in the 



2 Recently, minute quantities of what is claimed to be a new form of water, sometimes 
called "polywater," have been isolated, in which very strong hydrogen bonds are believed 
to exist, indeed stable to temperatures above 400°. It is not yet known to what degree one 
should expect to find a multiplicity of bond strengths for hydrogen bonds to a specific 
molecule or indeed whether " polywater " is in fact even a compound of formula (H 2 0)„ . 



chap 1 introduction 12 
Table 1 -4 Boiling and melting points of some oxygen compounds 



H,0 CH,OH CH,OCH, 



boiling point 100° 65° -24° 

melting point 0° -89° -139° 



series CH 4 , NH 3 , H 2 0, HF belongs to water and the lowest to methane, in 
which hydrogen bonding is completely absent (Table 1-2). 

The three oxygen compounds shown in Table 1-1 have melting and boiling 
points as shown in Table 1 -4. 

The trends are exactly opposite to those expected on the basis of molecular 
weight alone and are the result of having two hydrogens bonded to oxygen in 
each molecule of water, one in the alcohol, CH 3 OH, and none in the ether, 
CH 3 OCH 3 . 

The hydroxyl group also has an important influence on solubility character- 
istics. The alcohol CH 3 OH is completely miscible with water because the two 
kinds of molecule can form hydrogen bonds to one another. On the other 
hand, the ether CH 3 OCH 3 is only partly soluble in water. Its oxygen atom 
can interact with the protons of water but it has no OH protons itself to 
continue the operation. Hydrocarbons have extremely low solubilities in 
water. Hydrocarbon molecules would tend to interfere with the hydrogen 
bonding between water molecules and could offer in exchange only the much 
weaker van der Waals forces. 

The nitrogen compounds shown in Table IT have boiling and melting 
points as shown in Table 1 -5. There is not a great deal of difference between 
the values for the three amines. Hydrogen bonding N— H---N is not as 
effective as O— H* • • O and the reduction in hydrogen bonding in going from 
CH 3 — NH 2 to CH 3 — N— CH 3 is roughly compensated by the increase in 

CH 3 

van der Waals forces caused by increasing molecular size. 



C. ACIDITY AND BASICITY 

The acidity of the four compounds methane, ammonia, water, and hydrogen 
fluoride increases regularly as the central atom becomes more electronegative. 



Table 1-5 Boiling and melting points of some nitrogen compounds 





NH 3 


CH 3 NH 2 


CH,— NH 

1 
CH 3 


CH 3 - 


-N— CH 3 

1 
CH 3 


boiling point 
melting point 


-33° 
-78° 


-6.5° 
-93° 


7° 
-96° 




4° 
-124° 



summary 13 

Thus the acid dissociation constants for these compounds in water solution 
are: 

CH 4 NH 3 H 2 HF 

A"ha,25° ~1(T 55 ~10~ 35 . 5.5 x 1(T 15 3.5 x 1(T 4 

The ionization constant used here for water is the customary value of 10" 14 
divided by the concentration of water in pure water (55 M). The symbol 
K HA denotes the equilibrium constant for dissociation of a neutral acid HA, 
that is, HA?±H® + A e and K HA = [H e ][A e ]/[HA]. The values refer to 
water solution whether actually measurable in water or not and the symbol 
H® represents the oxonium ion H 3 0®. 

Methane, like most other hydrocarbons, has a negligible acidity in water. 
Amines resemble ammonia in being very feeble acids. Alcohols are somewhat 
stronger and have acidities similar to that of water. 

The basicities of these four compounds follow a different pattern which is 
not simply the reverse of that for acidity : 

CH 4 NH 3 H 2 HF 

K B 25° <1(T 30 1.8 x 10~ 5 5.5 x 1(T 15 ~1(T 25 

The symbol K B denotes the equilibrium constant for ionization of a neutral 
base B, that is, B + H 2 0?±BH® + OH® and K B = [BH®][OH®]/[B]. Base 
strengths are sometimes taken to be indicated by the acid strengths of the 
corresponding conjugate acids. When this is done the symbol K BH o should be 
used to denote the process being referred to; that is, K BH ® = [H®][B]/[BH®] 
represents the acid dissociation BH® ?± H® + B. Ordinary basic ionization 
constants, K B , will be used in this book. 

The increase in basicity from HF to H 2 to NH 3 is readily understandable 
in terms of the decreasing electronegativity of the central atom along the 
series. Why then is the basicity of methane so low ? The reason is that this 
molecule has no unshared pairs of electrons available for bonding to a proton 
as do ammonia and the other compounds with which we have compared it. 

H 3 N: + H 2 . NH 4 e + OH e 

If methane is to accept a proton to form the ion CH 5 ® the carbon atom must 
hold five hydrogen atoms with four pairs of electrons. (There is evidence that 
CH 5 ® can be generated and detected in the gas phase in a mass spectro- 
meter. It may also be a transient intermediate in solutions of methane in the 
so-called " super acids." Examples of the latter are mixtures of FS0 3 H and 
SbF 5 ; their protonating power far exceeds that of concentrated sulfuric acid.) 



summary 

Carbon is unique among the elements: it is able to form an enormous 
number of compounds by bonding to itself and to the atoms of other elements, 
principally hydrogen, oxygen, nitrogen, sulfur, and the halogens. Such 



chap 1 introduction 14 

bonding is almost always covalent, with each carbon atom having four bonds, 
each bond resulting from a pair of electrons in an orbital which encloses both 
the bonded nuclei. Repulsions between the four electron pairs in CH 4 , NH 3 , 
and H 2 determine the shapes of these molecules; CH 4 is tetrahedral with 
bond angles of 109.5° and NH 3 and H 2 have slightly smaller bond angles. 
Of these compounds only methane, CH 4 , because of its symmetry, has no 
dipole moment. 

The hydrogen atoms in CH 4 , NH 3 , and H 2 can be replaced by alkyl 
groups such as methyl, CH 3 — . Compounds formed from CH 4 this way are 
hydrocarbons like CH 4 itself. Those from NH 3 are called amines, CH 3 NH 2 
or CH3NHCH3 . Those from H 2 are alcohols if only one hydrogen is 
replaced, CH 3 OH; and ethers if both hydrogens are replaced, CH 3 OCH 3 . 

The physical properties of such compounds are determined chiefly by 
iVrtermolecular forces, van der Waals forces and hydrogen bonds, which are 
normally much weaker than those involved in covalent bond formation. In 
general, the larger the molecule, the greater the van der Waals forces and the 
higher the boiling point. The presence of one or more hydroxyl groups (or 
other good hydrogen-bonding groups) will raise the boiling point consider- 
ably and will tend to make the compound soluble in water. 

Acidity increases in the order CH 4 < NH 3 <H 2 < HF; basicity in the 
order CH 4 < HF <H 2 < NH 3 . 



exercises 

1 • 1 Write Lewis structures for each of the following compounds using dots for the 
electrons. Mark any atoms which are not neutral with charges of the ap- 
propriate sign. 

a. ammonia e. ozone (Z0-O-0 = 120°) 

b. ammonium bromide / hydroxylamine, H 2 NOH 

c. carbon dioxide g. hydrogen cyanide 

d. hydrogen peroxide h. boron trifluoride 

1-2 Tetramethyllead, Pb(CH 3 )4, is a volatile liquid, bp 106°, while lead fluoride, 
PbF 2 , is a high-melting solid, mp 824°. What kinds of bonding forces are 
present in the two compounds ? 

1-3 Which of the following substances are expected to possess a dipole moment? 
Why? 

(CH 3 ) 3 N, 3 , C0 2) BF 3 , CH 2 F 2 , CF 4 , CH3OCH3, CH3CH3 

1-4 Do you expect the compound hydrazine, NH 2 NH 2 , to be more or less basic 
than ammonia? Explain your answer. 

1 • 5 Which of the compounds in the following list are expected to be more soluble 
in water than in carbon tetrachloride? 

D 2 0, CH 3 NH 2 , CH 3 CH 3 , CH 3 C0 2 e Na e , HC1, CCU 



exercises IS 

1-6 Why is hydrogen peroxide a stronger acid than water? 

1-7 Arrange the following compounds in the order of increasing acidity in water 
solution. 

NH 3 -NH 3 S0 4 2e , CH 3 OH, CH 4 , NH 4 ®Cl e , CH3OCH3 

1-8 The term autoprotolysis means self- ionization by proton transfer, that is, 
2 H 2 *± H 3 O ffl + OH e . Write autoprotolysis reactions for the following 
liquids: NH 3 , CH 3 OH, H 2 S0 4 , HOOH, CH 3 NH 2 . 

1-9 Addition of 17.9 g of water to 100 g of pure liquid perchloric acid, HC10 4 , 
produces a crystalline solid. What is its formula ? 

1-10 There is evidence to suggest that the form of the solvated proton in water 
solution is better represented by the formula H 9 O 4 than H 3 O e . Draw a 
structural formula for H 9 4 ® and identify the kinds of bonds that might 
hold it together. 

1-11 On pp. 8-9 it is suggested that the repulsions between the bonded pairs of 
electrons in water will be less than between the nonbonded pairs, thus 
making the H— O— H angle less than 109.5°. Explain why this should be so. 



•*W*-.sft:vss 



tofaStss y.f/i\ 



iwrif'SS ©ft-.-il u^-'..-W *&£«0 %■*:.<&■ 

r . 7 'V -' , 

rj*-Jft?a? ^b^Sr^T. li^-^ ;&! " 






B5«I f a»iS;~ S'SSE i«K 











f-i?fi *T^i> U*a^ '*:*3'*' S V-At" r'sV'Vji I 



y , :- ! l;-'X' 




Vi'-i'" i.'.'iVi* - . i*:%V> 


**v: &*.*K ; - 


ViVr-i. rs.L : t''iV : -V ■'-'*•' <''* 


.;rU':. l *.'i- : -V-..-,C 




vir^--** -S.-.vw .v.*n*:7* ! 


4. . , > * 


• 


.'iVi f .?>V,"" '-'•* ; '!"if-' 




'..ii' *vrf 


S3n $*• {** «i? 




•-'••# 


•$#?»•■■" -a'j'-.S, 





;Si» sit siii Hasan mmi if 

s^«»^^^^«^^;sx^*..,«,,. v ...,:..,Js i^:$«**<*$? fep*,^ *>. -life Hit 



"•;?" : K-*-'i* *Sa Sfessiag?®? %> ^ia* iE iff §i 
i'":.3f^< ■;^^V f :>: *jg0^ $& 

-S'-i i .:-V5'^!' rt s "??, ©E 'ifSrsftis* :\Si.f S^ •- s il5 lW v S?s 
•'■->4.:-. .A ..•.*;■{ viST*ag©**« r*fe t*i*-:« -ftgi SS^fe 

'■ ',1 > I 

. K -. ?/ •..•-.*» .is^j riW':?5*vSi«>*i? -^sSj. r . ^«ste; &»sas 

^^'••••wtn^-^^i^rt ^■'i-'v.^^^^Jzkj!; i£^!S J 3>l3SV>^&?. i^i(V^' ; -v 



fe*>'^ 



K&:igl^ 






'^i* s '*''- V' • : :-'"-*V^* * ; .^V'- S'- : iV-. i 'f..':' :1 'vyv *'V-.'; "•*•• il^'-V-'- Vi'-V.v *;-• ^'.Oi- ^eV*"!'-''"''^": •-•ii T ^ f T*'^ y *''W; : '^"*&' J^-¥i--\?^ 



'& : <?& fe^S» « ; 5'i %V** ^fi'S i 



psa 



B :*;.HfeSSiSJ 



5&K=3 



.,,.,. ..= .,-. v,-. ••=--,. «-^v .-..-. ^%^W. TOW 



P^^^-v- ;*. >v*3\v» ■:*&(.:&$%$ ^v&-^^« ^0?.^vpf|; *^ 

a isi\i--v='if;»'il5 i -ra ; 2yy<fe-i?a ^"^^*- l >-VAxJgi A«yS#feSgj«i« tflfefe' Sfcte ^:Srt 

;^a I^Ia^ ^?Sil?S iS^M.Wi# ^lii;€ iife> 't^ liS 









K'sgss±a% s«g 








chap 2 the C x and C 2 hydrocarbons 19 

Hydrocarbons are compounds that contain only carbon and hydrogen. 
We shall consider in this chapter the four simplest known hydrocarbons — 
those with the lowest molecular weights — and we shall see that they represent 
three classes of compounds: the alkanes, in which each carbon atom has four 
single bonds ; the alkenes, in which two carbon atoms are joined by a double 
bond (two electron pairs); and the alkynes, in which two carbon atoms are 
joined by a triple bond (three electron pairs). We shall also see that these 
classes of compounds are physically similar but chemically rather different. 

There are four stable hydrocarbons of molecular weight 30 or less. They are 
all gases at room temperature, and analyses for carbon and hydrogen content 
coupled with determinations of their molecular weights show them to have 
the formulas CH 4 , C 2 H 6 , C 2 H 4 , and C 2 H 2 . The first of these is methane, 
CH 4 , whose physical properties and molecular shape were discussed in Chap- 
ter 1. The other three are all C 2 compounds and are called, respectively, 
ethane, ethene, and ethyne (ethyne rhymes with brine). These are the syste- 
matic names approved by the International Union of Pure and Applied 
Chemistry, IUPAC. 1 However, ethene is often called ethylene and ethyne 
called acetylene. It is to be hoped that both of these older names will pass out 
of use in time. 

If the carbon atoms in each of the three C 2 compounds are tetravalent, 
then there is only one possible way to bond the atoms together in each case : 

H H 

H H || 

ethane H:C:CH or H— C— C-H 

H " It li 

ethene H. .H \ / 

(ethylene) H : - C::C .-' H or / C = C \ 

H H 



ethyne 
(acetylene) H:C:::CH or H-C=C-H 

The tendency of carbon to form bonds at the tetrahedral angle results in 
compounds such as methane and ethane being nonplanar. The two-dimen- 
sional representation of ethane, above, is thus misleading and it is just as 
informative (and quicker) to write the formula as CH 3 — CH 3 . (Or, indeed, 
as C 2 H 6 , since there is only one stable compound known with this formula. 
We shall see that with some C 3 hydrocarbons and with all hydrocarbons 
having four or more carbons, some indication of structure is necessary because 
a designation such as C 4 H 8 is ambiguous, there being five known compounds 
with this formula.) 

1 IUPAC, with headquarters at Zurich, Switzerland, is an international organization 
concerned with creating worldwide standards in nomenclature, analytical procedures, 
purity, atomic weights, and so on. It is governed by a Congress of delegates from many 
countries, the number of delegates depending partly on a country's financial resources and 
partly on its scientific maturity. The following countries have the maximum number of 
delegates (six): Australia, Belgium, Canada, Denmark, France, Germany, Italy, Japan, 
Netherlands, Sweden, Switzerland, United Kingdom, United States, U.S.S.R. IUPAC was 
founded in 1918 at the famous Coq d'Or restaurant in London. 



chap 2 the C x and C 2 hydrocarbons 20 



# 



Figure 2- 1 Ball-and-stick model of CH 4 . 

Because of the importance of molecular structure in organic chemistry, we 
shall consider the three-dimensional shapes of these compounds in the next 
section. 

2-1 molecular shape of CH 4 , C 2 H 6 , C 2 H 4 , 
and C 2 H 2 

You can illustrate the shape of a tetrahedral molecule such as methane with 
ball-and-stick models (Figure 2-1). 

With ethene and ethyne, the model's carbon-to-carbon bonds are con- 
structed from stiff metal springs or flexible or curved plastic connectors be- 
cause more than one bond exists between the carbon atoms (Figure 2-2). 

These simple mechanical models are surprisingly good for predicting the 
shapes of molecules and, indeed, their reactivity. Ethene is known from 
spectroscopic measurements to be planar, and this is the shape the model 
naturally takes. The electronic analogy here is that the orbitals for each pair 
of electrons extend as far away from one another as possible. Ethyne, likewise, 
is known to be linear. The strain involved in making " bent bonds " for these 
models is reflected in a higher degree of chemical reactivity for these com- 
pounds than for ethane. 

Figure 2-2 Ball-and-stick models of ethene and ethyne. 






sec 2.2 rotational conformations of ethane 21 






Figure 2-3 Two rotational conformations of ethane. 



The arrangement of the linkages in the ethene model suggests that one CH 2 
group cannot twist with respect to the other CH 2 group without gross dis- 
tortion from the favored geometry. We shall see that this conclusion, too, is 
borne out by chemical evidence (Section 2-6B). By contrast, the model of the 
saturated compound, ethane, suggests that free rotation should be possible 
about the single bond joining the two carbon atoms if the sticks representing 
the bonds are allowed to rotate in the holes of the balls representing the 
atoms. Such rotation is considered in more detail in the next section. 



2-2 rotational conformations of ethane 

In organic chemistry, the word structure has a specific meaning; It designates 
the order in which the atoms are joined to each other. A structure does not 
necessarily specify the exact shape of a molecule because rotation about single 
bonds could lead, even for a molecule as simple as ethane, to an infinite 
number of different arrangements of the atoms in space. These are called 
conformations and depend on the angular relationship between the hydrogens 
on each carbon. Two extreme arrangements are shown in Figure 2-3. 

In end-on views of the models, the eclipsed conformation is seen to have the 
hydrogens on the forward carbon directly in front of those on the back carbon. 
The staggered conformation has each of the hydrogens on the forward carbon 
set between each of the hydrogens on the back carbon. It has not been possible 
to obtain separate samples of ethane which correspond to these or intermediate 
arrangements because actual ethane molecules appear to have essentially 
"free rotation" about the single bond joining the carbons. 

Free, or at least rapid, rotation is possible around all single bonds, except 
under special circumstances, as when the groups attached are so large that 
they cannot pass by one another, or when the attached groups are connected 
together by chemical bonds (e.g., in ring compounds). For ethane and its 
derivatives, the staggered conformation is always more stable than the eclip- 



chap 2 the C 1 and C 2 hydrocarbons 22 



H 


H 






jj 


X 


H 


HH 


H-y^H 


Hi rN 


H^ 


^ H 


r 


\ 


H^H 


H^H 


A 


^ 


"J< 


y% 


H 

staggered eclipsed 
"sawhorse" 


staggered eclipsed 
"Newman" 



Figure 2-4 Conventions for showing the staggered and eclipsed conformations 
of ethane. In " sawhorse " drawings the lower left-hand carbon is always 
taken to be towards the front. In " Newman " drawings the view is along the 
C — C bond axis with the most exposed bonds being towards the front. 



sed conformation because in the staggered conformation the atoms are as 
far away from one another as possible and offer the least interaction. 

Many problems in organic chemistry require consideration of structures in 
three dimensions, and it is very helpful to be able to use ball-and-stick models 
for visualizing the relative positions of the atoms in space. Unfortunately, 
we are very often forced to communicate three-dimensional concepts with 
drawings in two dimensions, and not all of us are equally gifted in making 
or visualizing such drawings. Obviously, communication by means of draw- 
ings such as the ones shown in Figure 2-3 would be impractically difficult and 
time consuming — some form of abbreviation is necessary. 

Two styles of abbreviating the eclipsed and staggered conformations of 
ethane are shown in Figure 2-4. Of these, we strongly favor the " sawhorse " 
convention because, although it is perhaps the hardest to visualize and the 
hardest to master, it is the only three-dimensional convention which is suitable 
for complex compounds, particularly natural products. With the sawhorse 
drawings, we always consider that we are viewing the molecule slightly from 
above and from the right, just as we have shown in Figure 2-4. 



2-3 space-filling models 



Ball-and-stick models of molecules are very useful for visualizing the relative 
positions of the atoms in space but are unsatisfactory whenever we also want 
to show how large the atoms are. Actually, atomic radii are so large relative 
to the lengths of chemical bonds that when a model of a molecule such as 
chloromethane is constructed with atomic radii and bond lengths, both to 
scale, the bonds connecting the atoms are not clearly evident. Nonetheless, 
this type of "space-filling" model, made with truncated balls held together 
with snap fasteners, is widely used to determine the possible closeness of 
approach of groups to each other and the degree of crowding of atoms in 
various arrangements (see Figure 2-5). 

A defect of both the ball-and-stick and space-filling models is their motion- 
less character. The atoms in molecules are in constant motion, even at absolute 



sec 2.4 combustion 23 

zero, and the frequencies of these vibrations give valuable information about 
molecular structure and shape. This subject is considered in greater detail in 
the section on infrared spectroscopy (Section 7-4). 

chemical reactions of the d and C 2 hydrocarbons 

Two of the four simple hydrocarbons we have been considering are saturated 
(contain only single bonds) and two are unsaturated (contain multiple bonds). 
All four are rather similar physically, being low-boiling, colorless gases that 
are insoluble in water, Chemically, however, they are rather different, the 
unsaturated compounds being much the more reactive. The three kinds of 
reactions we shall consider in the following sections are combustion (shared 
by all hydrocarbons), substitution reactions (more important for saturated 
compounds), and addition reactions (confined to the unsaturated compounds). 

2-4 combustion 

The rapid reaction of a chemical substance with oxygen to give an oxide, 
usually carbon dioxide, is called combustion. The burning of a candle, the 
explosion of a gasoline-air mixture in the cylinder of an automobile engine, 
and the oxidation of glucose in a living cell are all examples bf this process. 
In all of these cases, the result is liberation of energy. 

Water and carbon dioxide, the products of complete combustion of organic 
compounds, are very stable substances, relative to oxygen and hydrocarbons. 
This means that large amounts of energy are given out when combustion 
occurs. Most of the energy of combustion shows up as heat, and the heat 
liberated in a reaction occurring at constant pressure is called the enthalpy 
change, AH, or simply heat of reaction. By convention, AH is given a negative 
sign when heat is evolved (exothermic reaction) and a positive sign when heat 
is absorbed (endothermic reaction). Some examples are given below, with the 
state of the reactants and products being indicated by subscripts (g) for gas 
and (s) for solid. 

For each of these examples, AH can be visualized as the total heat given off 
when a mixture of gaseous hydrocarbon and excess oxygen at 1 atm pressure 
is exploded in a bomb at 25°, the contents allowed to expand or contract by 

Figure 2-5 Space-filling models of organic compounds. 




#*^>Ss^'' 



methane 




ethane (staggered conformation) 




-M^ 



ethane 



chap 2 the C x and C 2 hydrocarbons 24 

means of a piston to maintain the pressure at 1 atm, and allowed to cool to 

25°. 

CH 4 (g) + 2 2 (g) -> C0 2 (g) + 2 H 2 0(g) AH= - 1 92 kcal per mole of 

methane (—11.95 kcal per 
gram) 

CH 3 — CH 3 (g) + f 2 (g)^2 CO z (g) + 3 H 2 0(g) MU= -341.3 kcal per mole of 

ethane (—11.35 kcal per gram) 

H— C=C— H(g)+ 4 2 (g)->2 C0 2 (g) + H 2 0(g) &H= -300.1 kcal per mole of 

ethyne (— 1 1 . 50 kcal per gram) 

C s H ls (g)+ ^ 5 -0 2 (g)^ 8 C0 2 (g) + 9 H 2 0(g) AH= -1222 kcal per mole of 

octane (— 10.71 kcal per gram) 

C 6 Hj 2 6 (s) + 6 2 (g)-+6 C0 2 (g) + 6 H 2 0(g) AH= -610 kcal per mole of glucose 

(—3.39 kcal per gram) 

You can see that the amount of heat liberated per gram of fuel is not greatly 
different in the case of the four hydrocarbons, but is much lower for the com- 
pound C 6 H 12 6 (glucose), which is already in a partly oxidized state. In the 
next section, we shall consider how you can estimate heats of reaction, with 
particular reference to combustion. 



A. ESTIMATION OF HEAT OF COMBUSTION OF METHANE 

The experimental value for the heat of combustion of methane obtained as 
described above does not depend on speed of the reaction. Slow oxidation of 
methane over many years would liberate as much heat as that obtained in an 
explosion, provided the reaction were complete in both cases, and the initial 
and final temperatures and pressures are the same. At 25°, combustion of each 
mole of methane to carbon dioxide and water vapor produces 192 kcal of heat. 

We can estimate the heat of this and many other reactions by making use 
of the bond energies given in Table 2- 1 . Bond energies for diatomic molecules 
represent the energy required to dissociate completely the gaseous substances 
to gaseous atoms at 25° or, alternatively, the heat evolved when the bonds are 
formed from such atoms. For polyatomic molecules the bond energies are 
average values. They are selected to work with a variety of molecules and re- 
flect the fact that the bond energy of any particular bond is likely to be 
influenced to some extent by other groups in the molecule. 

It turns out that what is called conjugation (alternation of double and single 
bonds) can have a relatively large effect on bond strengths. We will see in 
Chapter 6 that this effect normally operates to increase bond energies ; that is, 
the bonds are harder to break and the molecule is made more stable. However, 
the effects of conjugation are so special that they are not normally averaged 
into bond energies, but are treated separately instead. 

To calculate Af/for the combustion of methane, first we calculate the energy 
to break the four C— H bonds as follows (using the average value of 99 kcal 
for the energy of a C— H bond): 

H 
H:CH(s) ► -C- (g) + 4 H-(g) A«=+4x99kcal 

H ' = + 396 kcal 



sec 2.4 combustion 25 
Table 2-1 Bond energies (kcal/mole at 25°)" 







diatomic 


molecules 






H-H 


104.2 


F-F 


36.6 


H-F 


134.5 


o=o 


119.1 


CI- CI 


58.0 


H-Cl 


103.2 


N^N 


225.8 


Br— Br 


46.1 


H-Br 


87.5 


c=o 


255.8 


I-I 


36.1 


H-I 


71.4 




bonds in polyat 


omic molycules 6 




C— H 


99 


c-c 


83 


C-F 


116 


N-H 


93 


C=C 


146 


C-Cl 


81 


O-H 


111 


C=C 


200 


C-Br 


68 


S— H 


83 


C-N 


73 


C-I 


51 


P— H 


76 


C=N 


147 


C-S 


65 


N— N 


39 


C-N 


213 


C=S 


128 


N=N 


100 


c-o 


86 


N-F 


65 


o-o 


35 


c=o c 


192 


N— CI 


46 


s-s 


54 


c=(y 


176 


O-F 


45. 


N— O 


53 


c=o e 


179 


O-Cl 


52 


N=0 


145 






O-Br 


48 


intermolecular forces 


hydrogen 


bonds 


3-10 









" The bond energies in this table are derived from those of T. C. Cottrell, The Strengths of 
Chemical Bonds, 2nd Ed., Butterworths, London, 1958, and L. Pauling, The Nature of the Chem- 
ical Bond, 3rd Ed., Cornell Univ. Press, Ithaca, N.Y., 1960. 

b Average values. 

c For carbon dioxide. 

d Aldehydes. 

e Ketones. 

Then 119 kcal is used for the energy required to cleave a molecule of oxygen 
(rounded off from the exact value of 119.1): 

2 0,(g) -♦ 4-6- (g) A//= +2 x 119 kcal 

= + 238 kcal 

Then we make bonds, using 192 kcal for each C=0 bond in carbon dioxide. 

■6 (g) + 2-6-(g) :6::C::6:(g) AH = - 2 x 192 kcal 

= -384 kcal 

We use 111 kcal for each of the H— O bonds of water: 

2-6-(g) + 4 H- (g) > 2 H:6:H(g) A//= -4 x 111 kcal 

= - 444 kcal 

The net of these AH changes is 396 + 238 - 384 - 444 = - 194 kcal, which 
is reasonably close to the value of 191.8 kcal for the heat of combustion of 
methane determined experimentally. 

The same type of procedure can be used to estimate AH values for many 
other kinds of reactions of organic compounds in the vapor phase at 25°. 



chap 2 the C x and C 2 hydrocarbons 26 

Moreover, if appropriate heats of vaporization or solution are available, it is 
straightforward to compute AH for liquid, solid, or dissolved substances. 

The steps shown above are not intended to depict the actual mechanism 
of methane combustion. The overall heat of reaction is independent of the way 
that combustion occurs and so the above calculations are just as reliable as 
(and more convenient than) those based on the actual reaction path. Some of 
the general questions posed by the reaction mechanism are taken up in Section 
2-5B. 



2-5 substitution reactions oj saturated hydrocarbons 

Of the four simple hydrocarbons we are considering in this chapter, only 
ethene and ethyne are unsaturated, meaning they have a multiple bond to 
which reagents may add. The other two compounds, methane and ethane, 
have their atoms joined together by the minimum number of electrons and 
can react only by substitution — replacement of a hydrogen by some other 
atom or group. 

There are only a few reagents which are able to effect the substitution of a 
hydrogen atom in a saturated hydrocarbon (an aikane). The most important 
of these are easily the halogens, and the mechanism and energetics of halogen 
substitution will be discussed in detail later. (Although the hydrogen atoms in 
alkenes such as ethene and alkynes such as ethyne are also subject to substi- 
tution, these reactions under normal conditions tend to be much slower than 
addition to the multiple bond and are therefore usually not important when 
compared to addition.) 

A complete description of a chemical reaction would include the structures 
of the reactants and products, the position of equilibrium of the reaction, its 
rate, and its mechanism. These four characteristics fall nicely into two groups. 
The equilibrium constant for a reaction depends only on the energies of the 
reactants and products, not on the rate of reaction nor on the mechanism. The 
rate of the reaction, on the other hand, is intimately related to the reaction 
mechanism and, in particular, to the energy of the least stable state along the 
reaction path. The subjects of equilibrium constants and reaction rates are 
treated in the next two sections. 



A. EQUILIBRIUM CONSTANTS 

In Section 2-4A we considered bond energies and showed how heats of reac- 
tion could be calculated. Reactions which give out large amounts of heat 
(highly exothermic processes) usually proceed to completion. Consequently 
it is reasonable to ask if the equilibrium constant, K, for a reaction is deter- 
mined only by the heat of reaction, AH. The study of thermodynamics tells 
us that the answer to this question is no. The equilibrium constant is, in fact, 
a function of the quantity free energy (AG), which is made up of AH and 
a second quantity called entropy (AS). These relations are 

AG = -RTlnK 
AG = AH -T AS 



sec 2.S substitution reactions of saturated hydrocarbons 27 

where 

AG = Free energy change for the reaction 

R = The gas constant (1.986 cal/deg mole) 

T = Temperature in degrees Kelvin 

K = Equilibrium constant 
AH = Heat of reaction 
AS = Entropy of reaction 

The heat of reaction term, AH, is readily understood but the meaning of 
the entropy term, AS, is more elusive. It is related to the difference in the num- 
bers of vibrational, rotational, and translational states available to reactants 
and products (see Sections 7-3 and 7-4). As we have seen, molecules are not 
lifeless objects but are in constant motion, each undergoing vibrational, 
rotational, and translational motions. Thes6 states of motion are quantized — 
that is, they can have certain energies only. The more of these states or degrees 
of freedom available to a molecule, the higher its entropy and the more favor- 
able the equilibrium constant for its formation. Thus, a positive entropy 
change in a reaction tends to make the free energy change more negative and 
increase the equilibrium constant, hence moving the reaction toward comple- 
tion. 

In simple terms, a negative entropy change (a AS" that is unfavorable for the 
reaction as written) means that the freedom of the atoms in the products 
(including the environment) is restricted more than in the reactants. A posi- 
tive entropy change (favorable for the reaction as written) means a greater 
freedom in the products. 

In practice, reactions which are fairly exothermic ( — AH> 15 kcal/mole) 
almost always proceed far to the right ; that is, K is large. An unfavorable 
entropy term will seldom overcome such a AH value at ordinary temperatures 
because a AG that is negative by only a few kilocalories per mole will still 
have a large K, This follows from the logarithmic relation between AG and K. 

The thermodynamic values for the chlorination of methane are 

CH 4 + C1 2 — > CH3GI + HCI 

AH= -27 kcal/mole 

(measurable experimentally or calculable from the data in Table 2-1); 

AS — — 6 cal/deg mole 

(estimated from the spectroscopic properties of reactants and products) ; 

AG= -25 kcal/mole 

(calculated from above values of AH and AS and the equation AG = AH — T 
AS); 

(calculated from above value of AG and the equation AG = — RT In K). 

In some cases, you can experimentally check an equilibrium constant 
calculated as above by measuring the concentrations of reactants and pro- 



chap 2 the Cj and C 2 hydrocarbons 28 

ducts when the system has come to equilibrium. Here, however, K is so large 
that no trace of the reactants can be detected at equilibrium, a situation often 
encountered in organic chemistry. 

Suppose bond energy calculations for a certain reaction indicate that the 
equilibrium strongly favors the desired products. Can we be assured that the 
reaction is a practical one to perform in the laboratory ? Unfortunately, no, 
because first, side reactions may occur (other reactions which also have 
favorable equilibrium constants); and second, the rate of the desired reaction 
may be far too low for the reaction to be a practical one. 

In the chlorination of methane, whose equilibrium constant we have seen 
overwhelmingly favors the products, the first of these two matters of concern 
is whether the substitution process may proceed further to give dichloro- 
methane, CH 2 C1 2 . 

CHjCl + Cl 2 ► CH 2 C1 2 + HC1 

In fact, given sufficient chlorine, complete substitution may occur to give 
tetrachloromethane (carbon tetrachloride), CC1 4 . Indeed, if the rate of chlori- 
nation of chloromethane greatly exceeds that of the first step, methane 
chlorination, there will be only traces of the mono substituted product in the 
mixture at any time. Using an excess of methane will help encourage mono- 
substitution only if the rates of the first two chlorination steps are comparable. 

With compounds and reagents that are more complex than methane and 
chlorine, you can imagine side reactions taking other forms. When devising 
synthetic schemes you must always consider possible side reactions that may 
make the proposed route an impractical one. 

The second question about the chlorination of methane that is left un- 
answered by the calculation of the equilibrium constant is whether or not the 
reaction will proceed at a reasonable rate. The subject of reaction rates is bound 
up intimately with the question of reaction mechanism and this subject is ex- 
plored in the next section. 

B. REACTION RATES AND MECHANISM 

Despite the enormously favorable equilibrium constant for the formation of 
chloromethane and hydrogen chloride from methane and chlorine, this 
reaction does not occur at a measurable rate at room temperature in the dark. 
An explosive reaction may occur, however, if such a mixture is irradiated with 
strong violet or ultraviolet light. Evidently, light makes possible a very 
effective reaction path by which chlorine may react with methane. 

Any kind of a theoretical prediction or rationalization of the rate of this or 
other reactions must inevitably take into account the details of how the 
reactants are converted to the products — in other words, the reaction mech- 
anism. One possible path for methane to react with chlorine would have a 
chlorine molecule collide with a methane molecule in such a way that hydro- 
gen chloride and chloromethane are formed directly (see Figure 2-6). The 
failure of methane to react with chlorine in the dark at moderate temperatures 
is strong evidence against this path, and indeed four-center reactions of this 
type are rather rare. 



sec 2.5 substitution reactions of saturated hydrocarbons 29 




* 







Figure 2*6 Possible four-center collision of chlorine with methane, as 
visualized with ball-and-stick models. 



If concerted four-center mechanisms for formation of chloromethane and 
hydrogen chloride from chlorine and methane are discarded, the remaining 
possibilities are all stepwise mechanisms. A slow stepwise reaction is dynami- 
cally analogous to the flow of sand through a succession of funnels with 
different stem diameters. The funnel with the smallest stem will be the most 
important bottleneck, and if its stem diameter is much smaller than the others, 
it alone will determine the flow rate. Generally, a multistep chemical reaction 
will have a slow rate-determining step (analogous to the funnel with the small 
stem) and other, relatively fast steps which may occur either before or after 
the slow step. The prediction of the rate of a reaction proceeding by a step- 
wise mechanism then involves, as the central problem, a decision as to which 
step is rate determining and an analysis of the factors which determine the 
rate of that step. 

A possible set of steps for the chlorination of methane follows : 

slow 

(1) Cl 2 > 2:CI- 

slow 

(2) CH 4 ► CH 3 - + H- 

fist 

(3) :CI- + CH 3 - ^—> CH 3 C1 

fast 

(4) :C1- + H- ► HCI 



chap 2 the C x and C 2 hydrocarbons 30 

Reactions (1) and (2) involve dissociation of chlorine into chlorine atoms, 
and the breaking of a C— H bond of methane to give a methyl radical and a 
hydrogen atom. The methyl radical, like chlorine and hydrogen atoms, has 
one odd electron not involved in bond formation. Atoms and free radicals 
are usually highly reactive, so that formation of chloromethane and 
hydrogen chloride should proceed readily by (3) and (4). The crux then will 
be whether steps (1) and (2) are reasonable under the reaction conditions. 

Our plan in evaluating the reasonableness of these steps is to determine how 
much energy is required to break the bonds. This will be helpful because, in 
the absence of some external stimulus, only collisions due to the usual thermal 
motions of the molecules can provide the energy needed to break the bonds. 
Below 100°C, it is very rare indeed that thermal agitation alone can supply 
sufficient energy to break any significant number of bonds stronger than 30 
to 35 kcal/mole. Therefore, we can discard as unreasonable any step, such as 
the dissociation reactions (1) and (2), if the A//'s for breaking the bonds are 
greater than 30 to 35 kcal. 

In most reactions, new bonds form as old bonds break and it is usually in- 
correct to consider bond strengths alone in evaluating reaction rates. (The 
appropriate parameters, the heat of activation, AH 1 , and the entropy of 
activation, AS*, are discussed in Section 8-9.) However, the above rule of 
thumb of 30 to 35 kcal is a useful one for thermal dissociation reactions such 
as (1) and (2), and we can discard these as unreasonable if their heats of 
dissociation are greater than this amount. 

For reaction (1) we can reach a decision on the basis of the CI— CI bond 
energy from Table 2-1, which is 58.0 kcal and clearly too large to lead to 
bond breaking as the result of thermal agitation at or below 100°. The C— H 
bonds of methane are also too strong to break at 100° or less. 

The promotion of the chlorination reaction by light must be due to light 
being absorbed by one or the other of the reacting molecules to produce a 
highly reactive species. Since a CI— CI bond is much weaker than a C— H 
bond, it is reasonable to suppose that the former is split by light to give two 
chlorine atoms. We shall see in Section 7-3 that the energy which can be 
supplied by ultraviolet light is high enough to do this ; photolytic rupture of 
the more stable C— H bonds requires radiation with much higher energy. It 
should now be clear why a mixture of methane and chlorine does not react in 
the dark at moderate temperatures. 

Cl 2 -^- 2:C1- 

Once produced, a chlorine atom can remove a hydrogen atom from a meth- 
ane molecule and form a methyl radical and a hydrogen chloride molecule 
(as will be seen from Table 2T, the strengths of C— H and CI— H bonds are 
quite close) : 

CH 4 + -'CI- ► CH 3 - + HCI 

The methyl radical resulting from the attack of atomic chlorine on a hydrogen 



sec 2.5 substitution reactions of saturated hydrocarbons 3 1 

of methane can then remove a chlorine atom from molecular chlorine and 
form chloromethane and a new chlorine atom: 

CH 3 - + Cl 2 ► CH3CI + :CI- 

An important feature of the mechanistic sequence postulated for the chlori- 
nation of methane is that the chlorine atom consumed in the first step is 
replaced by another chlorine atom in the second step. This type of process is 

CH 4 + :g- ► CH 3 - + HCI 

CH 3 - + Cl 2 ► CH3CI + :C1- 



CH 4 + Cl 2 ► CH3CI + HCI 



called a chain reaction since, in principle, one chlorine atom can induce the 
chlorination of an infinite number of methane molecules through operation 
of a "chain" or cycle of reactions. In practice, chain reactions are limited 
by so-called termination processes, where chlorine atoms or methyl radicals 
are destroyed by reacting with one another, as shown in these equations: 

CH 3 - + :C1- ► CH3CI 

2 CH 3 - > CH3CH3 

Chain reactions may be considered to involve three phases. First, chain 
initiation must occur, which for chlorination of methane is activation and 
conversion of chlorine molecules to chlorine atoms by light. In the second 
phase, the chain-propagation steps convert reactants to products with no 
net consumption of atoms or radicals. The propagation reactions occur in 
competition with chain-terminating steps, which result in destruction of atoms 
or radicals. 



light •• 

CI, ► 2:C1- chain initiation 



CH 4 + :Cl- ► CH 3 - + HCI 

CH3-+ Cl 2 ► CH3CI +:cV chai »P r <W ti0 » 



► CH3CI 

CHj- + CHj- ► CHjCH 3 



chain termination 



The two chain-termination reactions for methane chlorination as shown 
might be expected to be exceedingly fast, because they involve combination 
of unstable atoms or radicals to give stable molecules. Actually, combination 
of chlorine atoms does not occur readily in the gas phase because there is 
almost no way for the resulting molecule to lose the energy of reaction except 
by redissociating or colliding with some third body, including the container 



chap 2 the C± and C 2 hydrocarbons 32 

wall. The products in the other termination steps shown above, by contrast, 
can take care of this energy by redistributing it as vibrational excitation of 
their C— H bonds. Collisions with other molecules then disperse the excess 
vibrational energy throughout the system in the form of heat. 

If much chain propagation is to occur before the termination steps destroy 
the active intermediates the propagation steps must themselves be very 
fast. However, propagation is favored over termination when the concen- 
trations of radicals (or atoms) are low because then the chance of two radicals 
meeting (termination) is much less likely than encounters of radicals with 
molecules which are present at relatively high concentrations (propagation). 

The overall rates of chain reactions are usually slowed significantly by 
substances which can combine with atoms or radicals and convert then into 
species incapable of participating in the chain-propagation steps. Such 
substances are often called radical traps, or inhibitors. Oxygen acts as an 
inhibitor in the chlorination of methane by rapidly combining with a methyl 
radical to form the comparatively stable (less reactive) peroxymethyl radical, 
CH 3 00 • . This effectively terminates the chain. Under favorable conditions, 
the methane-chlorination chain may go through 100 to 10,000 cycles before 
termination occurs by free-radical or atom combination. The efficiency (or 
quantum yield) of the reaction is thus very high in terms of the amount of 
chlorination that occurs relative to the amount of the light absorbed. 



C. REACTIVE INTERMEDIATES 

We have seen that the chlorination of methane proceeds stepwise via highly 
unstable intermediate species such as chlorine atoms (CI • ) and methyl radicals 
(CH 3 • ). Are there other molecules or ions which are too unstable for isolation 
but which might exist as transient intermediates in organic reactions ? If we 
examine simple Q species we find the possibilities : 

H H H 

H:6 or CH 3 - H:C e or CH® H:C: e or CH 3 : e H:CH or:CH 2 

H H H 

methyl cation methyl anion methylene 

(a carbonium ion) (a carbanion) (a carbene) 

Each of these C L entities possesses some serious structural defect that makes 
it much less stable than methane itself. The methyl radical is unstable because 
it has only seven electrons in its valence shell. It exists for only brief periods of 
time at low concentrations before dimerizing to ethane : 

CH 3 - + CH 3 - ► CH 3 — CH 3 

The methyl cation, CH 3 ®, is an example of a carbonium ion. This species 
has only six valence electrons ; moreover, it carries a positive charge. Methyl 
cations react with most species that contain an unshared pair of electrons — 
for example, a chloride ion, CH 3 ® + Cl e -»■ CH 3 — CI. We shall see later, 
however, that the replacement of the hydrogens of the methyl with other 



sec 2.5 substitution reactions of saturated hydrocarbons 33 

groups, such as CH 3 — or C 6 H 5 — , can' provide sufficient stability to make 
carbonium ions important reaction intermediates. In fact, they can sometimes 
be isolated as stable salts (Section 24-5). Having only three pairs of electrons 
about the central carbon atom, carbonium ions tend toward planarity with 
bond angles of 120°. This gives the maximum separation of the three pairs 
of electrons. 



H 

k 120° 

methyl cation (a carbonium ion) 

The methyl anion, CH 3 e , does possess an octet of electrons but bears a 
negative charge, for which it is ill suited by virtue of carbon's low electronega- 
tivity. Such carbanions react rapidly with any species that will accept a share 
in an electron pair — any proton donor, CH 3 e -5^. CH 4 , for example. We shall 
see later that carbanions can be stabilized and rendered less reactive by having 
strongly electron- withdrawing groups, such as nitro ( — N0 2 ), attached to 
them, and we shall encounter such carbanions as reaction intermediates in 
subsequent chapters. Carbanions have four pairs of electrons about the cen- 
tral carbon atom. The mutual repulsion of the electrons gives the ion a pyra- 
midal shape. The methyl anion, for example, is isoelectronic with ammonia 
and is believed to have a similar shape : 



c 


N 


H-7 ^H 


h"/ 


H 


H 



The nonbonding electron pairs in each of the above molecules are expected 
to repel the bonding pairs more than the bonding pairs repel each other 
(see pp. 8-9 and Exercise 1-11). This accounts for the fact that the H— N— H 
bond angles in ammonia (Section 1-2A) are slightly less than the tetrahedral 
value. 

Methylene, :CH 2 , like a carbonium ion, possesses only a sextet of electrons 
in its valence shell and tends to react rapidly with electron donors (Section 9-7). 

It is important to be able to deduce the overall charge of a species from its 

H 

electronic arrangement and vice versa. A carbonium ion, such as H: C®, is 

a 

positively charged because, although the hydrogens are formally neutral, the 
carbon atom has a half-share of six valence electrons. It thus has, effectively, 
only three electrons in its own outer shell instead of the four electrons that a 

H 
neutral carbon atom possesses. By contrast, a carbanion such as H:C: e is 

H 

negatively charged because the carbon has an unshared electron pair in addi- 



chap 2 the C x and C 2 hydrocarbons 34 



H^ 


r^^n^ 


H 




H 






JA 




^c cC 




or 




- C v 


,cC 




H" 


"^fiXjS^ 


~"H 




H" 




— 'St 


^H 



Figure 2-7 Bent bonds in ethene. 

tion to its half-share of the six bonding electrons, giving it effective possession 
of five valence electrons, one more than that of a neutral carbon atom. 



2-6 addition reactions of unsaturated hydrocarbons 

The two simplest unsaturated compounds (those containing a multiple bond) 
are ethene (CH 2 =CH 2 ) and ethyne (HC=CH). The generally lower stability 
of multiply bonded compounds arises from the restriction that only one 
electron pair can occupy a given orbital. Because the most effective region for 
interaction between an electron pair and the two nuclei it links is along the 
bond axis, we expect to find this to be the region of highest electron density 
in a single bond. Bonds that run symmetrically along the bond axis are called 
sigma bonds (a bonds). 

In a molecule such as ethene, though, the two carbon atoms are linked by 
two electron pairs and it is quite clear that only one pair can occupy the prime 
space along the bond axis. There are two ways of looking at the electronic 
arrangement of ethene. The first, and simplest, is to consider the two orbitals 
used to link the carbons together as being identical, both being bent, somewhat 
like the arrangement taken up by springs in the ball-and-stick model of ethene 
(Figure 2-7). This simple view of the bonding of ethene accounts for its mole- 
cular geometry (repulsion between the electron pairs produces a planar 
molecule) and its chemical reactivity (the electron pairs are not bound as 
tightly to the nuclei as in the case of ethane where space along the bond axes 
can be used). 

The alternative way of looking at the bonding of ethene is to consider the 
double bond as being made up of an ordinary single-type bond along the bond 
axis, a a bond, and a second bond occupying the regions of space above and 
below the plane of the molecule, a it bond (pi bond). (A % bond is not cylin- 
drically symmetrical along the bond axis.) The shaded regions above and below 
the plane of the molecule (see Figure 2-8) represent just one bonding orbital : 

Figure 2-8 a and n bonding in ethene. 





7T 








W 














H^ 




^%k 


H 




( 


l> cC 




H" 


^ttn 


a 


^H 



sec 2.6 addition reactions of unsaturated hydrocarbons 35 

that which corresponds to the % bond. This model also accounts for the geo- 
metry of ethene and for its high chemical reactivity. Despite being less simple 
than the bent-bond model, it is extensively used by chemists to describe the 
bonding in unsaturated systems. In the case of ethyne, two n bonds and one 
a bond can be said to link the two carbons together. 

Neither model can be called correct or incorrect. And, indeed, refined theory 
suggests that they represent equivalent approximations. Each is useful accord- 
ing to the degree of clarity it gives the user and according to its ability to 
predict molecular behavior. The language of chemistry (particularly the 
theory of spectroscopy and bonding) is based to a great extent on the a, n 
model, and it must be considered by anyone wishing to explore organic 
chemistry in depth. Paradoxically, it is only recently that the simple bent- 
bond model has received much attention from theorists. 

Both models account for the shortening of the carbon-carbon bond 
distance as the number of bonds between the carbons increases — the greater 
the forces, the shorter the bond distance. We shall see that the analogy of 
springs and bonds also accounts for the vibrational energies of these com- 
pounds. A greater amount of energy is required to increase the vibration in 
the strong triple bond of ethyne than in the double bond in ethene or the single 
bond in ethane. 



/1.54 A 


/ 1.34 A 


.1.20A 


H / H 

\ / / 
-C-C-H 

/ \ 
H H 


H / H 

\ i / 

C = C 

/ \ 

H H 


/ 

H-C^C-H 



ethane ethene ethyne 

A. HYDROGENATION OF MULTIPLE BONDS 

The reaction of hydrogen with ethyne is highly exothermic: 

H-C=C-H + H 2 ► CH,=CH, A//„ p = - 42.2 kcal 

Alcaic = — 40 kcal (from bond 

energies in Table 2-1) 

Likewise, the further reduction of ethene to ethane is highly exothermic : 

CH 2 =CH 2 + H 2 ► CH3-CH3 A// cxp = - 32.8 kcal 

AH C1 | C = - 30 kcal (from bond 

energies in Table 2- 1) 

However, mixtures of either compound with hydrogen are indefinitely stable 
under ordinary conditions, and this again reminds us that reaction rates 
cannot be deduced from heats of reaction. Both ethyne and ethene, however, 
react rapidly and completely with hydrogen at low temperatures and pressures 
in the presence of metals such as nickel, platinum, and palladium. For 
maximum catalytic effect, the metal is usually obtained in a finely divided 
state. This is achieved for platinum and palladium by reducing the metal 
oxides with hydrogen before hydrogenating the alkene or alkyne. A specially 



chap 2 the C t and C 2 hydrocarbons 36 

active form of nickel (" Raney nickel ") is prepared from a nickel-aluminum 
alloy; sodium hydroxide is added to dissolve the aluminum, and the nickel 
remains as a black, pyrophoric powder. 



2N1-A1 + 2 0H e + 2H 2 



-» 2Ni + 2A10, B + 3H, 



Highly active platinum, palladium, and nickel catalysts can also be prepared 
by reducing metal salts with sodium borohydride. 

Besides having synthetic applications, catalytic hydrogenation is useful 
for analytical and thermochemical purposes. The analysis of a compound 
for the number of double bonds is carried out by measuring the uptake of 
hydrogen for a given amount of sample. 

The reaction occurs on the surface of the catalyst to which the reacting 
substances may be held loosely by van der Waals forces or, more tightly, by 
chemical bonds. The relatively loosely held electrons in a double or triple 
bond participate in forming carbon-metal bonds to the surface while hydrogen 
combines with the surface to give metal-to-hydrogen bonds (see Figure 2-9). 
The new bonds are much more reactive than the old ones and allow combina- 
tion to occur readily. The hydrogenated compound is then replaced on the 
surface by a fresh molecule of unsaturated compound, which has a stronger 
attraction for the surface. 

This is an example of heterogeneous catalysis — a type of reaction which 
involves adsorption on a surface of a solid or liquid and is often hard to 
describe in precise terms because the chemical nature of a surface is hard to 
define in precise terms. Homogeneous catalysis occurs in solution or in the 
vapor state. For such reactions, it is usually easier to trace the path from 
reactants to products in terms of intermediates with discrete structures. The 
light-induced chlorination of methane is an example of a homogeneous 
reaction. 

Ethene will add one mole of hydrogen while ethyne can add either one or 



Figure 2-9 Schematic representation of the hydrogenation of ethene on the 
surface of a nickel crystal. 



H 



H 



CH, 



CH 2 



yy\y y 



y\A ./ 



y y y '/ 

























y 




























nickel crystal 



sec 2.6 addition reactions of unsaturated hydrocarbons 37 

two. Special catalysts are available which will convert ethyne to ethene much 
faster than they convert ethene to ethane. 



B. ADDITION OF BROMINE TO MULTIPLE BONDS 

When ethene is treated with bromine (or with chlorine), a rapid addition occurs 
to give the 1 ,2-dihaloethane. The name given to the reaction product can be 



Br 2 + CH 2 =CH 2 ► BrCH 2 -CH 2 Br 

1 ,2-dibromoethane 

understood as follows. First, it is saturated (no double or triple bonds) and 
contains two carbon atoms and therefore is a derivative of ethane. The posi- 
tions of two of the hydrogen atoms of ethane have been taken by two bromine 
atoms ; hence it is a dibromoetham. The two bromines are located on different 
carbon atoms. We call one of the carbon atoms number 1 and the other num- 
ber 2, hence the name 1,2-dibromoethane. This naming system can be used to 
name the vast majority of organic compounds. In using it one should remem- 
ber these points . 

1. Pick out the parent hydrocarbon. Here it is ethane, not ethene, because 
we are interested in designating the structure, not the way the compound is 
formed in any particular reaction. 

2. Pick out those groups or atoms that replace any of the hydrogens of the 
parent compound and join their names to the front of the name of the parent 
hydrocarbon. If there are two such groups, as in the case we are working with, 
designate them with the prefix di; three such groups, tri; and so on. The name 
at this stage should be all one word — for example, dibromoethane. 

3. Locate the substituent groups by counting the carbon atoms from the 
end of the carbon chain. One can start numbering at either end in the example 
above, but this will not be generally true. The numbers designating the carbon 
atoms that bear substituents are separated from one another by commas and 
from the rest of the name by a hyphen — for example, 1,2-dibromoethane. 

Remember that the number of substituents is designated by the prefixes 
di, tri, and so on, but that their locations are designated by the numerals that 
identify particular carbon atoms. 

We shall examine nomenclature further in the next chapter, which deals with 
alkanes. 

The rapidity of the reaction of ethene with bromine illustrates the high 
reactivity of carbon-carbon double bonds. It would be natural to suppose 
that this reaction occurs by a simple simultaneous addition of both atoms of 
bromine to the double bond : 

Br- Br 
It will be recalled from the discussion of the methane-chlorine reaction, how- 



chap 2 the C x and C 2 hydrocarbons 38 

ever, that four-center reactions are rare, and in fact the addition of bromine to 
alkenes is known to occur in a stepwise manner, usually (but not always) to 
involve ionic intermediates, and usually to proceed by addition of the two 
bromines from opposite sides of the double bond. The evidence for this course 
of reaction is described in detail in Chapter 4 on alkenes. Suffice it to say here 
that an examination of the product of the ethene-bromine addition reaction 
will not tell us whether addition occurred from the same or the opposite 
side of the double bond, since rotation about the C— C bond converts one 
product into the other and thus obliterates the distinction between the 
two possible products (Equation 2-1). Both products are different confor- 
mations of the same compound, 1,2-dibromoethane. 



Br 



H 

J 

H 

Br 




-H 
H 




Addition of bromine to ethyne also occurs readily to give the compound 
1,2-dibromoethene. (Note that the parent compound is now ethene, not 
ethyne.) The product of this reaction is a liquid, bp 108°, mp —6.5°. It is 

Br 2 + H-CsC-H ► BrCH=CHBr 

1,2-dibromoethene 

immiscible with water and easily can be shown to possess no dipole moment. 
However, there is another known compound that has the same structure 
— that is, possesses two carbon atoms joined by a double bond and has a 
bromine and hydrogen atom on each carbon. This second compound is a 
liquid, bp 110°, mp —53°. It is also immiscible with water but has a large 
dipole moment. The two isomers arise because of a lack of rotation about the 
double bond. The molecule with the bromine atoms on the opposite side is 
called the trans isomer, and the other the cis isomer. 

Br H Br Br 

w 

/ \ / \ 

H Br H H 

trans- 1,2-dibromoethene cis- 1,2-dibromoethene 

These two compounds are said to have different configurations. It is worth 
reviewing here the meanings of the terms structure, conformation, and 
configuration. Structure designates the atoms that are linked together and the 
bonds that do this. Compounds with the same formula, C 2 H 4 Br 2 for example, 
but different structures, such as Br 2 CH— CH 3 and BrCH 2 —CH 2 Br, are 
called structural isomers. If the compound contains one or more carbon- 



sec 2.6 addition reactions of unsaturated hydrocarbons 39 

carbon single bonds, rotation can usually occur freely about these and give 
rise to different conformations. If the compound contains double bonds (or a 
ring of atoms), rotation is prevented and different configurations may then 
be possible. Compounds with different configurations are called stereoisomers 
— for example, cis- and ?ranj-l,2-dibromoethene — and they can only be 
interconverted by the rupture of chemical bonds. We shall encounter later in 
the book (Chapter 14) a more subtle form of stereoisomerism. The form of 
stereoisomerism we are discussing here is called either cis-trans isomerism or 
geometrical isomerism and, harking back to our ball-and-stick models, it is 
easy to rationalize why interconversion between geometrical isomers does 
not take place readily (Equation 2-2). For rotation to occur, one of the C— C 



Br H Br Br Br 

\ / A. A . H \ / 

/ C=C \ ' / C & C -Br ' r\ (2-2) 

H Br H H H 

trans cis 



bonds of the double bond must be broken. For this, the necessary energy 
input would be roughly equal to the difference in energy between a double 
and a single bond — 63 kcal/mole (see Table 2-1). Such an amount of energy 
is not available from molecular collision at ordinary temperatures. 

It is a simple matter to assign configurations to the two geometrical isomers 
of BrCH=CHBr. The one formed by the addition of bromine to ethyne has 
no dipole moment and hence must be the trans isomer. The boiling points of 
these compounds are nearly the same, but the melting points are vastly 



Br 



r<< Vv 

H BA H H 

trans isomer cis isomer 

dipoles cancel, fi = dipoles do not cancel, (i=1.3D 
bp 108°, mp - 6.5° bp 110°, mp - 53° 



different. Trans compounds often have somewhat higher melting points than 
the corresponding cis isomers, reflecting greater ease of crystal packing of 
their somewhat more symmetrical molecules. Dihaloethenes are exceptional 
in that the cis isomers tend to be slightly more stable than the trans. This is 
because the distances between the halogens in these compounds (but not 
usually between other substituents) are just right for operation of favorable 
van der Waals attractive forces. With most other substituents, particularly 
if they are bulky, the trans arrangement is preferred. Where three or four 
different groups are attached to the double bond, you must define what you 
mean by the terms cis and trans, and the generalizations given here about 
melting points and stabilities do not apply. 

Will 1,1-dibromoethene (CH 2 =CBr 2 ), which is a structural isomer of the 
above compounds, also exist in cis and trans forms? Clearly not, because 



chap 2 the C t and C 2 hydrocarbons 40 

interchange of the two bromines on one carbon or interchange of the two hyd- 
rogens on the other produces the same molecule. The requirement for the 
existence of geometrical isomers of an alkene is that the two groups on one 
end of the double bond be different from each other and the two groups on the 
other end be different from each other. 

Geometrical isomerism does not arise with triply bonded compounds, 
because the — C=C— bonds in these molecules are linear. 



summary 

There are four Q and C 2 hydrocarbons: methane (CH 4 ) and ethane (CH 3 — 
CH 3 ) are alkanes, ethene (CH 2 =CH 2 ) is an alkene, and ethyne (HC=CH) is 
an alkyne. Ethene is a planar molecule, ethyne is linear, and ethane nonplanar. 
Rotation around the single bond in ethane can be seen to give rise to an in- 
finite number of arrangements called conformations, two extreme forms of 
which are designated eclipsed and staggered. 

All hydrocarbons undergo combustion — complete oxidation to carbon 
dioxide and water. Heats of combustion and of other reactions can be cal- 
culated from bond energy data. 

The hydrogens in alkanes can be replaced by halogen atoms via a substi- 
tution process — for instance, CH 4 + Cl 2 -> CH 3 C1 + HC1. The position of 
equilibrium of this reaction is far to the right, the equilibrium constant K 
being determined by the free energy of reaction, AG, which in turn is related 
to the heat of reaction, the entropy factor, and the absolute temperature: 
AG = -RTln K = AH - T AS. Reactions that are fairly exothermic (-AH > 
15 kcal) usually have large K values. The rate of a reaction cannot be re- 
lated in a simple way to its overall AH or K. It depends on the reaction path. 
For the chlorination of methane, this involves a chain reaction with the follow- 
ing steps : Cl 2 is cleaved by light to give CI • atoms (initiation) ; a hydrogen 
atom is abstracted from CH 4 by CI- to give CH 3 -, a methyl radical; the 
radical, in turn, abstracts a chlorine atom from Cl 2 . 

CH 4 + CI- ► CH 3 - + HC1 

CH 3 - + Cl 2 ► CHjCI + CI- 

These two steps are the propagation steps in the chain reaction, CI • being con- 
sumed in the first step and regenerated in the second ; the chain is terminated 
when radicals or atoms combine. 

Some important intermediate species that will be encountered in other 
reactions, in addition to radicals such as CH 3 -, are carbonium ions, such as 
CH 3 ®; carbanions, such as CH 3 e ; and carbenes, such as :CH 2 . 

The double bonds in alkenes can be considered as two identical bent bonds, 
or as one bond along the bond axis (a a bond) and a second bond above and 
below the plane of the molecule (a % bond). 

Unsaturated hydrocarbons undergo addition reactions, as with hydrogen or 



exercises 41 

halogens. Addition of bromine is often very fast while the addition of hydro- 
gen requires the presence of a heterogeneous catalyst such as finely divided 
nickel : 

HC = CH __ ^ h 2 C=CH 2 — ^-* CH3-CH3 

Ni Ni 3 j 

HC = CH tj -* BrCH=CHBr Br2 > Br 2 CH — CHBr, 

Compounds such as 1,2-dibromoethene (BrCH=CHBr) can exist as two 
geometrical isomers, designated cis and trans. These two compounds have 

Br Br Br H 

\ / \ / 

c=c c=c 

/ \ / \ 

H H H Br 

cis- 1,2-dibromoethene /ranj-l,2-dibromoethene 

different physical properties and, because of the restricted rotation about the 
double bond, are quite stable to interconversion. The requirements for geo- 
metrical isomerism at a double bond are that two different groups be attached 
to one carbon atom and two different groups be attached to the other. Cis and 
trans isomers (geometrical isomers) have the same structure but different 
configurations. (The many arrangements that arise because of rotation about 
a single bond, as in ethane, are called conformations.) 



exercises 

2-1 Show how the two conventions of Figure 2-4 can be used to represent the 
possible staggered conformations of the following substances : 

a. CH 3 CH 2 C1 (chloroethane) 

b. CH 2 C1CH 2 C1 (1,2-dichloroethane) 

c. CH3CH2CH2CH3 (butane); consider rotation about the middle two 
carbon atoms in this compound. 

2-2 Use the bond-energy table to calculate A/7 for the following reactions in the 
vapor phase at 25° : 

4 H 2 



a. 


CH 3 CH 2 CH 3 + 5 2 ->3CO 


b. 


CH 4 + f 2 ->CO + 2H 2 


c. 


CO + 3H 2 -CH 4 + H 2 


d. 


CHU + 4 Cl 2 -»■ CCU + 4 HC1 


e. 


CH 4 + I 2 ->CH 3 I+HI 



2-3 Calculate AH for C(s)-+C(g) from the heat of combustion of 1 gram-atom 
of solid carbon (94.05 kcal) and the bond energies in Table 2- 1 . 

2-4 Write balanced equations for the complete and incomplete combustion of 
ethane to give, respectively, carbon dioxide and carbon monoxide. Use the 



chap 2 the C x and C 2 hydrocarbons 42 

table of bond energies to calculate the heats evolved in the two cases from 10 g 
of ethane. 

2-5 A possible mechanism for the reaction of chlorine with methane would be to 
have collisions where a chlorine molecule removes a hydrogen according to 
the following scheme: 

.. .. slow 
CH 3 :H + :Cl:Cl: «• CH 3 - + H:C1: + :C1- 

. , " ' fact .. 

(?' CH 3 - + :C1- ►■CH 3 :Cl: 

Use appropriate bond energies to assess the likelihood of this reaction mech- 
anism. What about the possibility of a similar mechanism with elemental 
fluorine and methane? 

2-6 Write the steps of a chain reaction for the light-catalyzed chlorination of 
ethane. 

2-7 Calculate A/f for each of the propagation steps of methane chlorination by 
a mechanism of the type 

(IV 

Cl 2 »• 2 CI- initiation 

C1+CH 4 ► CH 3 C1 + H- 



H- + C1, > HCl + Cl- ' P r °P a g ation 

Discuss the relative energetic feasibilities of these chain-propagation steps in 
comparison with those of other possible mechanisms. 

2-8 How many dichloro substitution (not addition) products are possible with 
(a) methane, (b) ethane, (c) ethene, (d) ethyne? 

29 Show that the methyl radical, CH 3 , has no charge. Show that the carbons of 
the neutral molecules CH and CH 2 are electron deficient, that is, they do not 
possess an octet of valence electrons. 

2-10 Which of the following molecules or ions contain a carbon atom that lack 
an octet of electrons: CH 3 CH 2 -, CH 3 ®, CH 3 e , CH 3 CH 3 , HC^CH, CH 2 :? 



2-11 Consider the feasibility of a free-radical chain mechanism for hydrogenation 
of ethene in the vapor state at 25° by the following propagation steps : 

CH 3 — CH 2 +H 2 ► CH 3 — CH 3 + H- 

CH 2 =CH 2 + H- ► CH 3 -CH 2 - 

2-12 Name the following compounds: 

a. C1CH 2 CC1 3 d , ClCsCCl 

b. C1 2 CHCHC1 2 e . Br 2 C=CBr 2 

c. BrCH=CH 2 



exercises 43 

2-13 Provide structures for the following compounds: 

a. 1,1,1-tribromoethane 

b. 1,1,2-tribromoethane 

c. l-chloro-2-fiuoroethene 

d. bromoethyne 

2-14 Is geometrical isomerism possible in the following cases? Draw formulas to 
show the configurations of the cis and trans isomers where appropriate. 

a. Br 2 C=CHCl d. C1C=CC1 

b. BrClC=CHBr e. CHF=CHF 

c. C1 2 CH— CHC1 2 t 

2-15 How many grams of bromine will react with (a) 20 g of ethene, (b) 20 g of 
ethyne? / 

2-16 What volume of carbon dioxide (dry, at 20° and 1 atm) will be obtained by the 
complete combustion of (a) 20 g of ethene, (b) 20 g of ethyne? 

2-17 The reaction HC=CH(<?) + H 2 (g)-> CH 2 =CH 2 (#) has the following 
thermodynamic parameters at 25 °C. 

AG = -33.7 kcal 
A#= -41.7 kcal 
^S = -26:8 eal 

a. Does the position of equilibrium favor reactants or product? 

b. Does the entropy term favor formation of reactants or product ? Explain 
the significance of your answer about the entropy in terms of the 
relative freedom of the atoms in the reactants and product. 



mmmm 



R5& 




RfK' «*ffi« : >a*7a«C WESaSiaSHJBMW S"'»SSS*4Si9: T^T "r-.»-R«-.i,F S3 pK»2IlSi 






ii§i 



*5? AsSi 



ii?i 






>s* *{;« »>M «#■ .ftii gei#fe«s sf» : #§s 1^1 

biisLS ^fe^ii g»#&sV4MI SaiKlfeag 5S^*Siffii_:KS&i:l: 









111! 

a& •ai 



K;*'% W 



[v^ %$iff& ; As 



lltllli 










.W.V* 

'-^j- : 


■¥? 


2f?M 


^^H 




**$£ 


<1& 


'V^iP'^-il 












v-SvF : 


ii'T.r; 


VJV- 


.\^5 


t&i 


iSfev.', 


lli 


^# 


& 


: .4*5vtS?d 




'"■!W:J 


?$■■? 


;,«; 








;-ji>^-: 










.¥££* 


tf'iR 










F vi^«.^a 










-* £••* -M 


,; .''a?'.' 






&*>& 


•:'^>*-g-l- 


■ -jjfcs* 


HUH 






&&&*#! 


'il^T 






iS 


^$S 



Ktv;A-*'.'.vv.^s",-ii.-v rv«Vvs»Y/?s>i,?. Avw:%ft-a &:■£& feSiS^&is 5i«j. '•«;*• ".Jfefe "©iSMiS 






P;s. ; 'i5iw] 



S$" •$ 



S' 4ifcSW 









1:8 S^SAllftSl^^^ 1P1 



chap 3 alkanes 47 

In the previous two chapters we have studied in some detail the properties 
of the two simplest saturated hydrocarbons, methane and ethane, and have 
shown how their simple derivatives are named using the rules of the Inter- 
national Union of Pure and Applied Chemistry (IUPAC rules). In this 
chapter we shall examine the larger alkanes, including those that are cyclic 
(cycloalkanes). Open-chain, or acyclic, alkanes have the general formula 
C„H 2 „+ 2 , whereas cycloalkanes have the formula C„H 2 „ . It is essential that 
one be able to name compounds correctly and we will begin our discussion . 
of alkanes with a survey of nomenclature. 



3 ' 1 nomenclature 

The IUPAC rules for naming alkanes are simple and easy to apply. -However, 
few people adhere strictly to these rules and it is necessary to be familiar with 
some other commonly used naming terms; these are often simple and con- 
venient when applied to simple compounds but become cumbersome or 
ambiguous with more complex compounds. The IUPAC name for a compound 
is always acceptable; hence, when asked to supply a name to a compound 
whose structure is given, it is best to follow the IUPAC rules. You should 
become familiar enough with other common terms, however, so you can 
supply the correct structure to a compound whose name is given in non- 
IUPAC terminology. 

The alkanes are classified as "continuous chain" (i.e., unbranched) if all 
the carbon atoms in the chain are linked to no more than two other carbons, 
or " branched chain " with one or more carbon atoms linked to three or four 
other carbons. Branching is only possible with alkanes C 4 and up. 1 



HiC CH* CHi 

II I 

CH 3 -CH 2 -CH 2 -CH 2 -CH 2 -CH 3 CH 3 — C-C-CH 3 CH 3 -C-CH 2 -CH 3 

H H CH 3 

continuous-chain hydrocarbon branched-chain hydrocarbons 

The first four continuous-chain hydrocarbons have nonsystematic names : 

CH 4 CH 3 -CH 3 CH 3 -CH 2 -CH 3 CH 3 -CH 2 -CH 2 -CH 3 

methane ethane propane butane 

The higher members, beginning with pentane, are named systematically with 
a numerical prefix (pent-, hex-, hept-, etc., to denote the number of car- 
bon atoms) and with the ending -ane to classify the compound as a saturated 
hydrocarbon. Examples are listed in Table 3T. These names are generic of 



1 The notation C 4 means a compound containing four carbon atoms whereas C-4 means 
the fourth carbon atom in a chain. 



chap 3 alkanes 48 
Table 3-1 Continuous-chain alkanes (C„H 2 „ +2 ) 







no. branched-chain 


no. of carbons, n 


name 


of isomers 


1 


methane 





2 


ethane 





3 


propane 





4 


butane 


1 


5 


pentane 


2 


6 


hexane 


4 


7 


heptane 


8 


8 


octane 


17 


9 


nonane 


34 


10 


decane 


74 


20 


eicosane 


366,318 


30 


triacontane 


4.11 X 10" 



both branched and unbranched hydrocarbons and, to specify a continuous- 
chain hydrocarbon, the prefix n- (for normal) is often attached. In the absence 
of any qualifying prefix the hydrocarbon is considered to be "normal" or 
unbranched. 

CH 3 -CH 2 - CH 2 - CH 2 - CH 3 

pentane 
(//-pentane) 

The possibility of branched-chain hydrocarbons isomeric with the con- 
tinuous-chain hydrocarbons begins with butane (n = 4). The total number of 
theoretically possible isomers for each alkane up to n = 10 is given in Table 
3-1, and is seen to increase very rapidly with n. All 75 possible alkanes from 
n = 1 to n = 9 inclusive have now been synthesized. 

There are two structural isomers of C 4 H 10 , one the continuous-chain 
compound called butane (or w-butane to emphasize its lack of branching) 
and the other called 2-methylpropane (or, in common terminology, isobutane). 



CH 3 -CH — CH 3 
CH 3 -CH 2 -CH 2 -CH 3 I 

CH 3 

butane 2-methylpropane 

( //-butane) (isobutane) 



The name 2-methylpropane is appropriate because the longest continuous 
chain in the molecule is made up of three carbons ; it is thus a derivative of 
propane. The second carbon (from either end) has one of its hydrogens 
replaced by a methyl group, hence the prefix 2-methyl. It should be re- 
membered that the shape of this molecule is not as depicted in the formula 
shown. The tetrahedral arrangement about the central carbon atom makes 
all three methyl groups equivalent (Figure 3-1). 



sec 3.1 nomenclature 49 




Figure 3-1 Shape of 2-methylpropane. 

There are three known compounds with the formula C 5 H 12 , and this is 
the number of isomers expected on the basis of carbon being tetravalent and 
hydrogen monovalent: 

CH 3 CH 3 

CH 3 -CH 2 -CH 2 -CH 2 -CH 3 CH 3 -CH-CH 2 -CH 3 CH 3 -C-CH 3 

CH 3 

pentane 2-methylbutane 2,2-dimethylpropane 

(;?-pentane) (isopentane) (neopentane) 

The prefix iso denotes a compound with two methyl groups at the end of a 
chain, and neo means three methyl groups at the end of a chain. 

When we come to the C 6 alkanes we find that there are five isomeric 
compounds, C 6 H 14 . Clearly, trivial prefixes become less and less useful as the 
length of the chain grows and the number of possible isomers increases. 
Accordingly, hexane and its four branched-chain isomers are here only given 
IUPAC names. (The second compound in the list might be called isohexane 
and the fourth neohexane.) 



CH 3 

I 
CH 3 — CH 2 — CH 2 -CH 2 — CH 2 — CH 3 CH 3 — CH-CH 2 — CH 2 — CH 3 

hexane 2-methylpentane 



CH 3 CH 3 CH 3 CH 3 

CH 3 -CH 2 -CH-CH 2 -CH 3 CH 3 -C-CH 2 -CH 3 CH 3 -CH-CH-CH 3 

CH 3 

3-methylpentane 2,2-dimethylbutane 2,3-dimethylbutane 

The branched-chain compounds considered so far all contain the simplest 
group, methyl, as substituent. Alkyl groups such as methyl are obtained by 
writing the alkane minus one of its hydrogens. Thus, we have methyl (CH 3 — ) 
and ethyl (CH 3 — CH 2 — ). A problem arises when we come to the alkyl 
groups of propane, CH 3 — CH 2 — CH 3 . The hydrogen atoms in this molecule 
are not all equivalent. Removing one of the six terminal hydrogens produces 
the n-propyl group, CH 3 — CH 2 — CH 2 — ; removing one of the two central 



chap 3 alkanes SO 

hydrogens produces the isopropyl group, CH 3 — CH— CH 3 (usually written 



V 



/ 



CH— or(CH 3 ) 2 CH— ). 



H 3 C 



Additional examples are listed in Table 3-2. These have been further classi- 
Table 3-2 Typical alkyl groups (C„H 2 „ + 1 ) 







primary (RCH 2 — ) 






CH 3 - 




CH3CH2 




CH3CH2CH2 — • 


methyl 




ethyl 




«-propyl 


CH3CH2CH2CH2 


CH 3 

\ 
CH-CH 2 - 










CH 3 






n-butyl 




isobutyl 






CH3CH2CH2CH2CH2 


CH 3 
CH— CH 2 -CH 2 - 




CH 3 

1 

CH3 — C CH 2 

1 




* 


CH 3 




CH 3 


pentyl 




isopentyl 




neopentyl 


(«-amyl) 




(isoamyl) 






CH3CH2CH2CH2CH2CH2 


CH 3 
V 
CH— CH 2 — CH 2 — CH 

CH 3 


CH 3 
1 
2 CH3 — C Cri2Cri2 








CH 3 


K-hexyl 




isohexyl 




neohexyl 


secondary (R 2 CH — ) 




CH 3 

\ 3 


CH 3 CH 2 








/C H 
CH 3 


- 

CH 3 








isopropyl s-butyl 






tertiary (R 3 C— ) 




CH 3 

1 


ch 3 

1 






CH 3 -C- 
1 


CH 3 CH 2 - 


c— 

1 






CH 3 


CH 3 






r-butyl 


r-pentyl 
(/-amyl) 







sec 3.1 nomenclature SI 

fied according to whether they are primary, secondary, or tertiary. An alkyl 
group is described as primary if the carbon at the point of attachment is 
bonded to only one other carbon, as secondary if bonded to two other carbons, 
and tertiary if bonded to three other carbons. The methyl group is a special 
case and is regarded as a primary group. 

In a few cases, where there is a high degree of symmetry, hydrocarbons are 
conveniently named as derivatives of methane or ethane. 



H 3 C v CH 3 

H >\ 7 -ii 

CH-C-H CH3-C-C-CH3 

/I II 

H 3 C CH H 3 C CH 3 

/ \ 
H 3 C CH 3 

triisopropyltnethane hexamethylethane 

In naming complex compounds, you must pick out the longest consecutive 
chain of carbon atoms. The longest chain may not be obvious from the way 
in which the structure has been drawn on paper. Thus the hydrocarbon [1] 
is a pentane rather than a butane derivative, since the longest chain is one 
with five carbons. 



H 3 C 



CH 3 
CH 2 



CH 3 -CH-CH- 



CH, 



[1] 
(dotted lines enclose 
longest chain of suc- 
cessive carbon atoms) 



The parent hydrocarbon is numbered starting from the end of the chain, 
and the substituent groups are assigned numbers corresponding to their 
positions on the chain. The direction of numbering is chosen to give the 
lowest sum for the numbers of the side-chain substituents. Thus, hydrocarbon 
[1] is 2,3-dimethylpentane rather than 3,4-dimethylpentane. Although the 
latter name would enable one to write the correct structure for this com- 
pound, it would not be found in any dictionary or compendium of organic 
compounds. 

5CH 3 'CH 3 

I I 

H 3 C *CH 2 H 3 C 2CH 2 

II II 

CH 3 — CH — CH— CHj not CH 3 -CH-CH-CH 3 

54 3 



2,3-dimethylpentane 3,4-dimethylpentane 



chap 3 alkanes 52 

Where there are two identical substituents at one position, as in [2], 
numbers are supplied for each. Remember that the numerals represent posi- 
tions and the prefixes di-, tri-, and so on represent the number of substituents. 
Note that there should always be as many numerals as there are substituents, 
that is, 2,2,3 (three numerals) and trimethyl (three substituents). 

H 3 C CH 3 
I I 
CH3-C-CH-CH3 
I 
CH 3 

2,2,3-trimethylbutane 
[2] 

Branched-chain substituent groups are given appropriate names by a simple 
extension of the system used for branched-chain hydrocarbons. The longest 
chain of the substituent is numbered starting with the carbon attached directly 
to the parent hydrocarbon chain. Parentheses are used to separate the number- 
ing of the substituent and the main hydrocarbon chain. The IUPAC rules 

23 
rliC CHi CHi 

CH 
CH 3 -CH 2 -CH 2 -CH 2 — CH-CH 2 — CH 2 -CH 2 -CH 2 -CH 3 

1 2 3 4 56 7 8 910 

5-(l-methy!propyl)decane 
(5-.s-butyldecanc) 

permit use of the substituent group names in Table 3-2, so that s-butyl can 
be used in place of (1-methylpropyl) for this example. 

When there are two or more different substituents present, the question 
arises as to what order they should be cited in naming the compound. Two 
systems are commonly used which cite the alkyl substituents (1) in order of 
increasing complexity or (2) in alphabetical order. We shall adhere to the 
latter system mainly because it is the practice of Chemical Abstracts. 2 Examples 
are given below. 

CH 3 CH 2 CH 3 
CH 3 -CH 2 — CH 2 — CH-CH-CH 2 — CH 3 

7 3 6 5 4 3 2 1 

4-ethyl-3-methyl heptane 
(i.e.. ethyl is cited before methyl) 



CH 3 -C-CH 3 
CH 3 -CH 2 -CH 2 — C-CH 2 -CH 2 -CH 2 — CH 2 — CH 2 -CH 3 

1 3 2 3 4| i Z b i lit9 i iQ 

CH 3 -CH 

CH 3 

4-/-butyl-4-isopropyldecane 

2 Biweekly publication of the American Chemical Society: an index to, and a digest of, 
recent chemical publications throughout the world. 



sec 3.2 physical properties of alkanes — concept of homology 53 

Derivatives of alkanes, such as haloalkanes (R— CI) and nitroalkanes 
(R— N0 2 ), where R denotes an alkyl group, are named similarly by the 
IUPAC system. Definite orders of precedence are assigned substituents of 
different types when two or more are attached to a hydrocarbon chain. Thus, 
alkanes with halogen and alkyl substituents are generally named as halo- 
alkylalkanes (not as alkylhaloalkanes) ; alkanes with halogen and nitro 
substituents are named as halonitroalkanes (not as nitrohaloalkanes). 

CH 3 

I 
CH 3 -CH 2 -CH,-CH,-CI CH 3 — CH,-CH-CH-CH 3 

N0 2 

l-chlorobutane 3-nitro-2-metliylpentane 

(n-bulyl ehloride) 

H 3 C CH 3 

\ I 

CH-CH 2 -Br CI-CH-CH 2 -CH 2 — N0 2 

H 3 C 

l-bromo-2-melhylpropane 3-chloro-l-nitrobutane 

(isobutyl bromide) 



Note that l-chlorobutane is written as one word whereas the alternate 
name «-butyl chloride is written as two words. In the former the chloro group 
is substituted for one of the hydrogens of butane and the position of sub- 
stitution is indicated by the numeral. The latter name is formed by simply 
combining names for the two parts of the compound just as one would do 
for NaCl, sodium chloride. 



3 ' 2 physical properties of alkanes — concept of homology 

The series of continuous-chain alkanes, CH 3 (CH 2 )„_ 2 CH 3 , shows a remark- 
ably smooth gradation of physical properties (see Table 3-3 and Figure 3-2). 
As you go up the series, each additional CH 2 group contributes a fairly 
constant increment to the boiling point and density and, to a lesser extent, 
to the melting point. This makes it possible to estimate the properties of an 
unknown member of the series from those of its neighbors. For example, the 
boiling points of hexane and heptane are 69° and 98°, respectively; a difference 
in structure of one CH 2 group therefore makes a difference in boiling point 
of 29°. This places the boiling point of the next higher member, octane, at 
98° + 29°, or 127°, which is close to the actual boiling point of 126°. 

Members of a group of compounds with similar chemical structures and 
graded physical properties and which differ from one another by the number 
of atoms in the structural backbone, such as the «-alkanes, are said to con- 
stitute a homologous series. The concept of homology, when used to forecast 
the properties of unknown members of the series, works most satisfactorily 
for the higher-molecular-weight members. For these members, the intro- 
duction of additional CH 2 groups makes a smaller relative change in the 
overall composition of the molecule. This is better seen from Figure 3-2, 



chap 3 alkanes 54 



200 



100 



-100 



-200 











































^•""'^ 






^^^ 










boiling 
point ^ 






yS&i 


nsity 










^ *~ 








/ 


, 


1 

melt in« point 








1 ^S 


















*^« 




ijllSIt 


/ 















0.75 



0.65 dl° 



10 



11 12 



0.55 



Figure 3 • 2 Dependence on n of melting points, and densities (c^ ) of straight- 
chain alkanes, CH 3 (CH 2 )„_ 2 CH 3 . 



which shows how the boiling points and melting points of the homologous 
series of normal alkanes change with the number of carbons, n. See also 
Figure 3-3. 

Branched-chain alkanes do not exhibit the same smooth gradation of 
physical properties as the «-alkanes. Usually, there is too great a variation in 



Figure 3-3 Dependence of A T (difference in boiling and melting points 
between consecutive members of the series of normal alkanes) on n (number 
of carbon atoms). 



80 i 


























60 




, bo 


ling f 


>oinl 


















V 




, 


























w 


— mei 


tinu p 


oint 














AT 40 




/ 






r 


/ 


\ 






















^„ / 


\^ 












20 










\ 


/ 














\ 








\ 


\f 
















10 



11 12 



sec 3.2 physical properties of alkanes — concept of homology 55 
Table 3-3 Physical properties of rc-alkanes, CH 3 (CH 2 )„ _ 2 CH 3 













refractive 






bp,°C 


mp, 


density, 


index, 


n 


name 


(760 mm) 


°C 


dl° 


n 20 D 


1 


methane 


-161.5 


-183 


0.424" 




2 


ethane 


-88.6 


-172 


0.546" 




3 


propane 


-42.1 


-188 


0.501" 




4 


butane 


-0.5 


-135 


0.579" 


1.3326" 


5 


pentane 


36.1 


-130 


0.626 


1.3575 


6 


hexane 


68.7 


-95 


0.659 


1.3749 


7 


heptane 


98.4 


-91 


0.684 


1.3876 


8 


octane 


125.7 


-57 


0.703 


1.3974 


9 


nonane 


150.8 


-54 


0.718 


1.4054 


10 


decane 


174.1 


-30 


0.730 


1.4119 


11 


undecane 


195.9 


-26 


0.740 


1.4176 


12 


dodecane 


216.3 


-10 


0.749 


1.4216 


15 


pentadecane 


270.6 


10 


0.769 


1.4319 


20 


eicosane 


342.7 


37 


0.786* 


1.4409 c 


30 


triacontane 


446.4 


66 


0.810 c 


1.4536 c 



At the boiling point. 

* Under pressure. 

c For the supercooled liquid. 



molecular structure for regularities to be apparent. Nevertheless, in any one 
set of isomeric hydrocarbons, volatility increases with increased branching. 
This can be seen from the data in Table 3-4, in which are listed the physical 



Table 3-4 Physical properties of hexane isomers 



isomer 


structure 


bp, 
°C 


mp, 

°C 


density 

at 20°, dl° 


hexane 


CH 3 (CH 2 ) 4 CH 3 

CH 3 

1 


68.7 


-94 


0.659 


3-methylpentane 


CH 3 CH 2 CHCH 2 CH 3 

CH 3 
1 


63.3 


-118 


0.664 


2-methylpentane 
(isohexane) 


CH 3 CHCH 2 CH 2 CH 3 

CH 3 CH 3 
1 1 


60.3 


-154 


0.653 


2,3-dimethylbutane 


CH 3 CH-CHCH 3 

CH 3 
| 


58.0 


-129 


0.661 


2,2-dimethylbutane 
(neohexane) 


CH 3 CCH 2 CH 3 

1 
CH 3 


49.7 


-98 


0.649 



chap 3 alkanes 56 

properties of the five hexane isomers; the most striking feature is the 19° 
difference between the boiling points of hexane and neohexane. 



33 alkanes and their chemical reactions 

As a class, alkanes are singularly unreactive. The name saturated hydrocarbon 
(or "paraffin," which literally means "little affinity" [L. par{um), little, 
+ affins, affinity]) arises because their chemical affinity for most common 
reagents may be regarded as saturated or satisfied. Thus none of the C— H 
or C— C bonds in a typical saturated hydrocarbon such as ethane are attacked 
at ordinary temperatures by a strong acid such as sulfuric, or by powerful 
oxidizing agents such as potassium permanganate, or by vigorous reducing 
agents such as lithium aluminum hydride (LiAlH 4 ). 

We have seen that methane and other hydrocarbons are attacked by oxygen 
at elevated temperatures and, if oxygen is in excess, complete combustion 
occurs to give carbon dioxide and water with the evolution of large amounts 
of heat. Vast quantities of hydrocarbons from petroleum are utilized as fuels 
for the production of heat and power, as will be described in the next section. 



A. PETROLEUM AND COMBUSTION OF ALKANES 

The liquid mixture of hydrocarbons delivered by oil wells is called petroleum. 
Its composition varies according to the location of the field but the major 
components are invariably alkanes. Natural gas is found in association with 
petroleum and also alone as trapped pockets of underground gas. Natural 
gas is chiefly methane, while crude petroleum is an astonishing mixture of 
hydrocarbons up to C 50 in size. This dark, viscous oil is present in interstices 
in porous rock and is usually under great pressure. 

Petroleum is believed to arise from the decomposition of the remains of 
marine organisms over the ages and new fields are continually sought to satisfy 
the enormous world demand. The effect of advanced technology on our 
environment is shown by the fact that combustion of fossil fuels, chiefly 
petroleum, has increased the carbon dioxide content of the atmosphere by 
10% in the past century and an increase of 25% has been predicted by the 
year 2000. These increases would be even more marked were it not for the 
fact that the rate of photosynthesis by plants becomes more efficient at 
utilizing carbon dioxide as the concentration increases. 

In addition to serving as a source of power — and being the only natural 
sources of suitable fuel for the internal combustion engine — petroleum and 
natural gas are extremely useful as starting materials for the synthesis of 
other organic compounds. These are often called petrochemicals to indicate 
their source but they are, of course, identical with compounds prepared in 
other ways or found in nature. 

Petroleum refining involves separation into fractions by distillation. Each 
of these fractions with the exception of the first, which contains only a few 
components, is still a complex mixture of hydrocarbons. The main petroleum 
fractions are given below in order of decreasing volatility. 



sec 3.3 alkanes and their chemical reactions 57 

1 . Natural Gas. Natural gas varies considerably in composition depending 
on the source but methane is always the major component, mixed with smaller 
amounts of ethane, propane, butane, and 2-methylpropane (isobutane). These 
are the only alkanes with boiling points below 0°C. Methane and ethane 
cannot be liquefied by pressure at room temperature (their critical tempera- 
tures are too low) but propane, butane, and isobutane can. Liquid propane 
(containing some of the C 4 compounds) can be easily stored in cylinders and 
is a convenient source of gaseous fuel. It is possible to separate natural gas 
into its pure components for sale as pure chemicals although the mixture is, 
of course, perfectly adequate as a fuel. 

2. Gasoline. Gasoline is a complex liquid mixture of hydrocarbons 
composed mainly of C 5 to C 10 compounds. Accordingly, the boiling range of 
gasoline is usually very wide, from approximately 40° to 180°. Because of the 
large number of isomers possible with alkanes of this size, it is much more 
difficult to separate gasoline into its pure components by fractional distillation 
than is the case with natural gas. Using a technique known as gas chromatog- 
raphy (Section 7-1), this separation can be done on an analytical scale. It has 
been shown that well over 100 compounds are present in appreciable amounts 
in ordinary gasoline. These include, besides the open-chain alkanes, cyclic 
alkanes (cycloalkanes, Section 3-4) and alkylbenzenes (arenes, Section 20T). 

The efficiency of gasoline as a fuel in modern high-compression internal 
combustion engines varies greatly with composition. Gasolines containing 
large amounts of branched-chain alkanes such as 2,2,4-trimethylpentane 
have high octane ratings and are in great demand, while those containing 
large amounts of continuous-chain alkanes such as octane or heptane have 
low octane ratings and perform poorly in a modern high-compression auto- 
mobile engine. The much greater efficiency of branched-chain alkanes is not 
the result of greater heat of combustion but of the smoothness with which 
they burn. The heats of combustion of octane and 2,2,4-trimethylpentane 
can be calculated from the data in Table 2T and, since each contains the 
same number of carbon-carbon bonds (seven) and carbon-hydrogen bonds 
(18), we would expect their heats of combustion to be identical. The calculated 
value is —1218 kcal/mole, and the experimental values are close to this. 



25. 

2 



CH 3 CH 2 CH 2 CH 2 CH 2 CH 2 CH 2 CH 3 + — 2 ► 8 C0 2 + 9 H 2 AH = -1222.8 kcal 



25. 
2 
CH, 



CH 3 -C-CH 2 -CH-CH 3 + — 2 ► 8C0 2 + 9H 2 AH= -1220.6kcal 



The combustion of vaporized branched-chain alkanes is slower and less 
explosive than that of continuous-chain compounds. In an automobile 
engine, too rapid combustion leads to dissipation of the combustion energy 
as heat to the engine block, rather than as movement of the piston. You can 
clearly hear the "knock" in an engine that is undergoing too rapid com- 
bustion of the fuel vapor in the cylinders. The problem is aggravated in high- 
compression engines. These will often continue to run (though inefficiently) 



chap 3 alkanes 58 

on low-octane gasoline even when the ignition switch has been turned off, 
the heat of compression in the cylinder being sufficient to ignite the fuel 
mixture. 

Combustion of hydrocarbons occurs by a complex chain reaction (Section 
2-5B). It is possible to slow the propagation of the chain by adding volatile 
compounds such as tetraethyllead, (CH 3 CH 2 ) 4 Pb, to gasoline. The fine 
particles of solid lead oxide formed by oxidation of tetraethyllead moderate 
the chain-carrying reactions and reduce the tendency for knock to occur. 3 
The octane rating of a gasoline is measured by comparing its knock with that 
of blends of 2,2,4-trimethylpentane whose octane rating is set at 100, and 
heptane whose octane rating is set at zero. Octane itself has a rating of — 20. 
The higher the octane rating, the smoother the ignition and, with high-com- 
pression engines, the more efficient the gasoline. With low-compression engines 
the effect is negligible and it is wasteful or worse (see footnote) to use gasoline 
of higher rating than is needed to eliminate knocking. 

The steadily increasing use of hydrocarbons in internal combustion engines 
has led to increasingly serious pollution problems. Some of these are associated 
with waste products from the refining operations used to produce suitable 
fuels from crude petroleum, others with spillage of petroleum in transit to the 
refineiies, but possibly the worst is atmospheric pollution from carbon mon- 
oxide and of the type known as "smog." The chemical processes in the 
production of smog are complex but appear to involve hydrocarbons (espe- 
cially branched-chain hydrocarbons), sunlight, and oxides of nitrogen. The 
products of the reactions are ozone, which produces rubber cracking and 
plant damage, particulate matter, which produces haze, oxides of nitrogen, 
which color the atmosphere, and virulent eye irritants (one being acetyl 
O 
II 
pernitrite, CH 3 — C— O— O— N=0). The hydrocarbons in the atmosphere 

which produce smog come principally from incomplete combustion in 
gasoline engines, although sizable amounts arise from evaporation and spill- 
age. Whether smog can be eliminated without eliminating the internal com- 
bustion engine is not yet known, but the prognosis is rather unfavorable. 

Pollution of the atmosphere from carbon monoxide is already so severe in 
heavy downtown traffic in large cities as to pose immediate health problems. 
The main reason for high concentrations of carbon monoxide in automobile 
exhaust is that the modern gasoline engine runs most efficiently on a slight 
deficiency in the ratio of oxygen to hydrocarbon which would produce 
complete combustion. A current solution to this problem is to introduce air 
and complete the combustion process in the exhaust manifold. 



3 Accumulation of lead oxide in a motor would rapidly damage the cylinder walls and 
valves. Tetraethyllead is normally used in conjunction with 1 ,2-dibromoethane in gasoline 
and this combination forms volatile lead bromide which is swept out with the exhaust 
gases. An unfortunate consequence of this means of improving octane rating is the addition 
of toxic lead compounds to the atmosphere. Isomerizing normal alkanes or cracking 
kerosene to produce branched-chain compounds would seem to be better solutions to the 
problem. The manufacture of engines capable of running on nonleaded gasoline is also 
being considered. 



sec 3.3 alkanes and their chemical reactions 59 

3. Kerosene. Kerosene consists chiefly of C u and C 12 hydrocarbons, 
compounds that do not vaporize well in automobile engines. It now finds 
considerable use as fuel for jet engines. It is also used in small heating units 
and can, if necessary, be converted to gasoline by a process known as cracking. 
This involves catalytic decomposition to smaller molecules, one of which is an 
alkene : 

heat 
CuH24 catalyst* C9H20 + CH 2 =CH 2 

4. Diesel Oil. The petroleum fraction which boils between about 250° and 
400° (C 13 to C 25 ) is known as diesel oil or fuel oil. Large amounts are used in 
oil-burning furnaces, some is cracked to gasoline, and much is used as fuel 
for diesel engines. These engines operate with a very high compression ratio 
and no spark system, so that they depend on compression to supply the 
heat for ignition of the fine spray of liquid fuel that is injected into the 
cylinder near the top of the compression stroke. Branched-chain compounds 
turn out to be too unreactive to ignite and for this reason diesel and auto- 
mobile engines have quite different fuel requirements. 

5. Lubricating Oils and Waxes. The high-molecular- weight hydrocarbons 
(C 26 to C 28 ) in petroleum have very high boiling points and can only be 
obtained in a reasonably pure state by distillation at reduced pressure. 
Thermal decomposition (pyrolysis) occurs if distillation is attempted at 
atmospheric pressure because the thermal energy acquired by collision of 
these compounds at their boiling points (400°) is sufficient to rupture carbon- 
carbon bonds. Almost all alkanes higher than C 20 are solids at room temper- 
ature, and you might be wondering why lubricating oil is liquid. This is 
because it is a complex mixture whose melting point is much below that of its 
pure components. Indeed, as the temperature is lowered, its viscosity simply 
increases although this undesirable property can often be corrected by special 
additives such as chlorinated hydrocarbons. 

Paraffin wax used in candles is a mixture of very high-molecular-weight 
hydrocarbons similar enough in structure to pack together to give a semi- 
crystalline solid. Vaseline is a mixture of paraffin wax and low-melting oils. 

6. Residue. After removal of all volatile components from petroleum, a 
black, tarry material remains which is a mixture of minerals and complex 
high-molecular- weight organic compounds; it is known as asphalt. 



B. SUBSTITUTION OF HALO AND NITRO GROUPS IN ALKANES 

The chlorination of methane was discussed in considerable detail in the 
previous chapter. This reaction actually occurs with all alkanes and can also 

be performed satisfactorily with bromine (but not with fluorine or iodine). 

I I 

For the general reaction — C— H + X, -» — C— X + H— X, where X = F, 

II 
CI, Br, or I, the calculated AH value is negative and very large for fluorine, 



chap 3 alkanes 60 
Table 3.5 Calculated heats of reaction for halogenation of hydrocarbons 

I I 

— C— H + X 2 ► — C— X + HX 

I I 



X 


A/f, kcal/mole 


F 


-115 


CI 


-27 


Br 


-10 


I 


12 



negative and moderate for chlorine and bromine, and positive for iodine (see 
Table 3-5). With fluorine, the reaction evolves so much heat that it is difficult 
to control, and products from cleavage of carbon-carbon as well as of carbon- 
hydrogen bonds are obtained. Indirect methods for preparation of fluorine- 
substituted hydrocarbons will be discussed later. Bromine is generally much 
less reactive toward hydrocarbons than chlorine, both at high temperatures 
and with activation by light. Nonetheless, it is usually possible to brominate 
saturated hydrocarbons successfully. Iodine is unreactive. 

As we have seen, the chlorination of methane does not have to stop with the 
formation of chloromethane, and it is possible to obtain the higher chlorin- 
ation products: dichloromethane (methylene chloride), trichloromethane 
(chloroform), and tetrachloromethane (carbon tetrachloride). In practice, 
all the substitution products are formed to some extent, depending on the 

CH 4 



► CHjCI 

chloro- 
methane 


— ► CH,CI 2 — 
dichloro- 
methane 


> CHCl., 

trichloro- 
methane 


► CCI 4 

tetrachloro 
methane 



chlorine-to-methane ratio employed. If monochlorination is desired, a large 
excess of hydrocarbon is advantageous. 

For propane and higher hydrocarbons, where more than one monosub- 
stitution product is generally possible, difficult separation problems may arise 
when a particular product is desired. For example, the chlorination of 
2-methylbutane at 300° gives all four possible monosubstitution products, 
[3], [4], [5], and [6]. On a purely statistical basis, we might expect the ratio of 
products to correlate with the number of available hydrogens at the various 
positions of substitution; that is, [3], [4], [5], and [6] would be formed in the 
ratio 6:3:2:1. However, in practice, the product composition is substanti- 
ally different, because the different kinds of hydrogens are not attacked at 
equal rates. Actually, the approximate ratios of the rates of attack of chlorine 
atoms on hydrogens located at primary, secondary, and tertiary positions are 
1.0:3.3:4.4 at 300°. These results indicate that dissociation energies of 
C— H bonds are not exactly the same but decrease in the order primary > 
secondary > tertiary. 



CH, 



I 
H 



CH, 

I 



sec 3.3 alkanes and their chemical reactions 61 

CH 3 CH 3 

Cl 2 I I 

— —> C1CH 2 -C-CH 2 -CH 3 + CH 3 -C-CH 2 -CH 2 C1 

300° 2 | 2 J ■* | i* 

H H 

l-chloro-2-methylbutane l-chloro-3-methylbutane 

[3] _ [4] 
CH, 



+ CH 3 -C— CHC1-CH, + CH 3 -C-CH 2 -CH 3 

H CI 

3-chloro-2-methylbutane 2-chIoro-2-methylbutane 
[5] [6] 



Bromine atoms are far more selective than chlorine atoms, and bromine 
attacks only tertiary hydrogens, and these not very efficiently. Thus, photo- 
chemical (light-induced) monobromination of 2-methylbutane proceeds 
slowly and gives quite pure 2-bromo-2-methylbutane. Bromine atoms might 

be expected to be more selective than chlorine atoms, because bond energies 

I .. I 

indicate that the process — C— H + :Br- ->•— C- + HBr is distinctly en- 

I •• I 

dothermic while the corresponding reaction with a chlorine atom is exothermic. 

In such circumstances it is not surprising to find that bromine only removes 

those hydrogens which are less strongly bonded to a carbon chain. 

Another reaction of commercial importance is the nitration of alkanes to 

give nitroalkanes. Reaction is usually carried out in the vapor phase at 

elevated temperatures using nitric acid or nitrogen tetroxide as the nitrating 

agent. All available evidence points to a radical-type mechanism for nitration 



RH + HNO, 



-425° 



RNO, 



H,0 



but many aspects of the reaction are not fully understood. Mixtures are 
obtained — nitration of propane gives not only 1- and 2-nitropropanes but 
nitroethane and nitromethane. 



CH 3 CH 2 CH 3 + HN0 3 



CH 3 CH 2 CH 2 N0 2 



1-nitropropane (25%) 

CH 3 CH 2 N0 2 

nitroethane (10%) 



I 
N0 2 

2-nitropropane (40%) 

CH 3 N0 2 

nitromethane (25%) 



In commercial practice, the yield and product distribution in nitration of 
alkanes are controlled as far as possible by the judicious addition of catalysts 
(e.g., oxygen and halogens) which are claimed to raise the concentration of 
alkyl radicals. The product mixtures are separated by fractional distillation. 



chap 3 alkanes 62 

3-4 cycloalkanes 

An important and interesting group of hydrocarbons, known as cycloalkanes, 
contain rings of carbon atoms linked together by single bonds. The simple 
unsubstituted cycloalkanes of the formula (CH 2 )„ make up a particularly 
important homologous series in which the chemical properties change in a 
much more striking way than do the properties of the open-chain hydro- 
carbons, CH 3 (CH 2 ) n _2CH 3 . The reasons for this will be developed with the 
aid of two concepts, steric hindrance and angle strain, each of which is simple 
and easy to understand, being essentially mechanical in nature. 

The conformations of the cycloalkanes, particularly cyclohexane, will be 
discussed in some detail, because of their importance to the chemistry of many 
kinds of naturally occurring organic compounds. 

Cyclohexane is a typical cycloalkane and has six methylene (CH 2 ) groups 
joined together to form a six-membered ring. Cycloalkanes with one ring 
have the general formula C„H 2 „ and are named by adding the prefix cyclo to 

H 2 C^ ^CH 2 

I I 

H 2 C \ C / CH 2 

H 2 

cyclohexane 

the name of the corresponding n-alkane having the same number of carbon 
atoms as in the ring. Substituents are assigned numbers consistent with their 
positions in such a way as to keep the sum of the numbers to a minimum. 

CH I 

2 C" ^CH 2 ..^ CH V 



II H 2 C CH 2 

^ **> h 2 V c— c'h 

\ 

C,H, 



CH 

I 
CH, 



1,4-dimethylcyclohexane l-ethyl-3-methylcyclopentane 

(not 3,6-dimethylcyclohexane) (not l-methyI-4-ethylcyclopentane) 

The substituent groups derived from cycloalkanes by removing one hydro- 
gen are named by replacing the ending -ane of the hydrocarbon with -yl to 
give cycloalkyl. Thus cyclohexane becomes cyclohexyl; cyclopentane, 
cyclopentyl; and so on. 

H 2 cA H2 
I CHCI 

h » c Vh, 

cyclopentyl chloride 
(or chlorocyclopentane) 



sec 3.4 cycloalkanes 63 

Frequently it is convenient to write the structure of a cyclic compound in an 
abbreviated form as in the following examples. Each line junction represents 
a carbon atom and the normal number of hydrogens on each carbon atom is 
understood. 





cyclobutane 2-methylcyclohexyl bromide cyclooctane 

( 1 -bromo-2-methylcyclohexane) 



A. PHYSICAL PROPERTIES OF CYCLOALKANES 

The melting and boiling points of cycloalkanes (Table 3-6) are somewhat 
higher than for the corresponding alkanes. The general "floppiness" of 
open-chain hydrocarbons makes them harder to fit into a crystal lattice 
(hence lower melting points) and less hospitable to neighboring molecules of 
the same type (hence lower boiling points) than the more rigid cyclic com- 
pounds. 

B. CONFORMATIONS OF CYCLOHEXANE 

If cyclohexane existed as a regular planar hexagon with carbon atoms at the 
corners, the C—C—C bond angles would be 120° instead of the normal 

Table 3*6 Physical properties of alkanes and cycloalkanes 



compounds 


bp,°C 


mp, °C 


di° 


propane 


-42 


-187 


0.580° 


cyclopropane 


-33 


-127 


0.689° 


«-butane 


- 0.5 


-135 


0.579" 


cyclobutane 


13 


- 90 


0.689" 


«-pentane 


36 


-130 


0.626 


cyclopentane 


49 


- 94 


0.746 


«-hexane 


69 


- 95 


0.659 


cyclohexane 


81 


7 


0.778 


H-heptane 


98 


- 91 


0.684 


cycloheptane 


119 


- 8 


0.810 


«-octane 


126 


- 57 


0.703 


cyclooctane 


151 


15 


0.830 


«-nonane 


151 


- 54 


0.718 


cyclononane 


178 


11 


0.845 



" At -40°. 

* Under pressure. 



chap 3 alkanes 64 



«* 



A 
# 



4 HjI 



ti ft. 



Figure 3-4 Two conformations of cyclohexane with 109.5° bond angles 
(hydrogens omitted). 

valence angle of carbon, 109.5°. Thus, a cyclohexane molecule with a planar 
structure could be said to have an angle strain of 10.5° at each of the carbon 
atoms. Puckering of the ring, however, allows the molecule to adopt confor- 
mations that are free of angle strain. 

Inspection of molecular models reveals that there are actually two extreme 
conformations of the cyclohexane molecule that may be constructed if the 

Figure 3-5 Boat form of cyclohexane showing interfering and eclipsed 
hydrogens. Top, scale model; center, ball-and-stick models ; bottom, sawhorse 
representations. 







— ♦ 



-ortydwscni 



r»>-aa>Mi!s 





side view 



end view 



sec 3.4 cycloalkanes 65 

carbon valence angles are held at 109.5°. These are known as the "chair" 
and "boat" conformations (Figure 3-4). These two forms are so rapidly 
interconverted at ordinary temperatures that they cannot be separated. It is 
known, however, that the chair conformation is considerably more stable 
and comprises more than 99 % of the equilibrium mixture at room tempera- 
ture. 

The higher energy of the boat form is not due to angle strain because all 
the carbon atoms in both forms have their bond angles near the tetrahedral 
angle of 109.5°. It is caused, instead, by relatively unfavorable interactions 
between the hydrogen atoms around the ring. If we make all the bond angles 
normal and orient the carbons in the ring to give the extreme boat con- 
formation shown in Figure 3-5, we see that a pair of 1,4 hydrogens (the 
so-called flagpole hydrogens) have to be so close together (1.83 A) that they 
repel one another. This is an example of steric hindrance. 

There is still another factor which makes the extreme boat form unfavorable ; 
namely, that the eight hydrogens around the "sides" of the boat are eclipsed, 
which brings them substantially closer together than they would be in a stag- 
gered arrangement (about 2.27 A compared with 2.50 A). This is in striking 
contrast with the chair form (Figure 3-6) for which adjacent hydrogens are 
seen to be in staggered positions with respect to one another all the way 
around the ring. The chair form is therefore expected to be the more stable of 
the two. Even so, its equilibrium with the boat form produces inversion about 
10 6 times per second at room temperature. If you make a molecular model of 

less stable 

cyclohexane you will find that the chair form has a considerable rigidity and the 
carbon-carbon bonds have to be slightly bent in going to the boat form. You 
will find that the boat form is extremely flexible and even if the bond angles 
are held exactly at 109.5°, simultaneous rotation around all the carbon- 
carbon bonds at once permits the ring to twist one way or the other to reduce 
the repulsions between the flagpole hydrogens and between the eight hydro- 
gens around the sides of the ring. These arrangements are called twist-boat 
(sometimes skew-boat) conformations (Figure 3-7) and are believed to be only 
about 5 kcal less stable than the chair form. 

It will be seen that there are two distinct kinds of hydrogens in the chair 
form of cyclohexane. Six are almost contained by the "average" plane of the 
ring (called equatorial hydrogens) and three are above and three below this 
average plane (called axial hydrogens). This raises an interesting question in 
connection with substituted cyclohexanes : For example, is the methyl group 
in methylcyclohexane equatorial or axial ? 



a CH 3 

^ T^ fast . t , 

£^J - \ ^CH 3 

(axial) (equatorial) 



chap 3 alkanes 66 




Figure 3-6 Chair form of cyclohexane showing equatorial and axial hydro- 
gens. Top left, scale model; bottom, ball-and-stick model; top right, sawhorse 
representations. Note that all the axial positions are equivalent and all the 
equatorial positions are equivalent. 



There is considerable evidence which shows that the equatorial form of 
methylcyclohexane predominates in the equilibrium mixture (K~15), and 
the same is generally true of all monosubstituted cyclohexane derivatives. 
The reason can be seen from scale models which show that a substituent 
group has more room in an equatorial conformation than in an axial con- 
formation (see Figure 3-8). The bigger the substituent, the greater the tendency 
for it to occupy an equatorial position. 

The forms with axial and equatorial methyl are interconverted about 10 6 



Figure 3-7 Drawings of the twist-boat conformations of cyclohexane. 



sec 3.4 cycloalkanes 67 




Figure 3-8 Scale models of equatorial and axial forms of the chair form of 
bromocyclohexane. 



times/second at room temperature. The rate decreases as the temperature is 
lowered. If one cools the normal mixture of chlorocyclohexane conform- 
ations dissolved in a suitable solvent to very low temperatures (—150°), the 
pure equatorial conformation crystallizes out. This conformation can then be 
dissolved in solvents at — 1 50° and, when warmed to — 60°, is converted to the 
equilibrium mixture in a few tenths of a second. However, the calculated 
half-time of the conversion of the equatorial to the axial form is 22 years at 
-160°. 



C. OTHER CYCLOALKANE RINGS 

The three cycloalkanes with smaller rings than cyclohexane are cyclopentane, 
cyclobutane, and cyclopropane, each with bond angles less than the tetrahedral 
value of 109.5°. If you consider a carbon-carbon double bond as a two- 
membered ring, then ethene, C 2 H 4 , is the simplest cycloalkane ("cyclo- 
ethane") and, as such, has carbon bond angles of 0° and, therefore, a very 
large degree of angle strain. 



H 2 C CH 2 

\ / 

H 2 C-CH 2 


H 2 C — CH 2 

1 1 
H 2 C — CH 2 


CH 2 
/ \ 
H 2 C— CH 2 


CH 2 =C 


cyclopentane 


cyclobutane 


cyclopropane 


ethene 


bond angle if planar: 108° 


90° 


60° 


0° 


angle strain 1 5° 
(109.5° -bond angle): 


19.5° 


49.5° 


109.5° 



Table 37 shows how strain decreases stability and causes the heat of com- 
bustion per methylene group (or per gram) to rise. 

The idea that cyclopropane and cyclobutane should be strained because 
their C—C—C bond angles cannot have the normal tetrahedral value of 
109.5° was advanced by Baeyer in 1885. It was also suggested that the diffi- 



chap 3 alkanes 
Table 3-7 Strain and heats of combustion of cycloalkanes 







angle strain 




heat of 








at each CH 2 


heat of 


combustion 


total 


cycloalkane, 




for planar 


combustion," 


per CH 2 , AH/n, 


strain, 6 


(CH 2 )„ 


n 


molecules, deg 


AH, kcal/mole 


kcal 


kcal/mole 


ethene 


2 


109.5 


337.23 


168.6 


22.4 


cyclopropane 


3 


49.5 


499.83 


166.6 


27.6 


cyclobutane 


4 


19.5 


655.86 


164.0 


26.4 


cyclopentane 


5 


1.5 


793.52 


158.7 


6.5 


cyclohexane 


6 


(10.5) c 


944.48 


157.4 


0.0 


cycloheptane 


7 


(19.0) c 


1108.2 


158.3 


6.3 


cyclooctane 


8 


(25.5) c 


1269.2 


158.6 


9.6 


cyclononane 


9 


(30.5) c 


1429.5 


158.8 


11.2 


cyclodecane 


10 


(34.5) c 


1586.0 


158.6 


12.0 


cyclopentaclecane 


15 


(46.5)' 


2362.5 


157.5 


1.5 


open-chain, «-alkane 


00 






157.4 





" For gaseous hydrocarbons to give liquid water at 25°, datafromS.Kaarsemakerand J. Coops, 
Rec. Trav. Chim. 71, 261 (1952), and J. Coops, H. Van Kamp, W. A. Lambgrets, B. J. Visser, and 
H. Dekker, Rec. Trav. Chim. 79, 1226 (1960). 

b Calculated by subtracting (« X 1 57.4) from the observed heat of combustion. 

c Angle strain calculated for planar ring as per the Baeyer theory. The strain that is present in 
the C7 to Cio compounds is not the result of angle strain (the molecules are puckered) but of 
eclipsing or interfering of hydrogen atoms. 

culties encountered up to that time in synthesizing cycloalkane rings from C 7 
upward was the direct result of the angle strain which would be expected if 
the large rings were regular planar polygons (see again Table 3-7). 

We now know that the Baeyer strain theory cannot be applied to large 
rings because cyclohexane and the higher cycloalkanes have puckered rings 
with normal or nearly normal bond angles. Much of the difficulty in synthe- 
sizing large rings from open-chain compounds is due to the low probability 
of having reactive groups on the two fairly remote ends of a long hydrocarbon 
chain come together to effect cyclization. Usually, coupling of reactive groups 
on the ends of different molecules occurs in preference to cyclization, unless 
the reactions are carried out in very dilute solutions. 

For cyclopentane, a planar structure would give bond angles of 108°, 
very close to the natural bond angle of 109.5°. Actually, the angle strain is 
believed to be somewhat greater than 1.5° in this molecule; the eclipsing of all 
of the hydrogens causes the molecule to distort substantially even though this 
increases the angle strain. Cyclobutane is also not completely flat for the 
same reason. (It should be remembered that molecules such as these are in 
vibrational motion at all times and the shapes that have been described refer 
to the mean atomic positions averaged over a period of time corresponding to 
several vibrations.) 



D. CHEMICAL PROPERTIES OF CYCLOALKANES 

We have already observed how strain in the small-ring cycloalkanes affects 
their heats of combustion. We can reasonably expect other chemical properties 



sec 3.4 cycloalkanes 69 

also to be affected by ring strain, and indeed cyclopropane and cyclobutane 
are considerably more reactive than saturated, open-chain hydrocarbons. In 
fact, they undergo some of the reactions which are typical of compounds with 
carbon-carbon double bonds, their reactivity depending on the degree of angle 
strain and the vigor of the reagent. 

The result of these reactions is always opening of the ring by cleavage of a 
C— C bond to give an open-chain compound having normal bond angles. 
Relief of angle strain may therefore be considered to be an important part 
of the driving force of these reactions. A summary of a number of ring- 
opening reactions is given in Table 3-8. Ethene is highly reactive, while cyclo- 
propane and cyclobutane are less so (in that order). The C— C bonds of the 
larger, relatively strain-free cycloalkanes are inert, so that these substances 
resemble the «-alkanes in their chemical behavior. Substitution reactions of 
these cycloalkanes are generally less complex than those of the corresponding 
alkanes because there are fewer possible isomeric substitution products. 
Thus, cyclohexane can give only one monochlorination product while 
«-hexane can give three. 



E. Cis-Trans ISOMERISM OF SUBSTITUTED CYCLOALKANES 

The form of stereoisomerism (isomerism caused by different spatial arrange- 
ments) called geometrical isomerism or cis-trans isomerism was discussed in 
the preceding chapter. This type of isomerism arises when rotation is prevented 
by, for example, the presence of a double bond. A ring prevents rotation 
equally well and we find that cis and trans isomers can also exist with ap- 
propriately substituted cycloalkanes. Thus, when a cycloalkane is disub- 



CH 2 

/ \ 2 

CH 3 CH-CHCH 3 

1,2-dimethylcyclopropane 



stituted at different ring positions, as in 1,2-dimethylcyclopropane, two 
isomeric structures are possible according to whether the substituents are 
both situated above (or both below) the plane of the ring {cis isomer), or one 
above and one below (trans isomer), as shown in Figure 3-9. 

The cis and trans isomers of 1,2-dimethylcyclopropane cannot be inter- 
converted without breaking one or more bonds. One way of doing this is to 
break open the ring and then close it again with a substituent on the opposite 
side from where it started. Alternatively, the bond to the substituent (or the 
hydrogen) can be broken and reformed on the opposite side of the ring. 
Examples of both processes will be discussed in later chapters. 

Cis and trans isomers of cyclohexane derivatives have the additional 
possibility of different conformational forms. For example, 4-?-butylcyclo- 
hexyl chloride can theoretically exist in four stereoisomeric chair forms, 



chap 3 alkanes 70 



Table 3-8 Reactions of cycloalkanes, (CH 2 )„ 



. o 










is 


JH 




u 

o 


in 


u X 


<D 


.s 


c 


c 


(►, u 


_c 








o X 





















i s 












^g 


J-l 

C 




CD 


u a 










i » 










° S 


l* 




sh 


© 


i— ( ra 


0J 




<D 


o 


£3 


.g 


C- 


,g 


(N 


O ,£! 










ID 

, (3 


>1 


^ 


w 




is. 

u o 




T3 


t-- 

.g 


© 

IN 


>l fcc 


fi 


I-< 






o a 










v 










c ^ 








& 


<d ^ 










5 II 








_>> 


^ 


_>, 


£E 
-o o 


•9 


t3 


•3 






CD 




a p 


CD »- 
> 


t-< 


^ 


V- 






X 










m 










o 








« « ffl 


•N O 


X I 


rs 




£ -5 £ 


m 1 ^V, 


°, vq, 


fl 1 fO 




Ux B /U 


^ 


^x>o 


^ 


E 


+ 


-t. 










_0 


T 






























■*j 
















o 












z 




« 
















HI 


I 














h 












u 


o 


o 


£ 




« 


c/3 


c 


+ 




+ 


£ 
+ 


+ 






£ i £ 










O-r-U 










\r/ 










'*~^j 










X 










u 



















sec 3.4 cycloalkanes 71 






\ ••■•.-.•. 


: :• #^li 


.-■.-■?ISS 



Figure 3*9 Ball-and-stick models of cis and /raws isomers of 1,2-dimethyl- 
cyclopropane. 



[7], [8], [9], and [10]. 



I 

H 3 L H c , 

[7] rrani [8] 




CI 



I 
H,C— C-CH, 




*> C ^ e y^ T^H ^^- H^~^\ 5 C1 




CH 

H 3 C y\ H 

[9] cis [10] 

Structures [7] and [8] have the substituents trans to one another, but in 
[7] they are both equatorial while in [8] they are both axial. Structures [9] 
and [10] have the cis relationship between the groups, but the /-butyl and 
chlorine are equatorial-axial in [9] and axial-equatorial in [10]. 7-Butyl groups 
are very large and bulky and much more steric hindrance results when a 
?-butyl group is in an axial position than when chlorine is in an axial position 
(Figure 3-10). Hence the equilibrium between the two conformational forms 
of the trans isomer strongly favors structure [7] over structure [8] because 
both ?-butyl and chlorine are equatorial. For the cis isomer, structure [9] is 
favored over [10] to accommodate the 7-butyl group in the equatorial position. 

When there are two substituents in the cis-1,4 arrangement on a cyclo- 
hexane ring, neither of which will go easily into an axial position, then it 



chap 3 alkanes 72 



X C "•■- "H 



Figure 3-10 1,3 Interactions in a cyclohexane ring 
with an axial 7-butyl group. 



appears that the twist-boat conformation (Section 3-4B) is most favorable 
(Figure 3-11). 



summary 

Alkanes are hydrocarbons possessing only single bonds. The open-chain 
alkanes have the formula C„H 2 „ + 2 . The 1UPAC names for alkanes are based 
on the longest continuous carbon chain with substituents being indicated by 
their position along the chain. The alkane names from C x to C 10 are methane, 
ethane, propane, butane, pentane, hexane, heptane, octane, nonane, and 
decane. Structural isomerism appears at C 4 , there being two compounds of 
formula C 4 H 10 — the continuous-chain compound CH 3 CH 2 CH,CH 3 (butane) 

CH 3 

and the branched-chain compound I (2-methylpropane or 

CH 3 — CH — CH 3 

isobutane). The larger the number of carbon atoms in a continuous-chain 

alkane, the larger the number of branched-chain isomers of it that will exist. 

Alkyl groups are obtained by removing a hydrogen atom from an alkane, 
and structural isomerism appears here at the C 3 level. The group CH 3 — 
CH 2 — CH 2 — is called the «-propyl group and CH 3 — CH— CH 3 the isopropyl 
group. ' 

The physical properties of the alkanes show a smooth gradation. At room 
temperature the C t to C 4 compounds are gases, the C 5 to C 18 continuous- 
chain (normal) alkanes are liquids, and higher-molecular-weight compounds 
are solids. The normal alkanes which are liquids often have branched-chain 
isomers which are solids. All alkanes are less dense than water and all are 
immiscible with water. 

Petroleum is a complex mixture of hydrocarbons which can be separated 
into fractions, according to volatility: natural gas, gasoline, kerosene, diesel 



Figure 3-11 Twist-boat conformation of c/,s-l,4-di-?-butylcyclohexane. 



summary 73 

oil, lubricating oils and waxes, and residual material (asphalt). The heats of 
combustion of the alkanes with the same molecular weights in the gasoline 
fraction are all very close but their efficiencies in producing power in high- 
compression internal combustion engines vary widely with structure. The 
normal alkanes, which knock in the cylinder, have low octane ratings; the 
branched alkanes, which burn less rapidly, have high octane ratings. 

In addition to combustion, alkanes undergo substitution reactions with 
halogens or nitric acid. These three reactions are illustrated using propane 
as the alkane: 



°' 2 > CH3CHCICH3 + CH 3 CH 2 CH 2 C1 (+HCI) 

N0 2 
HNO3 I 

* CH3CHCH3 + CH 3 CH 2 CH 2 N0 2 (+H 2 0) 

With higher alkanes, more complex mixtures of substitution products result, 
although the major products are usually those in which a tertiary hydrogen has 
been replaced. 

The cycloalkanes have similar physical and chemical properties to those of 
the open-chain alkanes except that the small-ring compounds such as cyclo- 
CH 2 

propane, / \ are more reactive because of bond-angle strain. 

H 2 C — CH 2 

Cyclohexane exists in two principal conformations that are rapidly inter- 
converted, the more stable and rather rigid chair form and the less stable 
and flexible twist-boat form. The twelve carbon-hydrogen bonds in the chair 



^=7 -= b^ 

chair form twist-boat form 

of cyclohexane of cyclohexane 

form are of two types, six axial bonds parallel to the vertical axis of the ring 
and six equatorial bonds pointing out from the equator of the ring. Of the 
two kinds of positions, the equatorial provides more room for bulky substi- 
tuent groups and, therefore, a substituent group will normally prefer to take 
an equatorial position. Some of the principal conformational forms of 
methylcyclohexane are shown here (all are in rapid equilibrium, with the form 
on the far right being the most stable). 

H 

a twist-boat form chair form chair form 

methyl group axial methyl group equatorial 



chap 3 alkanes 74 

Geometrical isomers can exist with appropriately substituted cycloalkanes 
since rotation about the C— C bonds in the ring is prevented by the ring itself. 
Cis isomers have the substituent groups on the same side of the ring and trans 
isomers on the opposite side. A number of forms are possible in cycloalkanes 
because of the combination of geometrical isomerism and conformational 
equilibria. 

exercises 

3-1 Name each of the foDowing hydrocarbons by the IUPAC system and in 
example e as an alkyl-substituted methane. 

CH 3 CH 3 

\ / 

a. CH-CH 2 -CH 2 -CH 
/ 2 \ 

CH 3 CH 3 

CH, CH, 

I I 

b. CH 3 -C-CH 2 -CH-CH 3 



CH 3 CH 2 .CH 3 

c. CH-CH 2 -CH 
/ \ 

CH 3 CH 3 

CH 3 — CH 2 
\ 

d. CH-CH, 
/ 

CH 3 — CH 2 



e. (CH 3 — CH 2 -CH 2 ^C 



H 3 C CH 3 
CH 



CH 2 CH 3 



I / 

CH 3 — CH 2 — CH 2 -CH 2 -CH— CH 2 -CH 2 -CH 

CH 3 



3-2 Write the structures of the eight branched-chain isomers of heptane 
CH 3 CH 2 CH 2 CH 2 CH 2 CH 2 CH 3 . Name each by the IUPAC system. 

3-3 What is the IUPAC name for (a) triethylmethane, (b) hexamethylethane ? 

3-4 Each of the names given below violates the rules of organic nomenclature. 
Supply the correct name in each case. 

a. 2,3-diethylbutane 

b. 2,4,4-trichlorohexane 

c. 1,4-diisopropylbutane 

d. 2,4,5-trimethyl-3-«-propylheptane 



exercises 75 

3-5 Fill in the appropriate prefix in the names given below and draw the structural 
formula in each case. 



a 


2,3- 


_ methylpentane 


h. 


1,1,1- 


chloroethane 


c. 


1,2,3,4,5,6- 


iodohexane 



3-6 Write structures for all seventeen possible monochlorohexanes and name them 
by the IUPAC system. 

3-7 Use the data of Tables 3-3 and 3-4 to estimate the boiling points of tetra- 
decane, heptadecane, 2-methylhexane, and 2,2-dimethylpentane. 

3-8 Is "neoheptane" an unambiguous name? Explain. 

3-9 Write structural formulas for each of the following and name each by the 
IUPAC system: 

a. ?-butyl-isobutyl-,s-butyl-M-butylmethane 

b. isononane 

c. the monochloropentane isomers ; also name each as best as you can as 
an alkyl chloride. 

3-10 Calculate AH for the following reactions in the vapor state at 25°: 

a. 2 CH 4 + 7 Cl 2 > CCh-C0 3 + 8 HC1 

b. CH 3 CH 3 + 5 2 * 2C0 2 + 3H 2 

c. CH 3 CH 3 + H 2 > 2CH 4 

d. CH 3 CH 3 + Br 2 > 2CH 3 Br 

e. CH 4 + 2C1 2 ► C(g) + 4HCl 

3-11 a. Would the calculated AHin Exercise 3-10e be greater or less if C (solid) 
were the reaction product ? Explain. 
b. What are the implications of the heats of reaction determined in Exercise 
3-10c and d to the "saturated" character of ethane? 

3-12 The C— F bond energy in Table 2-1 was computed from recent thermo- 
chemical studies of the vapor-phase reaction, 

CHU + 4 F 2 — ► CF 4 + 4 HF AH = -460 kcal 

Show how the AH value for this reaction may be used to calculate the energy 
of the C— F bond if all the other required bond energies are known. 

3-13 Investigate the energetics (AH) of possible chain mechanisms for the light- 
induced monobromination of methane and make a comparison with those 
for chlorination. What are the prospects for iodination of methane? 

3-14 The heat of combustion of cyclopropane (CH 2 ) 3 to give carbon dioxide and 
water vapor is 468.6 kcal. Show how this value can be used to calculate the 
average C— C bond energies of cyclopropane. 



chap 3 alkanes 76 

3-15 Combustion of a pure sample of a gaseous alkane produced a quantity of 
carbon dioxide whose weight and volume (under the same conditions) were 
exactly three times that of the gaseous alkane. What is the latter's formula ? 

3.16 Combustion of natural gas is generally a "cleaner" process (in terms of 
atmospheric pollution) than combustion of either gasoline or fuel oil. Explain 
why this is so. 

3-17 Write a mechanism in harmony with that usually written for hydrocarbon 
chlorination which would lead to production of hexachloroethane as in 
Exercise 3- 10a. (This reaction is used for commercial production of hexa- 
chloroethane.) 

3-18 Show the configurations of all of the possible cis-trans isomers of the following 
compounds : 

a. 1,2,3-trimethylcyclopropane 

b. 1,3-dichlorocyclopentane 

c. 1,1,3-trimethylcyclohexane 

3-19 Would you expect cis- or ?ra««-l,2-dimethylcyclopropane to be the more 
stable ? Explain. 

3-20 Write expanded structures showing the C— C bonds for each of the following 
condensed formulas. Name each substance by an accepted system. 

a. (CH 2 ) 10 

b. (CH 2 ) 5 CHCH 3 



c. (CH 3 ) 2 C(CH 2 ) 6 CHC 2 H 5 

d. The isomers of trimethylcyclobutane 

e. (CH 2 ) 6 CHCH 2 C(CH 3 ) 2 CH 2 C1 
/. [(CH 2 ) 2 CH] 2 C(CH 3 )C 2 H 5 

3-21 Draw structural formulas for all C 4 H 8 and all C 5 Hi compounds that con- 
tain a ring. Designate those that exist in cis and trans forms. 

3'22 The energy barrier for rotation about the C— C bond in ethane is about 3 
kcal, which suggests that the energy required to bring one pair of hydrogens 
into an eclipsed arrangement is 1 kcal. Calculate how many kilocalories the 
planar form and extreme boat form of cyclohexane would be unstable rela- 
tive to the chair form on account of H— H eclipsing interactions alone. 

3-23 Use the sawhorse convention and draw all the possible conformations of 
cyclohexyl chloride with the ring in the chair and in the boat forms. Arrange 
these in order of expected stability. Show your reasoning. 

3-24 Formation of a cycloalkane (CH 2 )„ by reactions such as Br-fCH 2 -KZnBr 
-» (CH 2 )„ + ZnBr 2 occurs in competition with other reactions such as 



, exercises 77 

2 Br-f(CH 2 ^-„ZnBr -> Br-eCH 2 )„(CH 2 >„ ZnBr + ZnBr 2 . Explain why 
cyclization reactions of this kind carried out in dilute solutions are likely to 
give better yields of (CH 2 )„ than in concentrated solutions. 

3-25 Use the data of Table 3-7 and other needed bond energies to calculate AH 
for the following reaction in the vapor state at 25° with n = 3, 4, and 5. 

(CH 2 )„ > CH 3 (CH 2 )„- 3 CH=CH 2 

3-26 What can you conclude about the stability of the cycloalkanes with n = 3, 4, 
and 5 with respect to corresponding open-chain compounds with double 
bonds ? 

3-27 Use the heats of combustion (to liquid water) given in Table 3-7 and appro- 
priate bond energies to calculate AH (vapor) for ring opening of the cyclo- 
alkanes with bromine over the range n =2 to n — 6: 

(CH 2 )„ + Br 2 * (CH 2 )„_ 2 (CH 2 Br) 2 

3-28 Show how the reactions described in Table 3-8 could be used to tell whether 
a hydrocarbon of formula C 4 H 8 is methylcyclopropane, cyclobutane, or 
1-butene (CH 3 CH 2 CH=CH 2 ). Write equations for the reactions used. 

3-29 Draw the possible chair conformations of trans- and cw-l,3-dimethylcyclo- 
hexane. Is the cis or the trans isomer likely to be the more stable? Explain. 

3-30 An empirical rule known as the von Auwers-Skita rule was used to assign 
configurations of pairs of cis and trans isomers in cyclic systems at a time 
when cis isomers were thought to be always less stable than trans isomers. 
The rule states that the cis isomer will have the higher boiling point, density, 
and refractive index. However, the rule fails for 1,3-disubstituted cyclohexanes, 
where the trans isomer has the higher boiling point, density, and refractive 
index. Explain how the von Auwers-Skita rule might be restated to include 
such 1,3-systems. 

3-31 Would you expect cyclohexene oxide to be more stable in the cis or trans 
configuration ? Give your reasons. 



.O cyclohexene oxide 



3-32 Write structural formulas for substances (one for each part) which fit the 
following descriptions. Make sawhorse drawings of the substances where 
conformational problems are involved. 

a. a compound of formula C 4 H 8 which reacts slowly with bromine and 
sulfuric acid but not with potassium permanganate solution 

b. the most highly strained isomer of C 5 Hi 

c. the possible products from treatment of l-ethyl-2-methylcyclopropane 
with bromine 




chap 3 alkan.es 78 

d. the least stable chair and the least stable boat conformations of trans- 

1 ,4-dichlorocyclohexane 

e, the most stable geometrical isomer of l,3-di-?-butylcyclobutane 

/. a compound with a six-membered ring which is most stable with the 

ring in a boat form 
g. the most stable possible conformation of fra«i , -l,3-di-Nbutylcyclo- 

hexane 



! •*.*"-. r-V*- ' V' , 



'^■l!l''---.-''l'm 



tiic'M 



chap 4 alkenes 81 

In the early days of organic chemistry, when it was found that the alkenes, 
but not the alkanes, readily undergo addition reactions with substances such 
as halogens, hydrogen halides, sulfuric acid, and oxidizing agents, the chemi- 
cal affinity of alkanes was said to be " saturated " while that of the alkenes 
was said to be " unsaturated." Now, even though we recognize that no chemi- 
cal entity (even the noble gases such as helium and xenon) can surely be 
classified as saturated, the description of alkanes and alkenes as saturated and 
unsaturated is still commonly used. However, in place of a nebulous chemical 
affinity, we ascribe the unsaturation of alkenes to the ease of cleaving half 
of a carbon-carbon double bond in an addition reaction. Additions occur with 
alkenes much more easily than with alkanes because (1) the carbon-carbon 
bonds of a double bond are individually weaker (more strained) than a normal 
carbon-carbon single bond and (2) the double-bond electrons are generally 
more accessible than single-bond electrons to an attacking reagent (see 
Section 2-6). 

The great variety and specificity of the addition reactions that compounds 
with double bonds undergo make these substances extremely important as 
intermediates in organic syntheses. We have already examined two of these 
reactions (addition of halogens and hydrogen) in connection with our study 
of ethene, the simplest alkene, in Chapter 2. 



4-1 nomenclature 

Open-chain alkenes containing one double bond have the general formula 
C„H 2 „ and are sometimes called olefins. According to the IUPAC system for 
naming alkenes, the longest continuous chain containing the double bond is 
given the name of the corresponding alkane with the ending -one changed 
to -ene. This chain is then numbered so that the position of the first carbon 
of the double bond is indicated by the lowest possible number. 



CH,CH,CH=CH, CH,-CH,-CH, 



2CH 



1-butene 3-propyl-l-heptene 

(not 3-butene) (the dotted lines indicate longest continuous 

chain containing the double bond) 

Other, less systematic names are often used for the simpler alkenes. By 
one method, alkenes are named as substituted ethylenes. This nomenclature, 



(CH 3 ) 2 C = C(CH 3 ) 2 C1 2 C = CHC1 

ethylene tetramethylethylene trichloroethylene 

based on the older name " ethylene," is given here because, even though not 
in accord with modern practices, it has been widely used in the literature. A 



chap 4 alkenes 82 

little reflection will show that attempts to name alkenes as derivatives of pro- 
pylene (propene) or butylene (as will be seen, there are four open-chain C 4 H 8 
isomers) will require special rules or be hopelessly ambiguous. 

The hydrocarbon groups derived from alkenes carry the suffix -enyl, as in 
alkenyl, and numbering of the group starts with the carbon atom with the 
free bond : 



CH 3 ~CH = CH— CH 2 - 

2-bntenyl 3-butenyl 

However, there are a few alkenyl groups for which trivial names are commonly 
used in place of systematic names. These are vinyl, allyl, and isopropenyl 
groups : 

CH 3 

I 

CH 2 =CH- CH 2 =CH— CH 2 — CH 2 =C — 

vinyl allyl isopropenyl 

(ethenyl) (2-propenyl) (1-methylethenyl) 



Cycloalkenes with double bonds in the ring (endocyclic double bonds) are 
named by the system used for the open- chain alkenes, except that the num- 
bering is always started at one of the carbons of the double bond and con- 
tinued on around the ring through the double bond so as to keep the sum of the 
index numbers as small as possible. 

More complex nomenclature systems are required when the double bond 
is exocyclic to the ring, especially if a ring carbon is one terminus of the 
double bond. Usually the parent compounds of this type are called methylene- 
cycloalkanes. 



H 2 C6 s iCH 2 H 2 C-C 

I. J II 



CUJCH 
H CH 3 



H 3 cr "c/ \ 



1 ,3-dimethylcyclohexene methylenecyclobutane 

(not 1,5-dimethylcydohexene) 

Many compounds contain two or more double bonds and are known as 
alkadienes, alkatrienes, alkatetraenes, and so on, the suffix denoting the num- 
ber of double bonds. The location of each double bond is specified by appro- 
priate numbers. 

CH 2 =C=CH-CH 3 CH 2 =CH-CH=CH 2 CH 2 =C = C = CH 2 

1.2-butadiene 1 ,3-buladiene 1,2.3-butatriene 



sec 4.2 isomerism in C 4 H 8 compounds 83 

A further classification is used according to the relationships of the double 
bonds, one to the other. Thus, 1,2-alkadienes and similar substances are 
said to have cumulated double bonds. 1,3-Alkadienes and other compounds 

\ / 

CH 2 =C=CH 2 C = C = C 

/ \ 

allene cumulated double bonds 

(propadiene) 

with alternating double and single bonds are said to have conjugated double 
bonds, and this arrangement leads, as we shall see in Chapter 6, to com- 
pounds having rather special properties. 

CH 3 

\ I I / I 

C = C — C — C CH,= CH— C=CH 2 

/ \ 

1,3-pentadiene conjugated 2-methyl-l,3-butadiene 

double bonds (isoprene) 



Compounds with double bonds that are neither cumulated nor conjugated 
are classified as having isolated double-bond systems. 

\ I I I / 
CH,=CH — CH 2 — CH=CH, C = C-eC^rC = C 

/ | \ 

1.4-penladiene isolated double bond system (n £ 1) 



4-2 isomerism in C + H 8 compounds 

In the homologous series of alkanes, isomerism first appears at the C 4 level, 
two compounds of formula C 4 H 10 being known. These are structural 
isomers: 



CH, 
I 
CH 3 — CH 2 — CH 2 — CH 3 CH 3 — CH-CH 3 

butane, bp —0.5 2-methylpropane, bp —12° 



There are in all six isomers of C 4 H 8 . Some are structural isomers and some 
stereoisomers (see Section 2-6B). Their boiling points and general physical 
properties are similar to those of butane and 2-methylpropane. Four of these 
compounds react quickly with bromine; one reacts slowly, and one not at all. 
The latter two compounds must be methylcyclopropane and cyclobutane, 
respectively (Section 3-4D), and these compounds are cycloalkanes, not 



chap 4 alkenes 84 

alkenes. Note that the 2-butene structure is the only one that can exist in 



/CH 2 
CHj-HC | 

^CH 2 


H 2 C-CH 2 

1 1 
H 2 C — CH 2 


methylcyclopropane. bp 4° 


cyclobutane. bp 1 1 


CH 3 — CH 2 — CH=CH 2 


CH 3 CH 3 

\ / 

c=c 

/ \ 

H H 




l-butenc. bp —6.3° 


ra-2-butene, bp 3.7° 


CH 3 H 

\ / 

c=c 

/ \ 

H CH 3 


CH 3 
1 
CH 3 -CH = CH 2 



trans-2-butene, bp 0.9° 2-methylpropene, bp - 6° 

two different configurational arrangements. The other two isomers, 1-butene 
and 2-methylpropene, have at least one carbon atom of the double bond with 
identical groups attached to it. Thus, a rotation about the double bond, even 
if it could occur, would produce an identical arrangement. 

It is worth reviewing once again the meanings of the terms structure, 
configuration, and conformation (Sections 2-2 and 2-6B). Of the six known com- 
pounds of formula C 4 H 8 , there are five different structures. These are cyclo- 
butane, methylcyclopropane, 1-butene, 2-butene, and 2-methylpropene. One 
of these structures, 2-butene, has two different stable configurations or spatial 
arrangements. All of these substances have many different possible conforma- 
tions because rotation can occur to at least some degree about their single 
bonds. Putting it another way, the C 4 H 8 compounds illustrate structural 
isomerism, geometrical isomerism, and conformational variation. Structural 
and geometrical isomers (but not conformational isomers), because of their 
stability to interconversion and their somewhat different physical constants, 
can be separated by physical techniques such as fractional distillation or, 
better, by chromatography (Section 7T). 



4- 3 cis and trans isomers 

By convention, the configuration of complex alkenes is taken to correspond to 
the configuration of the longest continuous chain as it passes through the 
double bond. Thus the following compound is 4-ethyl-3-methyl-/ran,s'-3- 
heptene, despite the fact that two identical groups are cis with respect to each 

CH3 CH 2 CH 2 CH3 



;-CH 2 


CH 2 


\ 


/ 


c= 


= C 


/ 


\ 


H 3 C 


CH 2 



CH 2 -CH 3 

4-ethyl-3-methyl-/ra«5-3-heptene 



sec 4.3 cis and trans isomers 85 



repulsions between 
methyl groups 



.)(". 



/C c' 

CH * c c CHs 

/ \ 

H H 



Figure 4-1 Repulsive interactions between the methyl groups of cis-sym- 
di-/-butylethylene (2,2,S,S-tetramethyl-m-3-hexene). 



other, because the longest continuous chain is trans as it passes through the 
double bond. 

The trans isomers of the simple alkenes are usually more stable than the 
corresponding cis isomers. The methyl groups in trans-2-butene are far 
apart; in cz's-2-butene, they are much closer to one another. Scale models, 
which reflect the sizes of the methyl groups, indicate some interference be- 
tween the methyl groups of the cis isomer. The cis alkenes with large groups 
have very considerable repulsive interactions (steric hindrance) between the 
substituents, and are much less stable than the corresponding trans isomers 
(see Figure 4-1). 

The generally greater stability of trans over cis isomers (see, however, 
Section 2-6B) is reflected in their lower heats of combustion. Table 4-1 com- 
pares the heats of combustion and the boiling and melting points of some cis 
and trans isomers. The data also reveal that trans isomers tend to have higher 
melting points and lower boiling points than cis isomers. Although the 
differences are not large, they may be of some help in assigning configurations. 
When electron-withdrawing groups such as halogens are attached to the 



Table 4-1 Comparison of properties of cis and trans isomers 















heat of 










bp, 


mp, 


combustion, 


alkene 


formula 






°C 


°C 


AH, kcal 


cw-2-butene 


CH 3 -CH=CH-CH 3 






3.7 


-139 


-606.4 


/ra/z.s-2-butene 


CH 3 -CH=CH-CH 3 






0.9 


-106 


-605.4 


c«-2-pentene 


CH 3 -CH 2 -CH=CH- 


-CH 3 




37.9 


-151 


-752.6 


trans-2-pentene 


CH 3 — CH 2 -CH=CH- 


-CH 3 




36.4 


-140 


-751.7 


cw-3-hexene 


CH 3 -CH 2 -CH=CH- 


-CH 2 - 


-CH 3 


66.4 


-138 


-899.7 


rran.s-3-hexene 


CH 3 -CH 2 -CH=CH- 


-CH 2 - 


-CH 3 


67.1 


-113 


-898.1 



chap 4 alkenes 86 

double bond, the dipole moments of cis and trans isomers are different 
(Section 2-6B), allowing an assignment of configuration to be made. Infrared 
spectroscopy (Section 7-4) is also useful for distinguishing cis and trans 
isomers. 

Occasionally, a chemical method, ring closure, can be used to determine the 
configuration of cis-trans isomers. In general, cis isomers can undergo ring 
closure much more readily than the corresponding trans isomers because it 
is not possible to prepare a five- or six-membered ring compound with a 
trans double bond in the ring. The kind of difference which is observed is well 
illustrated by maleic acid, which has a cis double bond and, on heating to 
1 50°, loses water to give maleic anhydride. The corresponding trans isomer, 
fumaric acid, does not give an anhydride at 150°. In fact, fumaric anhydride, 
which would have a trans double bond in a five-membered ring, has never 
been prepared. Clearly, of this pair, maleic acid has the cis configuration and 



o 

II 



c 

II 
o 



OH 
OH 



150° 



- H 2 



maleic 
acid 



O 

II o 
c^ / 

H-" C 

w 
o 

maleic 
anhydride 



O 



Y ^OH 150° 



HO. X. 
^C^ ^H 

II 
O 

fumaric 
acid 



*o 






• rk° 



o" 



fumaric 
anhydride 
(unknown) 



fumaric acid the trans configuration. 



4-4 chemical reactions of alkenes 

We have previously examined briefly two addition reactions of ethene, the 
first member of the homologous series of alkenes. These were addition of 
hydrogen, catalyzed by surfaces of finely divided metals such as nickel, and the 
addition of bromine. These reactions also occur with the higher homologs. 
For example, the colorless, volatile liquid 4-methyl-2-hexene reacts as follows : 



CH 3 
CH 3 -CH 2 -CH-CH=CH-CH 3 

4-methyl-2-hexene Br 




CH 3 
I 
CH 3 - CH 2 - CH- CH 2 - CH 2 - CH 3 

3-methylhexane 

CH 3 Br Br 

II I 

CH 3 -CH 2 -CH-CH-CH-CH 3 

2,3-dibromo-4-methylhexane 



Note that in the names of the three compounds shown, the number 1 carbon 



sec 4.4 chemical reactions of alkenes 87 

atom in one case is at the opposite end of the chain to that for the other two 
compounds. This is necessary to make the names conform to systematic 
usage (Sections 3-1 and 4-1). 

The ease with which addition reactions to alkenes occur is the result of 
the repulsions between the two pairs of electrons that make up the double 
bond. Cleavage of one half of a carbon-carbon double bond requires 63 
kcal, while cleavage of a carbon-carbon single bond requires 83 kcal (Table 
2-1). Furthermore, because the repulsions push the electrons to average 
positions further from the bond axis than the electron positions of a single 
bond, the alkenes will be more readily attacked by electrophiles, that is, 
reagents that act to acquire electrons. On the other hand, nucleophiles 
("nucleus-loving" reagents) are rather poor at reacting with carbon-carbon 
double bonds, unless one or more groups with a high degree of electron- 
withdrawing power are attached to one of the carbon atoms. 

Of the two reagents so far considered, H 2 and Br 2 , the latter, like all the 
halogens, is electrophilic, as we shall see when the mechanism of the reaction 
is considered in the next section. We have already noted that the addition of 
hydrogen to alkenes occurs on activated surfaces, and the availability of the 
electrons in the double bond is here reflected in part in the ease of adsorption 
of the alkene on the metallic surface. 



A. ELECTROPHILIC ADDITION TO ALKENES. THE STEPWISE POLAR 
MECHANISM 

Reagents such as the halogens (Cl 2 , Br 2 , and, to a lesser extent, I 2 ), hydrogen 
halides (HC1, HBr, and HI), hypohalous acids (HOC1 and HOBr), water, 
and sulfuric acid commonly add to the double bonds of alkenes to give 




H OSO,H 



chap 4 alkenes 88 

saturated compounds. These reactions have much in common in their mech- 
anisms and have been much studied from this point of view. They are also of 
considerable synthetic and analytical utility. The addition of water to alkenes 
(hydration) is particularly important for the preparation of a number of com- 
mercially important alcohols. Thus ethyl alcohol and ?-butyl alcohol are made 
on a very large scale by hydrating the corresponding alkenes (ethene and 2- 
methylpropene), using sulfuric or phosphoric acids as catalysts. 



H 2 0, 10% H 2 S0 4 



CH 2 =CH 2 — 


► CH 3 CH 2 OH 

240° 3 2 


ethene 


ethyl alcohol 


H 3 C 

C=CH 2 
H 3 C 


H 2 0, 10% H 2 SO, H 3 C X ^/ CH 3 

25° / \ 

H 3 C OH 


!-methylpropene 


/-butyl alcohol 



We shall pay particular attention here to addition of bromine to alkenes. 
This reaction is conveniently carried out in the laboratory and illustrates a 
number of important points about addition reactions. The characteristics of 
bromine addition are best understood through consideration of the reaction 
mechanism. A particularly significant observation concerning the mechanism 
is that bromine addition (and the other additions listed above) proceeds in the 
dark and in the presence of radical traps (reagents such as oxygen that react 
rapidly with radicals to produce reasonably stable compounds). This is evi- 
dence against a radical chain mechanism analogous to the chain mechanism 
involved in the halogenation of alkanes (Section 2-5B). It does not, however, 
preclude operation of radical addition reactions under other conditions. In 
fact, there are light-induced radical-trap inhibited reactions of bromine and 
hydrogen bromide with alkenes which we shall describe later. 

The alternative to a radical-type chain reaction is an ionic, or polar, reaction 
in which electron-pair bonds are regarded as being broken in a heterolytic 
manner in contrast to the radical, or homolytic, processes discussed pre- 
viously. 

XJ:Y ► X ffl + :Y° heterolytic bond-breaking 

X r H Y * X- + -Y homolytic bond-breaking 

Most polar addition reactions do not seem to be simple four-center, one- 
step processes for two important reasons. First, it should be noted that such 
mechanisms require the formation of the new bonds to be on the same side 
of the double bond and hence produce cis addition (Figure 4-2). However, 
there is ample evidence to show that bromine and many other reagents give 
trans addition. For example, cyclohexene adds bromine and hypochlorous 
acid to give ?ra«.s'-l,2-dibromocyclohexane and /ran.y-2-chlorocyclohexanol. 
Such trans additions can hardly involve simple four-center reactions between 



sec 4.4 chemical reactions of alkenes 89 



plane of the 
ethene molecule 


X— Y 


H */ 





Figure 4-2 Schematic representation of cis addition of a reagent X- 
ethene by a four-center mechanism. (Most reagents do not add ii 
manner.) 



•Y to 
this 



one molecule of alkene and one molecule of an addend X— Y, because the 
X— Y bond would have to be stretched impossibly far to permit the formation 
of trans C— X and C— Y bonds at the same time. 



CH 



Br, + H,C 



2-CH, x 



CH, 



TH==CH' 



H 2 H 2 

c — c 

/ \ 

H,C Br H CH, 

c— c 

I I 

H Br 



HOC1 + H,C 

N 



,CH,-CH 



2\ 



CH=CH 



/ 



CH, 



H 2 H 2 

/ C - C \ 
H 2 C OH H CH 2 

^C— c 

I I 

H CI 



The second piece of evidence against the four-center mechanism is that 
mixtures of products are often formed when addition reactions are carried 
out in the presence of reagents able to react by donation of a pair of electrons 
(nucleophilic reagents). Thus, the addition of bromine to an alkene in methyl 
alcohol solution containing lithium chloride leads not only to the expected 
dibromoalkane, but also to products resulting from attack by chloride ions 
and by the solvent. This intervention of extraneous nucleophilic agents in the 
reaction mixture is evidence against a one-step mechanism. 



Br 2 



-► BrCH,CH,Br 



CH 2 =CH, + Br, 



Br, 



Cl e 
Br, 



-+ ClCH,CH 2 Br + Br e 



CH,-0-H 



♦ CH 3 OCH 2 CH 2 Br + HBr 



chap 4 alkenes 90 

A somewhat oversimplified two-step mechanism that accounts for most of 
the facts is illustrated for the addition of bromine toethene. (The curved arrows 
are not considered to have real mechanistic significance but are used primarily 
to show which atoms can be regarded as nucleophilic — donating electrons — 
and which as electrophilic — accepting electrons. The arrowheads point to the 
atoms that accept electrons.) 



electrophilic 

attack (4-1) 



-> BrCH 2 CH,Br nucleophilic 

. ... " attack (4-2) 

1,2-dibromoetnanc 

(ethylene dibromidc) 



The first step (which involves electrophilic attack on the double bond and 
heterolytic breaking of both a carbon-carbon and a bromine-bromine bond) 
as shown in Equation 4T) produces a bromide ion and carbonium ion. The 
latter is electron deficient (Section 2-5C) and, in the second step of the pos- 
tulated mechanism shown in Equation 4-2, it combines rapidly with an 

available nucleophile (:Br: e ) to give the reaction product. 

Clearly, if other nucleophiles (e.g., : CI : e , CH 3 OH) are present in solution, 
they may compete with the bromide ion for the carbonium ion, as in Equa- 
tions 4-3 and 4-4, and mixtures of products will result. 



:C1. +^ 6 CH 2 — CH 2 Br ► CICH,CH 2 Br (4-3) 

.. ,. — x .. lie 

CH 3 :0: + ffi CH,-CH,Br > CH 3 :0:CH,-CH,Br > CH 3 OCH 2 CH,Br 

H H (4-4) 



In short, we must conclude that the reagents mentioned add across the double 
bond in a trans and stepwise manner and that the two steps take place from 
opposite ends of the double bond. 



B. WHY TRANS ADDITION? 

The simple carbonium-ion intermediate of Equation 4T does not account 
for formation of the ^ra/M-addition product. For one thing, there is no obvious 
reason why free rotation should not occur about the C— C bond of the cation 

I I m 
— C— C® derived from an open-chain alkene; if such occurs, all stereospeci- 

I I 
ficity is lost. In the case of cyclic alkenes, addition of Br° might be expected 

to occur from either side of the ring. 



sec 4.4 chemical reactions of alkenes 91 




H 2 H 2 

C — C 

/ \ 

H 2 C CH 2 

C=C 
/ \ 

H H 



To account for the stereospecificity of bromine addition to alkenes, it has 
been suggested that a cyclic intermediate is formed in which bromine is 
bonded to both carbons of the double bond. This "bridged" ion is called 
a bromonium ion because the bromine formally carries the positive charge. 



H 2 


H 2 


C- 


-c\ 


/ 


\ 


H 2 C Br 


Br CH, 


V 


i y 


C- 


-c 


1 


I CIS 


H 


H 


H 2 


H 2 


c- 


-c s 


/ 


\ 


H,C Br 


H CH 2 


V 


] y 


C- 


-c 


i 


| trans 


H 


Br 




I 



,Br w + :Br: 



.9 



-c 

bromonium ion 

Attack of a bromide ion, or other nucleophile, at the carbon on the side 
opposite the bridging group results in formation of the 7ra/M-addition 
product. 1 

I 
Br— C — 

► I 

— C — Br 




By analogy, a hydrogen-bridged intermediate can be used to account for 
trans addition of acids such as HBr, HC1, H 3 O ffi , and H 2 S0 4 , to alkenes. 
These intermediates are sometimes called protonium ions and might appear 
to violate the usual generalization that hydrogen can form only one stable 



^ -C. 

II + HA ► I: ® ; H + A e 

c — c' 

protonium ion 

1 It is clear that the cis and trans addition routes can be distinguished in the case of 
addition to cycloalkenes on the basis of the stereochemistry of the product. You might 
wonder, however, how this can possibly be done for open-chain alkenes because free 
rotation can occur about the C— C bond of the product. This will be made clear when we 
examine optical isomerism in Chapter 14. 



chap 4 alkenes 92 

bond. It should be emphasized, however, that the bonding between the 
bridging hydrogen and the two carbon atoms is not considered to be normal 
electron-pair covalent bonding. It is different in that one electron pair effects 
the bonding of three atomic centers rather than the usual two. Protonium 
ions of this structure may be regarded as examples of " electron-deficient 
bonding," there being insufficient electrons with which to form all normal 
electron-pair bonds. 

During the past few years, a number of species having electron-deficient 
bonds to hydrogen have been carefully investigated. The simplest example is 
the H 2 ® ion, which may be regarded as a combination of a proton and a 
hydrogen atom. This ion has been detected and studied spectroscopically in 
the gaseous state. Another and very striking example is afforded by the stable 
compound diborane (B 2 H 6 ), which has been shown to have a hydrogen- 
bridged structure. The bonds to each of the bridge hydrogens in diborane, 
like those postulated to the bridge hydrogen of an alkene-protonium ion, are 
examples of three-center electron-pair bonds. Spectroscopic evidence is also 

H x / H x / H 

B; ;B 

H X V X H 

diborane 

available for the stable existence of alkylhalonium ions, (CH 3 ) 2 X®, in solu- 
tions containing extremely weak nucleophiles. 

Whether the intermediates in alkene-addition reactions are correctly 
formulated with bridged bromonium, chloronium, or protonium structures 
is still a controversial matter. Certainly, there are many other reactions of 
carbonium ions that are known to be far from being stereospecific, and 
therefore carbonium ions are not to be considered as necessarily, or generally, 
having bridged structures. It should also be remembered that all ions in 
solution, even those with only transitory existence, are strongly solvated, and 
this in itself may have important stereochemical consequences. In subsequent 
discussion, we shall most frequently write carbonium ions with the charge 
fully localized on one carbon atom, but it should be understood that this may 
not always be either the most accurate or the most desirable representation. 



C. ORIENTATION IN ADDITION TO ALKENES; MARKOWNIKOFF'S 
RULE 

Addition of an unsymmetrical substance such as HX to an unsymmetrical 
alkene can theoretically give two products : 

(CH 3 ) 2 C = CH 2 + HX ► (CH 3 ) 2 C — CH 2 and/or (CH 3 ) 2 C — CH 2 

X H H X 

One of the most important early generalizations in organic chemistry was 
Markownikoff's rule (1870), which may be stated as follows: During the 
addition of HX to an unsymmetrical carbon-carbon double bond, the hydrogen 



sec 4.4 chemical reactions of alkenes 93 

of HX goes to that carbon of the double bond that carries the greater number 
of hydrogens. Thus, Markownikoff' s rule predicts that hydrogen chloride will 
add to propene to give 2-chloropropane (isopropyl chloride) and to 2-methyl- 
propene to give 2-chloro-2-methylp'ropane (?-butyl chloride). These are, in 
fact, the products that are formed. The rule by no means has universal 



CH,-CH = CH, 



HC1 



-> CH3-CH-CH3 

I 
CI 



(CH 3 ) 2 C = CH 2 + HC1 



— (CH 3 ) 2 C-CH 3 

CI 

additions in accord with MarkownikofT's rule 



application, but it is of considerable utility for polar additions to hydrocarbons 
with only one double bond. 



D. A THEORETICAL BASIS FOR MARKOWNIKOFF'S RULE 

To understand the reason for MarkownikofT's rule, it will be desirable to 
discuss further some of the principles that are important to intelligent pre- 
diction of the course of an organic reaction. Consider the addition of hydro- 
gen bromide to 2-methylpropene. Two different carbonium-ion intermediates 
could be formed by attachment of a proton to one or the other of the double- 
bond carbons. Subsequent reaction of the cations so formed with bromide 



\ r^ 

C-CH 2 

/ 



H 3 C 



C-CH, 

/ 



r-butyl cation 



H,C CH 3 

X 

H,C Br 



/-butyl bromide 



H,C 



C-CH 2 + H s 

/ 



H 3 C v H 
H,C CH, 9 



isobutyl cation 



Br e 



H 3 C x /H 
H,C CH,Br 



isobutyl bromide 



ion gives ?-butyl bromide and isobutyl bromide. In the usual way of running 
these additions, the product is, in fact, quite pure 7-butyl bromide. 

How could we have predicted which product would be favored ? The first 
step is to decide whether the prediction is to be based on which of the two 
products is the more stable, or which of the two products is formed more 
rapidly. If we make a decision on the basis of product stabilities, we take into 
account AH and AS values to estimate an equilibrium constant K between the 
reactants and each product. When the ratio of the products is determined by 
the ratio of their equilibrium constants, we say the overall reaction is subject 
to equilibrium (or thermodynamic) control. This will be the case when the 
reaction is carried out under conditions that make it readily reversible. 

When a reaction is carried out under conditions in which it is not reversible, 



chap 4 alkenes 94 

the ratio of the products is determined by the relative rates of formation of the 
products. Such reactions are said to be under kinetic control. To predict rela- 
tive reaction rates, we take into account steric hindrance, stabilities of possible 
intermediates, and so on. 

Addition of hydrogen bromide to 2-methylpropene is predicted by Mark- 
ownikoff's rule to give /-butyl bromide. It turns out that the equilibrium 
constant connecting /-butyl bromide and isobutyl bromide is 4.5 at 25°, 
meaning that 82% of an equilibrium mixture is /-butyl bromide and 18% is 
isobutyl bromide. 

[/-butyl bromide] 
[isobutyl bromide] 

Addition of hydrogen bromide to 2-methylpropene actually gives 99 + % 
/-butyl bromide in accord with Markownikoff's rule. This means that the 
rule is a kinetic-control rule and may very well be invalid under conditions 
where addition is reversible. 

If Markownikoff's rule depends on kinetic control of the product ratio in 
the polar addition of hydrogen bromide to 2-methylpropene, then it is proper 
to try to explain the direction of addition in terms of the ease of formation 
of the two possible carbonium-ion intermediates. There is abundant evidence 
that tertiary carbonium ions are more easily formed than secondary car- 
bonium ions and these, in turn, are more easily formed than primary carbo- 
nium ions. A number of carbonium salts have been prepared, and the tertiary 
ones are by far the most stable. Thus, the theoretical problem presented by 
Markownikoff's rule is reduced to predicting which of the two possible 
carbonium-ion intermediates will be most readily formed. With the simple 
alkenes, formation of the carbonium ion accords with the order of preference 
tertiary > secondary > primary. 



E. ADDITIONS OF UNSYMMETRICAL REAGENTS OPPOSITE 
TO MARKOWNIKOFF'S RULE 

The early chemical literature concerning the addition of hydrogen bromide to 
unsymmetrical alkenes is rather confused, and sometimes the same alkene was 
reported to give addition both according to and in opposition to Markow- 
nikoff's rule under very similar conditions. Much of the uncertainty about the 
addition of hydrogen bromide was removed by the classical researches of 
Kharasch and Mayo (1933), who showed that there must be two reaction 
mechanisms, each giving a different product. Under polar conditions, 
Kharasch and Mayo found that hydrogen bromide adds to propene in a rather 
slow reaction to give pure 2-bromopropane (isopropyl bromide): 



CH 3 CH=CH 2 + HBr ■ ► CH 3 CHCH 3 

polar J I * 

conditions 



slow 
polar 

Br 



With light or peroxides (radical initiators) and in the absence of radical 



sec 4.4 chemical reactions of alkenes 95 



traps, a rapid radical chain addition of hydrogen bromide occurs to yield 
80% or more of 1-bromopropane (n-propyl bromide): 



HBr ^!— ► CH 3 CH,CH 2 Br 

peroxides 



Similar effects have been occasionally noted with hydrogen chloride but 
never with hydrogen iodide or fluoride. A few substances apparently add to 
alkenes only by radical mechanisms and always give addition opposite to 
Markownikoff 's rule. 

The polar addition of hydrogen bromide was discussed in the previous 
section and will not be further considered now. Two questions with regard to 
the so-called abnormal addition will be given special attention: why the 
radical mechanism should give a product of different structure from the polar 
addition, and why the radical addition occurs readily with hydrogen bromide 
but rarely with the other hydrogen halides (see Exercise 4-17). 

The abnormal addition of hydrogen bromide is strongly catalyzed by 
peroxides, which have the structure R— O— O— R and decompose thermally 
to give radicals : 

R — 6:5 — R ► 2 R— O- Atf=+35kcal 

The RO • radicals can react with hydrogen bromide in two ways : 

^ ROH + Br- A//=~23kcal 



RO- + HBr ^^ 

^"^ ROBr + H- AH= +39kcal 



Clearly, the formation of ROH and a bromine atom is energetically more 
favorable. The overall process of decomposition of peroxide and attack on 
hydrogen bromide, which results in the formation of a bromine atom, can 
initiate a radical chain addition of hydrogen bromide to an alkene : 



Chain 


CH 3 CH=CH 2 + Br- > 


CH 3 CH-CH 2 Br 


AH = 


- 5 kcal 


propa- 


t^Ttr j"irr /in n i tin 


fc PU pit ptr 13 1 t> . 


Atf = 


- 1 1 kcal 


gation: 


LH3CH Crl 2 or + HBr 


* Cri3^rl2^rl2Dr ^ or 


Chain 


rt 1 I r> ' » P ' T> ' 


R' • = atom or radical 






termi- 
nation: 


K ' t K ' K K. 





The chain-propagating steps, taken together, are exothermic by 16 kcal 
and have a fairly reasonable energy balance between the separate steps, 
which means that one is not highly exothermic and the other highly endother- 
mic. Both steps are, in fact, comparably exothermic. The reaction chains 
appear to be rather long, since only traces of peroxide catalyst are needed and 
the addition is strongly inhibited by radical traps. 

The direction of addition of hydrogen bromide to propene clearly depends 



chap 4 alkenes 96 

on which end of the double bond the bromine attacks. The choice will depend 
on which of the two possible carbon radicals that may be formed is the 

CH 3 -CH— CH 2 -Br CH 3 — CH-CH 2 - 

Br 

[1] [2] 

more stable, the l-brom'o-2-propyl radical [1] or the 2-bromo-l -propyl 
radical [2]. As with carbonium ions, the ease of formation and stabilities of 
carbon radicals follow the sequence tertiary > secondary > primary. There- 
fore, the secondary l-bromo-2-propyl radical [1] is expected to be more 
stable and more easily formed than the primary 2-bromo-l -propyl radical [2]. 
The product of radical addition should be, and indeed is, 1-bromopropane. 

It may seem strange to refer to certain radicals or ions as being stable and 
therefore more likely to be reaction intermediates than other less stable 
radicals or ions. Would not the unstable radicals or ions actually be more 
likely as intermediates because they would react more rapidly to give products ? 
"Stable" is used here only in a relative sense. All of the radicals and ions 
which we are invoking as intermediates react very quickly to give products 
and never attain high concentrations in the reaction mixture. This means that 
the reactions will go more readily by way of the relatively more stable in- 
termediates because these are formed most easily and react rapidly to give 
products. The less stable intermediates are not formed as readily and the fact 
that they would react more rapidly does not increase the overall reaction rate 
in processes which would involve them. This point will be considered in other 
connections later. The important thing to recognize is that there may be 
large differences in the ease of formation of different kinds of reaction in- 
termediates — so much so that mechanisms which imply that primary car- 
bonium ions or radicals (RCH 2 ffi or RCH 2 *) are formed in preference to 
secondary or tertiary carbonium ions and radicals should be regarded as 
suspect. 

F. ADDITION OF BORON HYDRIDES TO ALKENES 

A recently developed and widely used reaction is that of diborane (B 2 H 6 ) 
with alkenes. Diborane (Section 19-5) is the dimer of the electron-deficient 
species BH 3 , and it is as BH 3 that it adds to the double bond to give tri- 
alkylboron compounds (organoboranes). With ethene, triethylborane 
results : 

6 CH 2 =CH 2 + B 2 H 6 — ^-» 2(CH 3 CH 2 ) 3 B 

This reaction is called hydroboration ; it proceeds in three stages, but the 
intermediate mono- and dialkylboranes are not generally isolated, as they 
react rapidly by adding further to the alkene. 



CH 2 =CH 2 + CH 3 CH 2 BH 2 ► (CH 3 CH 2 ) 2 BH 

CH 2 =CH 2 + (CH 3 CH 2 ) 2 BH ► (CH 3 CH 2 ) 3 B 



sec 4.4 chemical reactions of alkenes 97 

With an unsymmetrical alkene such as propene, hydroboration occurs so 
that boron becomes attached to the less substituted end of the double bond — 
with propene forming tri-«-propylborane. 

6 CH 3 CH=CH 2 + B 2 H 6 2 (CH 3 CH 2 CH 2 ) 3 B 

Hydroborations have to be carried out with some care, since diborane and 
alkylboranes are highly reactive substances ; in fact, they are spontaneously 
inflammable in air. For most synthetic purposes it is not necessary to isolate 
the addition products, and diborane can be generated either in situ or exter- 
nally through the reaction of boron trifluoride with sodium borohydride. 



3 NaBH 4 + 4 BF 3 ► 2 B 2 H 6 + 3 NaBF 4 

Boron trifluoride is conveniently used in the form of its stable complex 
with diethyl ether, (C 2 H 5 ) 2 0:BF 3 , the reactions usually being carried out in 
ether solvents such as diethyl ether, (C 2 H 5 ) 2 0; diglyme, (CH 3 OCH 2 CH 2 ) 2 0; 
or tetrahydrofuran, (CH 2 ) 4 0. 

The most common synthetic reactions of the resulting alkylboranes are 
oxidation with alkaline hydrogen peroxide to the corresponding primary 
alcohol, and cleavage with aqueous acid (or, better, anhydrous propanoic 
acid, CH 3 CH 2 C0 2 H) to give alkanes. Thus, for tri-n-propylborane : 

(CH 3 CH 2 CH 2 ) 3 B + 3H 2 2 -^^ 3 CH 3 CH 2 CH 2 OH + B(OH) 3 

n -propyl alchohol 
(a primary alcohol) 

(CH 3 CH 2 CH 2 ) 3 B + 3H 2 "^ > 3 CH 3 CH 2 CH 3 + B(OH) 3 

The first of these processes achieves " anti-Markownikoff " addition of water 
to a carbon-carbon double bond as the overall result of the two steps. The 
second reaction provides a method of reducing carbon-carbon double bonds 
without using hydrogen and a metal catalyst. Both of these conversions are 
difficult to do any other way and this accounts for the extensive use that 
organic chemists have made of diborane in recent years. 



G. OXIDATION OF ALKENES 

Most alkenes react readily with ozone, even at low temperatures, to cleave 
the double bond and yield cyclic peroxide derivatives known as ozonides. 

O 
3 -> H.CHC^ X CHCH, 



-80° \ I 

o-o 

2-butene ozonide 
Considerable evidence exists to indicate that the overall reaction occurs in 



chap 4 alkenes 98 

three main steps, the first of which involves a cz's-cycloaddition reaction that 
produces an unstable addition product called a molozonide. 



H 3 C CH 3 

3 \ / 3 
CH,-HC=CH-CH, HC-CH 

/ \ 



<x o e o. o 

2 ° 

molozonide (unstable) 



HjC \ // CHj H 3 C rH, 



HQfCH CH HC /O 

/^'C.\ > II + IL " CH 3 HCT X CHCH 3 

°< /° o o! _ \ / 

fro' ^o e o-o 



€ 



ozonide 



Ozonides, like most substances with peroxide (O— O) bonds, may explode 
violently and unpredictably. Ozonizations must therefore be carried out with 
due caution. The ozonides are not usually isolated but are destroyed by 
hydrolysis with water and reduction with zinc to yield carbonyl compounds 
that are generally quite easy to isolate and identify. (In the absence of zinc, 
hydrogen peroxide is formed which may degrade the carbonyl products by 
oxidation). The overall reaction sequence provides an excellent means for 



° O 

/O x h 2 // \ 

H 3 CHC CHCH 3 ► CH 3 C + CCH 3 + ZnO 

\ / Zn \ / 

O-O H H 



locating the positions of double bonds in alkenes. The potentialities of the 
method may be illustrated by the difference in reaction products between the 
1- and 2-butenes: 

O 
i.o 3 / 

CH 3 CH=CHCH 3 ► 2CH 3 C 

3 2. Zn,H 2 3 \ 

H 

O H 

1.O3 // \ 

CH 3 CH,CH=CH, ► CH 3 CH 2 C + C = 

32 2 2. Zn,H 2 \ / 

H H 

Natural rubber (polyisoprene, Section 28-2) is a substance with many 
double bonds, and ozone formed in the atmosphere by sunlight or by smog- 
producing reactions (Section 3-3A) combines with the double bonds of the 
rubber and causes the rubber to crack. This destructive action can be re- 
duced by antioxidants mixed with the rubber, or by use of rubberlike materials 
without double bonds (see Section 4-4H). 

Several other oxidizing reagents react with alkenes under mild conditions 
to give, overall, addition of hydrogen peroxide as HO— OH. Of particular 



sec 4.4 chemical reactions of alkenes 99 

importance are permanganate ion and osmium tetroxide, both of which 
react in an initial step by a cw-cycloaddition mechanism like that postulated 
for ozone: 



°s P 

Mn 



o o 



r \ 


/ 1 


'f- 


-s- 


o 





\ 


/ 


Mn 


// 


V 


L o 


J 



unstable 



O O 

stable osmate 
ester 



H 2 o 
OH e 



mild 
reduction 

(Na 2 S0 3 ) ' 



— C- 

I 
HO 



I 
-C — 

I 
OH 



— C- 

I 
HO 



I 
-C- 



MnO^ 



L 



MnO, + MnO< 



+ Os 



OH 



Each of these reagents produces m-dihydroxy compounds (diols) with 
cycloalkenes : 




l,OsO*,25° 

2.Na 2 S0 3 



(or KMnO, , OH e , H 2 0) 




OH 



OH 

cis- 1 ,2-cyclopentanediol 



An alternate scheme for oxidation of alkenes with hydrogen peroxide in 
formic acid follows a different course in that trans addition occurs. (The 
mechanism of this reaction is analogous to the addition of bromine to a 
carbon-carbon double bond, which also takes place by trans addition.) 



<CJ 



35%H 2 Q 2 
HC0 2 H, 25° 



OH 



OH 

trans-\ ,2-cyclopentanediol 



H. POLYMERIZATION OF ALKENES 

One of the most important industrial reactions of alkenes is their conversion 
to higher-molecular-weight compounds (polymers). A polymer is here de- 
fined as a long-chain molecule with recurring structural units. Polymerization 
of propene, for example, gives a long-chain hydrocarbon with recurring 
CH 3 units. Most industrially important polymerizations of alkenes 

-CH-CH 2 - 



chap 4 alkenes 100 

CH 3 

CH 3 CH=CH 2 ► -j-CH— CH 2 - AH= -n ■ 20kcal 

propene polypropene 

(polypropylene) 

occur by chain mechanisms and may be classed as anion, cation, or radical- 
type reactions, depending upon the character of the chain-carrying species. 
In each case, the key steps involve successive additions to molecules of the 
alkene. The differences are in the number of electrons that are supplied by 
the attacking agent for formation of the new carbon-carbon bond. For 
simplicity, these steps will be illustrated by using ethene, even though it does 
not polymerize very easily by any of them : 



R— CH 2 -CH? + ch 2 =Q:h 2 




R-CH 2 — CH 2 + CH 2 =CH 2 > R-CH 2 -CH 2 -CH 2 -CH 2 , etc. 

R-CH 2 -CH 2 + CH 2 -CH 2 > R-CH 2 -CH 2 -CH 2 -CH 2 

Anionic Polymerization. Initiation of alkene polymerization by the anion- 
chain mechanism may be formulated as involving an attack by a nucleo- 
philic reagent Y: e on one end of the double bond and formation of a carban- 
ion. Attack by the carbanion on another alkene molecule gives a four- 

Y: + CH,^CH 



2 — v~n 2 ^ i«^n 2 

carbanion 



carbon carbanion, and subsequent additions to further alkene molecules 
lead to a high-molecular-weight anion. The growing chain can be terminated 



► Y:CH 2 -CH 2 -CH 2 -CH 2 

n (CH 2 =CH 2 ) n 
> Y:CH 2 -CH 2 -fCH 2 -CH 2 ^rCH 2 -CH 2 e 

by any reaction (such as the addition of a proton) that would destroy the 
carbanion on the end of the chain : 



Y:CH 2 -CH 2 -eCH 2 -CH 2 ^-CH 2 -CH 2 e ► YiCHj-CHj-fCHj-CH^CHj-CHj 

Anionic polymerization of alkenes is quite difficult to achieve, since few 
anions (or nucleophiles) are able to add readily to alkene double bonds 
(see p. 87). Anionic polymerization occurs readily only with ethenes 
substituted with sufficiently powerful electron-attracting groups to expedite 
nucleophilic attack. 



sec 4.4 chemical reactions of alkenes 101 

Cationic Polymerization. Polymerization of an alkene by acidic reagents 
can be formulated by a mechanism similar to the addition of hydrogen 
halides to alkene linkages. First, a proton from a suitable acid adds to an 
alkene to yield a carbonium ion. Then, in the absence of any other reasonably 
strong nucleophilic reagent, another alkene molecule donates an electron 
pair and forms a longer chain cation. Continuation of this process can lead 
to a high-molecular-weight cation. Termination can occur by loss of a proton. 



CH, = CH, 



CH 3 -CH 2 -CH 2 — CH 

-H® 



CH3 CH2 ~t~ CH 2 — CH 2 

n(CH 2 =CH 2 ) 



* CH 3 -CH 2 -fCH 2 -CH 2 ^CH 2 -CH 2 tt 



CH 3 -CH 2 -eCH 2 -CH 2 -^-CH=CH 2 



Ethene does not polymerize by the cationic mechanism, because it does 
not have groups that are sufficiently electron donating to permit ready 
formation of the intermediate growing-chain cation. 2-Methylpropene has 
electron-donating alkyl groups and polymerizes much more easily than ethene 
by this type of mechanism. 

The usual catalysts for cationic polymerization of 2-methylpropene are 
sulfuric acid, hydrogen fluoride, or boron trifluoride plus small amounts of 
water. Under nearly anhydrous conditions, a very long-chain polymer is 
formed called " polyisobutylene." Polyisobutylene fractions of particular 



CH, 



CH, 



CH, 



tr. h 2 o 



CH, 



CH, 



-C- 

I 



CH, 



CH 2 



CH, 

I 
C — 

I 



CH, 



CH 3 

polyisobutylene 



CH, 



CH, 



molecular weights are very tacky and are used as adhesives for pressure- 
sealing tapes. 

In the presence of 60 % sulfuric acid, 2-methylpropene is not converted to 
a long-chain polymer, but is dimerized to a mixture of C 8 alkenes. The mech- 
anism is like the polymerization reaction described for polyisobutylene, 
except that chain termination occurs after only one alkene molecule has been 
added. The short chain length is due to the high water concentration; the 
intermediate carbonium ion loses a proton to water before it can react with 



CH, 



CH, 



60%H 2 SO, 

> 

70° 



CH, 



CH, 

I 
-c-c 

I 

CH, 



/H 2 

C 
\ 
CH, 



+ CH 3 - 



CH 3 
I 
-C- 

I 
CH, 



CH 3 

/ 3 



CH = C 

\ 



CH, 



20% 



"diisobutylene" 



chap 4 alkenes 102 

another alkene molecule. Because the proton can be lost two different ways, 
a mixture of alkene isomers is obtained. The alkene mixture is known as 
" diisobutylene " and has a number of commercial uses. Hydrogenation gives 
2,2,4-trimethylpentane (often erroneously called " isooctane "), which is used 
as the standard "100 antiknock rating" fuel for internal-combustion gasoline 
engines (Section 3-3A). 



CH, 



CH, = C 



CH, 



CH, 



CH 2 =C 



CH 3 

/ 



- CH,-C S 



\ 
CH, 



CH, 



\ CH 3 

CH 3 I / 

► CH 3 -C-CH 2 -C ffi 

I \ 

CH 3 CH 3 



CH, 



CH, 



CH, 



CH, 



CH 3 
I 
+ CH,-C-CH=C 



CH, 



CH, 



CH, 



CH 3 CH 3 

H 2 (Ni) I I 

diisobutylene isomers ► CH 3 — C— CH, — CH — CH, 

50° J j 2 3 

CH 3 

2,2,4-trimethylpentane 

Radical Polymerization. Ethene may be polymerized with peroxide cata- 
lysts under high pressure (1000 atmospheres or more, literally in a cannon 
barrel) at temperatures in excess of 100°. The initiation step involves forma- 
tion of RO radicals, and chain propagation entails stepwise addition of radi- 
cals to ethene molecules. 



Initiation: R:Q:0:R 



-> 2R=0- 



Propa- 
gation: 



R:0- + CH 2 = CH 2 - 
R:p:CH 2 -CH 2 + «(CH 2 =CH 2 ) 



♦ R:0:CH,-CH, 



-» RO-r-CH 2 -CH 2 ^-CH 2 -CH 2 



Termi- 
nation: 



2ROfCH 2 — CH 2 ^;CH 2 — CH 2 



- [RO-eCH.-CH^CH.-CH^ 

combination 



RO-(-CH 2 -CH 2 i-„CH=CH 2 + ROf CH 2 CH 2 )„- 
disproportionation 



-CH,-CH, 



Chain termination may occur by any reaction resulting in combination or 
disproportionation of two radicals. (Disproportionation means that two 
identical molecules react with one another to give two different product 
molecules.) The polymer produced this way has from 100 to 1000 ethene 
units in the hydrocarbon chain. The polymer, called Polythene (or sometimes 
polyethylene), possesses a number of desirable properties as a plastic and is 



summary 103 

widely used for electrical insulation, packaging films, piping, and a variety 
of molded particles. The very low cost of ethene (a few cents a pound) makes 
Polythene a commercially competitive material despite the practical diffi- 
culties involved in the polymerization process. Propene and 2-methylpropene 
do not polymerize satisfactorily by radical mechanisms. 

Coordination Polymerization. A relatively low-pressure low-temperature 
ethene polymerization has been achieved with an aluminum-molybdenum 
oxide catalyst, which requires occasional activation with hydrogen (Phillips 
Petroleum). Ethene also polymerizes quite rapidly at atmospheric pressure 
and room temperature in an alkane solvent containing a suspension of the 
insoluble reaction product from triethylaluminum and titanium tetrachloride 
(Ziegler). Both the Phillips and Ziegler processes produce a very high- 
molecular-weight polymer with exceptional physical properties. The unusual 
characteristics of these reactions indicate that no simple anion, cation, or 
radical mechanism can be involved. It is believed that the catalysts act by 
coordinating with the alkene molecules in somewhat the way hydrogenation 
catalysts combine with alkenes. 

Polymerization of propene by catalysts of the Ziegler type gives a most 
useful plastic material. It can be made into durable fibers or molded into a 
variety of shapes. Copolymers (polymers with more than one kind of monomer 
unit in the polymer chains) of ethene and propene made with Ziegler cata- 
lysts have highly desirable rubberlike properties and are potentially the cheap- 
est useful elastomers (elastic polymers). A Nobel Prize was shared in 1963 by 
K. Ziegler and G. Natta for their work on alkene polymerization. 



summary 

Alkenes are hydrocarbons possessing a carbon-carbon double bond. Simple 
open-chain alkenes have the formula C„H 2 „ . The IUPAC names for alkenes 
are obtained by finding the longest continuous carbon chain containing the 
double bond and giving it the name of the corresponding alkane with the 
ending changed from -ane to -ene. The numbering of the carbon chain is 
started at the end that will provide the lowest number for the position of the 
first carbon of the double bond. Thus, 



I 
CH 3 - CH- CH 2 - CH=CH- CH 3 

is 5-methyl-2-hexene. Some common alkenyl groups (an alkene minus a 
hydrogen atom) are vinyl (CH 2 =CH-), allyl (CH 2 =CH-CH 2 -), and 
isopropenyl (CH 2 =C— CH 3 ). Compounds with two carbon-carbon double 

bonds are named as alkadienes; if the double bonds are adjacent they are 
called cumulated ; if they are separated by a single bond they are conjugated ; 
and if they are separated by more than one single bond they are isolated. 



chap 4 alkenes 104 

Both structural and geometrical isomerism appear at the C 4 level in the 
alkene series, there being three structural isomers. One of these (2-butene) 
exists in cis and trans forms. Trans isomers have the substituents on the 

H 3 C CH 3 H 3 C^ H 

c=c c-c 

/ \ / \ 

H H H CH 3 

cu-2-butene trans-2-buiene 

opposite side of the double bond and usually are of lower energy (more stable) 
than their cis isomers. Trans isomers usually have the higher melting points, 
lower boiling points, lower dipole moments (often zero) and, because their 
substituent groups are far apart, they do not undergo ring closure reactions 
which may occur with some cis compounds. 

The physical properties of alkenes are similar to those of the corresponding 
alkanes, but alkenes are much more reactive chemically. Because of the 
concentration of electrons in the double bond, alkenes are subject to attack 
by electrophiles (reagents that seek electrons). These addition reactions can 
be illustrated with a typical alkene such as propene : 



CH 3 CH=CH 2 " 2 > CH,CH 2 CH 

3 i Nl J z 



3^112^113 

Br 2 



(other halogens 
behave similarly) 

HOC1 

H 2 so 4 OS0 3 H 

► I 

CH 3 CHCH 3 

HC1 

-♦ CH3CHCICH3 



■> CH 3 CHBrCH,Br 



(HI behaves 
similarly) 



HRr 

-> CHXHBrCH, 



(peroxide-free, in dark) 
HBr 



(peroxides) 
H 2 



H 8> 



-♦ CH 3 CHOHCH, 



H2O2 
B a H 6 _ /OH e 



-* (CH 3 CH 2 CH 2 ) 3 B 



\H 2 

CH 3 CH 2 CH 3 



/°v. 



-^— CH 3<\ H CH 2 Jh^ C H 3 CHO + CH 2 

O-O Zn 



Mn0 4 e 9 

> I (cis addition) 

CH 3 CHCH 2 OH 



exercises 10S 



HC0 2 H OH 

., _ ' I (trans addition) 

H *°> CH 3 CHCH 2 OH 



R e 



RO- 



^y 



CH 3 


CH, CH, 


(polymer; none of 


1 


1 1 


these reactions works 


CH-CH 2 - 


-CH-CH 2 -CH- 


very well for ethene 
or propene) 



The reactions shown involve attack by electrophilic reagents at the double 
bond with the following four exceptions: the metal-induced reaction with 
H 2 ; hydrogen bromide with peroxides; and polymerization initiated by 
radicals, R-, or anions, Y e , both of which are difficult to achieve. 

Addition of most electrophiles occurs stepwise by way of ionic intermediates 
(heterolytic bond breaking), with the groups being connected to the carbons 
in the trans manner. 

When unsymmetrical electrophilic reagents add to unsymmetrical alkenes, 
Markownikoff's rule can be used to predict the principal product. Thus, 
during the addition of HX, the hydrogen goes to that carbon of the double 
bond that carries the greater number of hydrogens — for example, CH 3 CH= 
CH 2 + HX -> CH3CHXCH3 . The basis for this rule is the tendency for that 
part of the electrophile that initiates the reaction (H® from HX) to add in 
such a way as to produce the lowest-energy carbonium ion. (Tertiary carbon- 
ium ions are of lowest energy and primary carbonium ions are of highest ener- 
gy.) For this reason, the first step of the HX addition is CH 3 CH=CH 2 + 

9 
H® -» CH 3 CHCH 3 , to give the secondary carbonium ion (not CH 3 CH— 

CH 2 + H® -> CH 3 CH 2 CH 2 ffi to give the primary carbonium ion); the final 

© 
step involves addition of X e , CH 3 CHCH 3 + X e -* CH 3 CHXCH 3 . 

The ratios of products in such reactions show that they are governed by 
the rates of the two possible reaction paths, not by the stabilities of the final 
products. This is called kinetic control of the reaction as opposed to equilib- 
rium (or thermodynamic) control. 

exercises 

4-1 Name each of the following substances by the IUPAC system and, if straight- 
forward to do so, as in examples a and e, as a derivative of ethylene : 



a. (CH 3 ) 2 C=CHCH 3 / ^"^^h 



CH 



b. C1 2 C=C(CH 3 ) 2 

c. (CH 3 ) 3 CCH 2 C(CH 3 )=CH 2 , 

d. (CH 3 ) 2 C=C=CHBr H 2 C /CH ^CH 

e. [(CH 3 ) 2 CH] 2 C=C[CH(CH 3 ) 2 ] 2 g- j| H 

XHT CH 3 



chap 4 alkenes 106 

4-2 Write structural formulas for each of the following substances: 

a. trifiuorochloroethylene d. l,l-di-(l-cyclohexenyl)-ethene 

b. 1,1-dineopentylethylene e. trivinylallene 

c. 1 ,4-hexadiene 

4-3 The trans alkenes are generally more stable than the cis alkenes. Give one or 
more examples of unsaturated systems where you would expect the cis form to 
be more stable and explain the reason for your choice. 

4-4 Write structural formulas for each of the following: 

a. The thirteen hexene structural isomers; name each by the IUPAC 
system. Show by suitable formulas which isomers can exist in cis 
and trans forms and correctly designate each. 

b. All trans- 1 , 1 8-di-(2,6,6-trimethyl-l -cyclohexenyl)-3,7, 11,1 5-tetramethyl 
1,3,5,7,9,11, 13,15,17-octadecanonaene(C 4 oH 56 ). 

4-5 Calculate, from the data in Table 3-7 and any necessary bond energies, the 
minimum thermal energy that would be required to break one of the ring 
carbon-carbon bonds and interconvert cis- and frays'- 1,2-dimethylcyclo- 
butanes (see pp. 67-69). 

4-6 What volume of hydrogen gas (STP) is required to hydrogenate 100 g of a 
mixture of 1 -hexene and 2-hexene? 

4-7 Supply the structure and a suitable name for the products of the reaction of 
2-methyl-2-pentene with each of the following reagents : 

a. H 2 , Ni d. HBr (plus peroxide) 

b. Cl 2 e. B 2 H 6 followed by aqueous acid 

c. Cl 2 in presence of NH 4 F 

4-8 How could bromoethane be prepared starting with ethyne? 

4-9 Show how each of the following compounds could be prepared starting with 
1,5-hexadiene. 

a. 1,2,5,6-tetrabromohexane 

b. 2,5-diiodohexane 

c. 2-iodohexane 

4-10 Write the structures of the products of the reaction of 3,4-dimefhyl-2-octene 
with each of the following reagents. 

a. diborane followed by hydrogen peroxide and base 

b. dilute aqueous sulfuric acid 

c. hypobromous acid 

d. aqueous potassium permanganate 

e. ozone followed by zinc and steam 

4-11 Calculate AH (vapor) for addition of fluorine, chlorine, bromine, and iodine 
to an alkene. What can you conclude from these figures about the kind of 
problems that might attend practical use of each of the halogens as a reagent 
to synthesize a 1,2-dihalide? 



exercises 107 

4-12 a. Write as detailed a mechanism as you can for the trans addition of hypo- 
chlorous acid (HOC1) to cyclopentene. 
b. How does the fact that HOC1 is a weak acid (K HA in water = 7 x 10~'°) 
make formation of CH 3 CH 2 OCl from ethene unlikely? 

4-13 Calculate AH for the addition of water to ethene in the vapor state at 25°. 
Why are alkenes not hydrated in aqueous sodium hydroxide solutions ? 

4-14 When r-butyl bromide is allowed to stand at room temperature for long 
periods, the material becomes contaminated with isobutyl bromide. Write a 
reasonable mechanism for the formation of isobutyl bromide under the 
influence of traces of water and/or oxygen from the atmosphere. 

4-15 Arrange ethene, propene, and 2-methylpropene in order of expected ease of 
hydration with aqueous acid. Show your reasoning. 

4-16 Write two different radical chain mechanisms for addition of hydrogen 
chloride to alkenes and consider the energetic feasibility for each. 

4-17 Calculate the A// values for initiation and chain propagation steps of radical 
addition of hydrogen fluoride, hydrogen chloride, and hydrogen iodide to an 
alkene. Would you expect these reagents to add easily to double bonds by 
such a mechanism ? 

4-18 Bromotrichloromethane, CBrCl 3 , adds to 1-octene by a radical chain 
mechanism on heating in the presence of a peroxide catalyst. Use bond 
energies (Table 2-1) to devise a feasible mechanism for this reaction and work 
out the most likely structure for the product. Show your reasoning. 

4-19 Determine from the general characteristics of additions to double bonds 
whether the direction of addition of B 2 H 6 to propene is consistent with a 
polar mechanism. 

4-20 The following physical properties and analytical data pertain to two isomeric 
hydrocarbons, A and B, isolated from a gasoline: 




Both A and B readily decolorize bromine and permanganate solutions and 
give the same products on ozonization. Suggest possible structures and 
configurations for A and B. What experiments would you consider necessary 
to further establish the structure and configuration of A and B? 

4-21 a. Write a mechanism for the sulfuric acid-induced dimerization of trimethyl- 
ethylene, indicating the products you expect to be formed. 
b. Ozonization of the mixture that is actually formed gives, among 



chap 4 alkenes 108 

o 

II 

other carbonyl products(— C— ), substantial amounts of 2-butanone 

( ' ) 

\CH 3 — C— CH 2 — CH 3 / Write a structure and reaction mechanism for 
formation of a Cio alkene that might reasonably be formed in the dimer- 
ization reaction and that, on ozonization, would yield 2-butanone and a 
C 6 carbonyl compound. (Consider how sulfuric acid might cause the 
double bond in trimethylethylene to shift its position.) 

4-22 A pure hydrocarbon of formula C 6 Hi 2 does not decolorize bromine water. 
Draw structures for at least six possible compounds that fit this description 
(including geometrical isomers, if any). 

4-23 Calculate the heats ( — A/7) of the following reactions in the gas phase at 25° : 

H 2 2 + (CH 2 ) 2 > HO-CH 2 -CH 2 -OH 

H 2 2 + (CH 2 ) 6 ► HO-(CH 2 ) 6 -OH 

a. What conclusion as to the rates of the above reactions can be made on 
the basis of the A// values? Explain. 

b. What change in the heats of the reactions would be expected if they were 
carried out in the liquid phase ? Why ? 

c. What agents might be effective in inducing the reactions in the liquid 
phase? Explain. 

4-24 Evaluate (show your reasoning) the possibility that the following reaction 
will give the indicated product: 




D 2 




If you do not think the indicated product would be important, write the 
structure(s) of the product(s) you think most likely to be found. 

4-25 Investigate the energetic feasibility of adding ammonia (NH 3 ) to an alkene 
by a radical chain mechanism with the aid of a peroxide (ROOR) catalyst. 
Would such a mechanism give addition in accord with Markownikoff's rule? 
Why? What practical difficulties might be encountered in attempts to add 
ammonia to 2-methylpropene with a sulfuric acid catalyst ? 

4-26 It has been found possible to synthesize two isomeric cycloalkenes of formula 
C 8 H 14 . Both of these compounds react with hydrogen in the presence of 
platinum to give cyclooctane, and each, on ozonization followed by reduction, 
gives 

ii ii 

H-C-CH 2 -CH 2 -CH 2 -CH 2 -CH 2 -CH 2 -C-H 

a. What are the structures and configurations of the two compounds ? 

b. Would the two substances give the same compound on hydroxylation 
with potassium permanganate? 



itlti 
lap, 

gin; 



MsgfS Saw 



Kg 



'MJs&s* *kSs* ,5-Jfeii 















lis'ssEsr VSSSSf RJ» 



S3 










• v- '::■/■*•■> 

■. •'.■'■'' 

"t">* \$'A' >V^G 






liii 







^^tf^iv «^ 






•v.i^i*" 1 



i /y*\4y?*:- : 






: '■Hi..* 1 3i 



k ■;."-■-: r.;\,i:';V .'**%) 



chap 5 alkynes 111 

Alkynes are hydrocarbons with carbon-carbon triple bonds. The simplest 
alkyne is ethyne, H— C=C — H, usually called acetylene, an important 
starting material for organic syntheses, especially on an industrial scale. We 
have previously discussed the geometry of ethyne and its addition reactions 
with hydrogen and bromine (Section 2-6). 



5-1 nomenclature 

The IUPAC system for naming alkynes employs the ending -yne in place of the 
-ane used for the name of the corresponding, completely saturated, hydro- 
carbon. Many alkynes are conveniently named as substitution products of 
acetylene, as shown in parentheses in these examples. 

H-C = C-H CH 3 -C = C-CH 3 

ethyne 2-butyne 

(acetylene) (dimethylacetylene) 

The numbering system for location of the triple bond and substituent 
groups is analogous to that used for the corresponding alkenes. 

CH 3 CH 3 

I I 

CH,-C-C=C — C-CH 3 

I I 

CH 3 H 

2,2,5-trimethyl-3-hexyne 
(isopropyl-(-butylacetylene) 

Open-chain hydrocarbons with more than one triple bond are called 
alkadiynes, alkatriynes, and so on, according to the number of triple bonds. 
Hydrocarbons with both double and triple bonds are called alkenynes, 
alkadienynes, alkendiynes, and so on, also according to the number of double 
and triple bonds. The order enyne (not ynene) is used when both double and 
triple bonds are present: 

HC = C-Cs=CH H 2 C = CH-C = CH 

butadiyne butenyne 

HC = C — CH=CH— CH=CH 2 HC=C — C = C — CH=CH 2 

1 ,3-hexadien-5-yne 1 -hexen-3,5-diyne 

The hydrocarbon substituents derived from alkynes are called alkynyl 
groups : 

HC = C— HC=C-CH 2 — 

ethynyl 2-propynyI 

(propargyl) 



chap 5 alkynes 112 

5-2 physical properties of alkynes 

Alkynes generally have physical properties rather similar to the alkenes and 
alkanes, as can be seen by comparing the boiling points of C 2 and C 6 repre- 
sentatives of these three classes of hydrocarbon : 



bp 68.7° 



Alkynes, like other hydrocarbons, are almost completely insoluble in water. 



CH3-CH3 CH 2 = CH 2 


HC^CH C 


bp -88.6° bp-105° 


bp -84° 


CH 3 CH 2 CH 2 CH 2 CH=CH 2 


CH 3 CH 2 CH 2 CH 2 C = CH 


bp 63.5° 


bp71.5° 



5-3 ethyne 



The simplest alkyne, HC^CH, is of considerable industrial importance and 
usually goes by the trivial name acetylene, rather than by the systematic 
name ethyne, which will be used here. Ethyne is customarily obtained on a 
commercial scale by hydrolysis of calcium carbide (CaC 2 ) or, in low yield, by 
high-temperature cracking (or partial combustion) of petroleum gases, 
particularly methane. Calcium carbide is obtained from the reaction of 
calcium oxide with carbon at about 2000° : 

2000° 

CaO + 3 C ► CaC 2 + CO 

It is cleaved by water (acting as an acid) to give ethyne and calcium hydroxide : 

CaC 2 + 2H 2 ► HC-CH + Ca(OH) 2 

Ethyne is much less stable with respect to the elements than ethene or ethane : 

HC=CH (g) — 

H 2 C=CH 2 (<?) 

H 3 C-CH 3 (<7) 



2 C (s) + H 2 (g) 


AH = 


-54.2kcal 


2C(s) + 2H 2 (g) 


AH = 


-12.5kcal 


2C(s) + 3H 2 (g) 


AH = 


+ 20.2 kcal 



An explosive decomposition of ethyne to carbon and hydrogen may occur 
if the gas is compressed to several hundred pounds per square inch (psi). Even 
liquid ethyne (bp — 83°) must be handled with care. Ethyne is not used com- 
mercially under substantial pressures unless it is mixed with an inert gas and 
handled in rugged equipment with the minimum amount of free volumes 
Large-diameter pipes for transmission of compressed ethyne are often packed 
with metal rods to cut the free volume. Ethyne for welding is dissolved under 

O 

II 
200 psi in acetone (CH 3 — C— CH 3 , bp 56.5°) and contained in cylinders 

packed with diatomaceous earth. 

Flame temperatures of about 2800° can be obtained by combustion of 
ethyne with pure oxygen. It is interesting that ethyne gives higher flame temp- 
eratures than ethene or ethane even though ethyne has a substantially lower 
heat of combustion than these hydrocarbons. The higher temperature of 



sec 5.4 addition reactions of alkynes 113 

ethyne flames, compared with those of ethene or ethane, is possible despite 
the smaller molar heat of combustion : 

C 2 H 2 (g) + i 2 (g) ► 2 C0 2 (g) + H 2 (I) AH = -31 1 kcal 

C 2 H 4 (<?) + 3 2 (g) ► 2 C0 2 (g) + 2H 2 (/) AH = -337 kcal 

C 2 H 6 (g) + 1 2 (g) ► 2 C0 2 (g) + 3H 2 (/) A//= -373 kcal 

This is because the heat capacity of the products is less. Less water is formed 
and less of the reaction heat is used to bring the combustion products up to 
the flame temperatures. Alternatively, what this means is more heat liberated 
per unit volume of stoichiometric hydrocarbon-oxygen mixture. The com- 
parative figures are ethyne, 3.97; ethene, 3.76; and ethane, 3.70 kcal/liter of 
gas mixture at standard temperature and pressure. 

5-4 addition reactions of alkynes 

That ethyne (acetylene) undergoes addition reactions with one or two moles of 
hydrogen or with halogens such as bromine was discussed in Chapter 2. The 
higher alkynes react similarly as can be illustrated with 4,4-dimethyl-l- 
pentyne : 

CH 3 CH 3 

I H, I 

CH,-C-CH 2 -C = CH ^— ♦ CH 3 -C-CH 2 -CH = CH 2 

\ Ni I 

CH 3 CH 3 

4,4-dimet hyl- 1 -pentyne 4,4-dimethyI- 1 -pentene 



I 
CH 3 -C-CH 2 -CH 2 -CH 3 

CH 3 

2,2-dimethylpentane 



CH 3 CH 3 

_ C = CH -Jfi^ CH 3 -C-CH 2 -CBr = CHBr 
I 
CH 3 

4,4-dimethyl- 1 ,2-dibromo- 1 -pentene 
(trans isomer formed predominantly) 



Br 2 



CH 3 
I 
— C— CH 2 — CBr 2 - CHBr 2 

CH 3 

1 , 1 ,2,2-tetrabromo-4,4- 
dimethylpentane 



chap 5 alkynes 114 

Alkynes undergo addition reactions with many other reagents which add to 
alkenes, particularly those which are electrophilic. The susceptibility of alkenes 
to electrophilic reagents was explained earlier on the basis of repulsions 
between the electrons in the double bond (Section 4-4), and you might rea- 
sonably expect alkynes, with triple bonds, to be even more susceptible to 
electrophilic attack. Actually, however, the reaction rates of alkynes with elec- 
trophilic reagents are rather less than those of alkenes. Whereas the deep 
color of liquid bromine is almost instantly discharged when it is added to an 
alkene, the bromine color persists for a few minutes with most alkynes. 
Similarly, the addition of water (hydration) to alkynes not only requires the 
catalytic assistance of acids (as do alkenes), but also mercuric ions: 



HC = CH 



H,0 



H 2 S0 4 
HgS0 4 ' 



OH 

vinyl alcohol 
(unstable) 



/ 

H 

acetaldehyde 



Mercuric, cuprous, and nickel ions are often specific catalysts for reactions 
of alkynes, perhaps because of their ability to form complexes with' triple 
bonds. 



Hg 2 



R-C=C-R + He 



± R — CssC— R 



The product of addition of one molecule of water to ethyne is unstable and 
rearranges to a carbonyl compound, acetaldehyde. With an alkyl-substituted 
ethyne, addition of water always occurs in accord with Markownikoff's rule: 



CH 3 -C=CH + H 2 

propyne 



H 2 S0 4 
HgSoT 



OH 
I 
CH 3 — C = CH 2 



O 

II 

CH 3 -C— CH 3 

acetone 



Alkynes react with potassium permanganate with formation of man- 
ganese dioxide and discharge of the purple color of the permanganate ion just 
as do alkenes but, again, the reaction is generally not quite as fast with alkynes. 
Hydrogen halides also add to the triple bond. These additions, like the ones to 
alkenes, occur in accord with Markownikoff's rule: 



HC=CH 



HF 



H 2 C = CHF 

fluoroethene 
(vinyl fluoride) 



♦ CH 3 -CHF 2 

1,1-difluoroethane 



Ethyne dimerizes under the influence of aqueous cuprous ammonium 
chloride. This reaction is formally analogous to the dimerization of 2-methyl- 
propene under the influence of sulfuric acid (see Section 4-4H), but the details 



sec 5.4 addition reactions of alkynes 115 

of the reaction mechanism are not known : 

H H 

Cu(NH 3 ) 2 ®Cl e \ I 

2HC=CH ► C=C— C = CH 

/ 
H 

butenyne 
(vinylacetylene) 

Alkynes, like alkenes, react with boron hydrides by addition of B— H 
across the carbon-carbon triple bond (Section 4-4F) and give vinylboranes : 

CH,CH, H 
\ / 

CH,CH 2 — C = C — H + BH 3 ► C = C 

/ \ 

H BH 2 

(a vinylborane) 

Vinylboranes react readily with acetic acid under mild conditions to give 
alkenes. The overall process is quite stereospecific, for a disubstituted alkyne 
gives only a cis alkene. Evidently the boron hydride adds in a cis manner to 
the triple bond, and the vinylborane produced then reacts with acid to give the 
corresponding cis alkene. 

CH 3 CH 2 H H 

I \/ 

C B 

III + I 

C H 

I 
CH 3 CH 2 

3-hexyne 

Whereas nucleophilic reagents do not generally add to alkenes, they add 
readily to alkynes, particularly to conjugated diynes and triynes. For example, 
1,3-butadiyne adds methanol in the presence of a basic catalyst such as 
sodium hydroxide. The mechanism of this type of reaction resembles the 

NaOH 

HC=C— C=CH + CH3OH > CH 3 OCH=CH— C^CH 

1,3-butadiyne l-methoxy-l-buten-3-yne 

ionic addition reactions discussed previously (Section 4- 4A) except that the 
initial step involves attack of the nucleophile (CH 3 O e ) on a terminal carbon 
to form a carbanion intermediate. The nucleophile is initially formed by the 
reaction of methanol with the basic catalyst. The carbanion intermediate is a 



H 






CH 3 CH 2 ^ 

II 
CH 3 CH 2 H 


H 


CH 3 CH 2 H 
CH 3 C0 2 H C 

— ~ II 
c 

CH 3 CH 2 ^H 

90 % c/.s-3-hexene 



CH 3 OH+NaOH r 



CH 3 O e + HC^C-C=CH •■ CH 3 0-CH=C-C=CH 

very strong base and reacts rapidly with methanol to remove a proton and 
reform the nucleophile, CH 3 O e , and generate the product, 1-methoxy-l- 
buten-3-yne. 

x a 

CH 3 0— CH=C— C=CH + CH 3 OH ► CH 3 OCH=CH-C=CH+ CH 3 O e 



chap S alkynes 116 

5-5 alkynes as acids 

A characteristic and synthetically important reaction of 1 -alkynes is salt (or 

"alkynide") formation with very strong bases. Calcium carbide, CaC 2 , 

can be regarded as the calcium salt of ethyne with both hydrogens removed, 

/e e\ 
Ca 2 ®\C=C/. Thus, the alkynes behave as acids in the sense that they give up 

protons to suitably strong bases: 

liquid 
ffi e NH-, e ffi 

R-C = C:H +K:NH 2 ^_ ' R— C = C:K + :NH 3 

R = H or alky] a potassium. alkynide 

Alkynes are much less acidic than water. In other words, water is much 
too weakly basic to accept protons from 1 -alkynes; consequently, even if 
alkynes were soluble in water, no measurable hydrogen-ion concentration 
would be expected from the ionization of 1 -alkynes in dilute aqueous solu- 
tions. However, 1-alkynes are roughly 10 13 times more acidic than ammonia, 
and alkynide salts are readily formed from 1-alkynes and metal amides in 
liquid ammonia. 

Alkynes are at least 10 18 times more acidic than ethene or ethane. The 
high acidity of ethyne and 1-alkynes relative to other hydrocarbons can 
be simply explained in terms of lower repulsion between the electron pair 
of the C — H bond of 1-alkynes and the other carbon electrons. In the triple 
bond, three pairs of bonding electrons are constrained to orbitals between 
the two carbon nuclei. As a result, they are, on the average, farther away 
from the C — H electron pair than are the C — C electrons from the C — H 
electron pairs in alkenes or alkanes. 



Consequently, less electron repulsion is expected for the C — H electron pair 
of a 1-alkyne: the less the electron repulsion, the more closely the C — H 
electrons will be held to the carbon nucleus ; and the more strongly they are 
held to carbon, the more easily the hydrogen can be removed as a proton by a 
base. By this reasoning, 1-alkynes are expected to be stronger acids than 
alkenes or alkanes. It should be clear from this discussion that the ability of 
an atom to attract bonding electrons — its electronegativity (Section 1 • 2) — 
will depend on whether it is singly, doubly, or triply bonded. For carbon, a 
triply bonded atom will have the highest electronegativity, and a saturated 
carbon the lowest. 

In terms of degree, the very much larger acidity of alkynes, compared with 
alkanes, is easily understood if you remember that electrostatic forces depend 
upon the inverse square of the distance. A small displacement of electrons will 
cause a very large electrostatic effect at the short distances which correspond 
to atomic diameters. 

A simple and useful chemical test for a 1-alkyne is provided by its reaction 



sec 5.6 synthesis of organic compounds 117 

with silver ammonia solution. Alkynes with a terminal triple bond give 
solid silver salts, while disubstituted alkynes do not react : 

R-C = C-H + Ag(NH 3 ) 2 ► R-C = C-Ag(j) + NH 3 + NH 4 

(R = H or alkyl) silver alkynide 

The silver alkynides appear to have substantially covalent carbon-metal bonds 
and are not true salts like calcium, sodium, and potassium alkynides. Silver 
ammonia solutions may be used to precipitate terminal alkynes from mixtures 
with disubstituted alkynes. The monosubstituted alkynes are easily regener- 
ated from the silver precipitates by treatment with mineral acids or sodium 
cyanide : 



R— C=C— Ag + NaCN + H 2 > R-C=C-H + NaOH + AgCN(s) 

It should be noted that dry silver alkynides may be quite shock sensitive and 
can decompose explosively. 



5-6 synthesis of organic compounds 



In this chapter we have described the chemistry of one of the important 
reactive groups (or "functional groups") in organic chemistry, the carbon- 
carbon triple bond. The previous chapter covered another important func- 
tional group, the carbon-carbon double bond. The numerous reactions of 
these groups are part of the complex web of reactions that allows us to convert 
one compound to another, often to one which has a quite different structure, 
by utilizing a number of reactions in sequence. Choosing the best route to 
convert compound A to compound Z is the forte of the synthetic organic 
chemist and, when he combines a high degree of intellectual skill in sifting 
through the many possible reactions with a high degree of skill in the labor- 
atory, his work can fairly be described as both elegant and creative. 

The purposes of syntheses are widely divergent: Thus one might desire to 
confirm by synthesis the structure of a naturally occurring substance, and at 
the same time, develop routes whereby analogs of it could be prepared for 
comparisons of chemical and physiological properties. Another aim might be 
to make available previously unreported substances that would be expected 
on theoretical grounds to have unusual characteristics because of abnormal 
steric or electronic effects, for example, the following compounds: 









H 


C(CH 3 ) 3 






1 
C 


(CH 3 ) 3 C-C-C(CH 3 ) 3 


C=C 

\ / 


C(CN) 4 


H"Cd 7 C- 


C(CH 3 ) 3 


CH 2 




c 




cyclopropyne 


tetracyanomethane 


H 


tetra-/-butylmethane 






tetrahedrane 



chap 5 alkynes 118 

Much research is also done to develop or improve processes for synthesizing 
commercially important compounds : in such work, economic considerations 
are obviously paramount. 

Regardless of why a compound is synthesized, the goal is to make it from 
available starting materials as efficiently and economically as possible. 
Naturally, what is efficient and economical in a laboratory-scale synthesis 
may be wholly impractical in industrial production; and, while we shall 
emphasize laboratory methods, we shall also indicate industrial practices in 
connection with the preparation of many commercially important substances. 

An essential difference between industrial and laboratory methods is that 
the most efficient industrial process is frequently a completely continuous 
process, in which starting materials flow continuously into a reactor and 
products flow continuously out. By contrast, research in a laboratory is 
usually unconcerned with sustained production of any single substance, and 
laboratory preparations are therefore normally carried out in batches. 
Another difference concerns by-products. In laboratory syntheses, a by- 
product such as the sulfuric acid used to hydrate an alkene is easily disposed 
of. But, on an industrial scale, the problem of disposal or recovery of millions 
of pounds per year of spent impure acid might well preclude use of an other- 
wise satisfactory synthesis. 

We believe that it is important in writing out projected syntheses to specify 
reagents and reaction conditions as closely as possible because different sets 
of products are sometimes formed from the same mixture of reagents, de- 
pending upon the solvent, temperature, and so on. The addition of hydrogen 
bromide to alkenes (Section 4-4E) provides a cogent example of how a change 
in conditions can change the course of a reaction. 

Most syntheses involve more than one step — indeed, in the preparation of 
complex natural products, it is not uncommon to have 30 or more separate 
steps. The planning of such syntheses can be a real exercise in logistics. The 
reason is that the overall yield is the product of the yields in the separate 
steps ; thus, if each of any three steps in a 30-step synthesis gives only 20 % of 
the desired product, the overall yield is limited to (0.20) 3 x 100 = 0.8% even 
if all the other yields are 100 %. If 90 % yields could be achieved in each step, 
the overall yield would still be only (0.90) 30 x 100 = 4%. Obviously in a 
situation of this kind, one should plan to encounter the reactions that have 
the least likelihood of succeeding in the earliest possible stages of the syntheses. 

In planning multistep syntheses, you must have a knowledge of how 
compounds are formed and how they react. For example, if asked to convert 
compound A to compound C, one would quickly check the reactions of 
compounds of type A and the preparations of the compounds of type C to 
see if any of them coincide. If they do not, the next step is to see if there is a 
compound B that can be prepared from A which could itself give C. In this, 
one is guided by the changes, if any, that are required in the carbon skeleton. 
Some simple examples are shown. 



Possible starting materials: 



CH 3 
HC^CH CH 3 — CH— C = C— H any inorganic reagents 



sec 5.6 synthesis of organic compounds 119 

Desired products : 



CH 3 






CH 3 Br 

1 1 


CH-CH 2 - 


CH 2 Br 


CH 3 - 


-CH-C = CH 2 



CH 3 

A quick glance at the carbon skeletons of the starting materials and the de- 
sired products shows that the product listed first, butane, has a different 
carbon skeleton than that of either of the possible organic starting materials 
but that the other two, l-bromo-3-methylbutane and 2-bromo-3-methyl-l- 
butene, have the same skeleton as one of the possible starting materials, 
-3-methyl-l-butyne. Let us then begin with an example where no change in 
carbon skeleton is necessary and try to think of reactions which will in one 
step convert 3-methyl-l-butyne to the desired products. 

In going back over the previously discussed addition reactions, we note that 
addition of hydrogen bromide to a triple bond gives a bromoalkene (Section 
4-4). Furthermore, if this addition occurs in the Markownikoff manner, the 
third product would be obtained in one step : 

CH 3 CH 3 Br 

I II 

CH 3 — CH-C = CH + HBr — — — » CH 3 — CH-C = CH 2 

dark 

With regard to the second desired product, the bromoalkane, there is no 
reaction that we have thus far encountered (nor does one exist) that will 
convert a triple bond to this arrangement in one step. Alkenes, however, can 
be made by the catalytic hydrogenation of alkynes and one could thus convert 
the alkyne to the desired product by the following route : 

CH 3 CH 3 CH 3 

I H 2 I HBr I 

CH 3 -CH-C = CH > CH 3 -CH-CH = CH 2 — ->■ CH 3 -CH-CH 2 -CH 2 -Br 

Ni peroxides 

The other desired product, butane, requires a change in the carbon skeleton. 
The only such synthetic reaction that we have thus far encountered is the 
dimerization of ethyne under the influence of cuprous ion. Thus, butenyne 
can be made from ethyne and this, when hydrogenated, gives butane: 

Cu(NH 3 ) 2 e 3H 2 

2HC=CH > 



Ni 

Each of the reactions used in the above syntheses gives reasonably high 
yields of products so that troublesome separation of isomers is not necessary. 
We have met one reaction thus far which is not specific and which usually 
leads to the production of isomeric mixtures — the halogenation of alkanes. If 
we had included 2-methylbutane in the list of available starting materials for 
the above syntheses one might have been tempted to try to prepare 1-bromo- 
3-methylbutane by light-catalyzed bromination because the carbon skeletons 
of these two compounds are the same : 

CH 3 CH 3 

CH,-CH-CH,-CH 3 + Br 2 > CH 3 — CH-CH 2 -CH 2 Br + HBr 



chap 5 alkynes 120 

Unfortunately, there is no way to ensure that substitution will occur at the de- 
sired place in the alkane. Indeed, you can be sure that a mixture of bromoalkanes 

CH 3 
I 
will be produced with the principal product being CH 3 — C— CH 2 — CH 3 , 

Br 

resulting from substitution at the tertiary carbon atom. Thus, it is 
important to consider the "practicality of synthetic schemes that you devise. 
The extremely important matters of identification and purification of products 
are considered in Chapter 7. 

It should be clear that a reaction of one type of compound may well be a 
method of preparation of another. So far, we have considered the reactions 
of alkanes, alkenes, and alkynes and have not listed separately their methods 
of preparation. There are two reasons for this. First, some of these compounds, 
particularly alkanes, can be obtained quite pure by careful fractionation of 
petroleum and we would seldom need to prepare them in the laboratory. 
Second, and more important, we have as yet encountered only a few of the 
many functional groups which can efficiently be converted to alkanes, alkenes, 
and alkynes. However, when we discuss a new functional group we shall list 
for purposes of convenience the common preparative methods that lead to it 
even though they may involve reactions or types of compounds to be described 
later in the book. 

In the interest of completeness, some useful methods of forming carbon- 
carbon single, double, and triple bonds are summarized below. 

C-C bonds: 

Addition, as of hydrogen to C=Cor C=C bonds (Sections 4-4, 5-4) 
C=C bonds: 

a. Dehydration of alcohols (Section 10-5B) 

b. Elimination of hydrogen halides from haloalkanes and related 
compounds (Sections 8-12, 8-13) 

c. Partial additions, such as catalytic hydrogenation, to C=C bonds 
(Section 5-4) 

C^C bonds: 

a. Elimination of two moles of hydrogen halide from 1,2-dihaloalkanes 
(Section 9-5) 

b. Elimination of one mole of hydrogen halide from haloalkenes (Section 
9-5) 



summary 

Alkynes possess carbon-carbon triple bonds. Simple open-chain alkynes have 
the formula C„H 2 „_ 2 and, in the IUPAC system, are named the same way as 
alkenes except that the -ene ending becomes -yne. Ethyne, HC=CH, is usually 
called acetylene. The physical properties of alkynes, alkenes, and alkanes are 
similar. Chemically, alkynes resemble alkenes in undergoing addition reactions 
with electrophilic reagents although they do not usually react as rapidly as 



exercises 12 1 



alkenes. However, they will undergo additions with nucleophilic reagents 
more readily than alkenes. Several typical addition reactions of alkynes are 
illustrated with propyne as an example. The final reaction produces cis 
alkenes when nonterminal triple bonds are reduced. 



CH 3 C=CH " 2 > CH 3 CH=CH 2 



H 3 C Br 

Br 2 \ / 
> c = c 

(other halogens / \ 

behave similarly) Br H 



— ► CH 3 CC1=CH 2 



(other hydrogen halides 
behave similarly) 

H 2 _ CH3 _^_ CH3 



H®, Hg 2 

H 3 C H 

B 2 H 6 \ / acid 

> C = C > CH 3 CH=CH 2 

/ \ 

H BH 2 



Ethyne can be dimerized to butenyne (vinylacetylene) by the action 
of ammoniacal cuprous ion : 



2 HC3C H Cu(NH3)2a > CH 2 = CH-C^CH 



1 -Alkynes are feebly acidic and can be converted to anions (alkynide ions) 
by the action of powerful bases. 



-»• RC = C e + NH, 



Insoluble silver salts are produced by the action of ammoniacal silver ion on 
1 -alkynes and this is a useful means of detecting terminal triple bonds. 

Some of the principles of organic synthesis have been discussed with the aid 
of several specific examples. 



exercises 

5-1 Name each of the following substances by the IUPAC system and as a 
substituted acetylene. 

a. CH 3 C=CC1 HX-CH HC-CH 2 

b. (CH 3 )3CC=CCH 2 C(CH3) 3 e. H 2 C^ C-C = C-C^ CH 2 

c. CH 2 =CHCH 2 C=CCH=CH 2 H 2 C-CH 2 H 2 C-CH 2 

d. HC=CCH 2 C=C-C(CH 3 )=CH 2 /^^c 

/• (CH 2 ) 8 III 

V c 



chap 5 alkynes 122 

5-2 Write structural formulas for each of the following, showing possible geo- 
metrical isomers : 

a. 1 ,2-dibromocyclopropane 

b. 2,4-hexadiene 

c. cyclooctyne 

d. dibromoethyne 

e. l,5-hexadien-3-yne 

5-3 Calculate AH values from the bond-energy table in Chapter 2 for the follow- 
ing reactions in the vapor state at 25° : 

a. HO=CH + Br 2 > CHBr=CHBr 

b. CHBr=CHBr+Br 2 > CHBr 2 -CHBr 2 



c. 



HC=CH + H 2 ► CH 2 =CHOH 



P 

d. CH 2 = CHOH ► CH 3 C 

H 

e. 2HC=CH ► H 2 C=CH-C=CH 

/ CH 3 -C=C-H ► CH 2 =C=CH 2 

g. Calculate also a AH for reaction f from the experimental AH values for 
the following reactions : 

CH 3 C=CH + 2H 2 ► CH 3 CH 2 CH 3 AH = -69.7 kcal 

CH 2 =C=CH 2 + 2H 2 * CH3CH2CH3 A J ff=-71.3kcal 

Explain why the value of AH calculated from bond energies might be un- 
reliable for the last reaction. 

5-4 Ethyne has an acid ionization constant (K U a) of ~ 10~ 20 in water. 

a. Calculate the concentration of ethynide ion expected to be present 
in 1 M solution of aqueous potassium hydroxide that is 10 " 4 M in 
ethyne (assuming ideal solutions). 

b. Outline a practical method (or methods) that you think might be 
suitable to determine an approximate experimental value of K H a for 
ethynes, remembering that water has a K H a of about 10 ~ 16 . 

c. Would you expect H — C=N to be a stronger acid than H — C= C — H ? 
Why? 

5-5 Suppose you were given four unlabeled bottles, each of which is known to 
contain one of the following compounds: «-pentane, 1-pentene, 2-pentyne, 
and 1-pentyne. Explain how you could use simple chemical tests (preferably 
test tube reactions that produce visible effects) to identify the contents of 
each bottle. (Note that all four compounds are low-boiling liquids.) 

5-6 How would you distinguish between the compounds in each of the following 
pairs using chemical methods (preferably test tube reactions). 

a. CH 3 CH 2 C=CH and CH 3 C=CCH 3 

b. CH 3 CH 2 C=CH and CH 2 =CH-CH=CH 2 

c. C 6 H 5 C=CC 6 H S and C 6 H 5 CH 2 CH 2 C 6 H 5 

5-7 Write balanced equations for the reaction of 3,3-dimethyl-l-butyne with each 



exercises 123 

of the following reagents and, in the first three cases, name the product. 

a. H 2 (two moles), Ni c. Cl 2 (one mole) 

b. HC1 (two moles) d. H 2 0, H e , Hg 2 ® 

5-8 Show how each of the following compounds could be synthesized from the 
indicated starting material and appropriate inorganic reagents. Specify the 
reaction conditions, mention important side reactions, and justify the practi- 
cality of any isomer separations. 

CI 
I 

a. CH 2 =CH— C=CH 2 from ethyne (the product is an intermediate in 

the synthesis of the artificial rubber, neoprene) 

b. CH 3 CHFCH 2 Br from propyne 

c. CH 3 CH 2 COCH 3 from 1-butyne 

CH 3 CH 3 

\ 3 / 3 

d. C=Q from 2-butyne 

H H 

CH 3 CH 2 H from ,. but y ne and 

e. C=C deuteroacetic acid 
H / \ D (CH 3 C0 2 D) 

5-9 Starting with ethyne, 3-methyl-l-butyne, and any inorganic reagents (the 
same starting materials used in the examples on pp. 118-119) show how you 
could prepare the following compounds. 

CH 3 
I 

a. CH 3 -CH-CHBr-CH 3 

CH 3 
I 

b. CH 3 -CH-CC1 2 -CH 2 C1 

CH 3 O 
I I! 

c. CH 3 -CH-C-CH 3 

d. CH 3 CH 2 CHOHCH 3 

5-10 When 0.100 g of an unsaturated hydrocarbon was treated with an excess of 
hydrogen in the presence of a platinum catalyst, 90.6 ml of hydrogen was 
absorbed at atmospheric pressure and 25°. Furthermore, the compound gave 
a precipitate with ammoniacal silver nitrate. What is its structural formula ? 

5-11 Indicate how you would synthesize each of the following compounds from 
any one of the given organic starting materials and inorganic reagents. 
Specify reagents and the reaction conditions, and justify the practicality of 
any isomer separations. If separations are not readily possible, estimate the 
proportion of the desired compound in the final product. Starting materials : 
ethene, propene, isobutane, 2-methylpropene. 

CH, 



OH 



CH, 



b. CH 3 — C-CH 3 



CH 3 CH-CH 2 

c I I 

OH Br 

CH, CH 3 

I I 

d. CH 3 -C — CH 2 -C— CH, 
I I 

CH, I 



chap 5 alkynes 124 



CH 3 
I 
e. CH 3 -C-CH 2 Br 

I 
H 



CH 3 CH 3 

I I 

/. C1CH 2 -C-CH 2 -CH-CH 3 

CH 3 



g. CH 3 — CH 2 — CH 2 OH 



* o t 

u :::: mm-' 

. £* <S i 



chap 6 bonding in conjugated unsaturated systems 127 

The nomenclature of alkenes, including those with more than one carbon- 
carbon double bond, was discussed in Chapter 4. The compounds with two 
double bonds separated by just one single bond were categorized as con- 
jugated dienes; those with double bonds separated by more than one single 
bond, as isolated dienes. Dienes with isolated double bonds have properties 
similar to those of simple alkenes except that there are two reactive groups 
instead of one. Thus, 1,5-hexadiene reacts with one mole of bromine to form 
5,6-dibromo-l-hexene and with two moles to give 1 ,2,5,6-tetrabromohexane. 

CH 2 = CH-CH 2 -CH 2 -CH=CH 2 ^ > CH 2 = CH-CH,-CH 2 - CHBr-CH 2 Br 
'^—> BrCH 2 — CHBr — CH 2 — CH 2 — CHBr — CH 2 Br 

Furthermore, the heat of hydrogenation of 1,5-hexadiene is almost exactly 
double that of the heat of hydrogenation of 1-hexene, indicating the normalcy 
of each of the double bonds. 

On the other hand, 2,4-hexadiene, the conjugated isomer, has properties 
which indicate that two double bonds are not wholly independent of one 
another. Addition of one mole of bromine produces a mixture of products. 

Br Br 

I I 

CH 3 — CH — CH=CH— CH — CH 3 

2.5-dibromo-3-hexene (major product) 

Br Br 
I I 

CH 3 — CH = CH — CH — CH — CH 3 

4.5-dibromo-2-hexene (minor product) 

The major product results from addition not to adjacent positions (1,2 
addition) but to positions separated by two intervening carbon atoms (1,4 
addition, or conjugate addition). Furthermore, hydrogenation of one mole of 
2,4-hexadiene liberates about 6 kcal less heat than hydrogenation of one 
mole of 1,5-hexadiene. 

A more striking anomaly is the case of benzene. This compound is a liquid 
hydrocarbon of formula C 6 H 6 . It is now known from a variety of experi- 
mental studies to be a cyclic molecule in the shape of a flat hexagon. The only 
way that carbon can preserve a tetravalent bonding arrangement here is by 
having three double bonds in the ring; for example, 

H 

I 
H C H 

^C^ ^C^ (T^*! (where a carbon and hydrogen atom 

|| I or I I are understood to be at each corner 

,<X ^>C \^ of the hexagon) 

H C H 

I 
H 

This structure for benzene was advanced in 1865 by August Kekule, only a 
few years after his postulate of tetravalent carbon. Those were the years in 



CH 3 — CH = CH— CH = CH — CH 3 + Br 2 




chap 6 bonding in conjugated unsaturated systems 128 

which enormous strides were made in establishing structural organic chem- 
istry as a sound branch of science. The objections to Kekule's structure that 
were soon put forward centered on the lack of reactivity of benzene toward 
reagents such as bromine. A number of alternative proposals were made 
in the following five years, of which the structures proposed by Ladenburg, 
Claus, and Dewar received the most attention. The Ladenburg and Dewar 





Ladenburg (1869) Claus (1867) 



Dewar (1867) 



formulas will be met again later in the book. The Claus structure is difficult 
to formulate in modern electronic theory and must be regarded as an attempt 
to formulate a substance with formula C 6 H 6 (no carbon is implied at the 
center of the ring) with all saturated tetravalent carbons. Actually, another 
40 years were to pass before the electron was discovered and an additional 20 
before the need would be recognized to identify bonds with pairs of electrons. 
Over this period, Kekule's structure for benzene was generally accepted, 
despite the dissimilarity between the reactivities of benzene and the alkenes. 
It was thought that the conjugated arrangement of the bonds must somehow 
be responsible for its inertness. Whereas the orange color of bromine vanishes 
instantly when cyclohexene and bromine are mixed, a benzene-bromine 
mixture remains colored for hours. When the mixture eventually becomes 
colorless, an examination of the reaction product shows it to be not the addi- 
tion compound, but a substitution compound. * Clearly benzene is much more 



+ Br, 



fast 




Br 



Br 



cyclohexene 
(C ( ,H 10 ) 



1 ,2-dibromocyclohexanc 
(C 6 H 10 Br,) 



+ Br, 



benzene 
(QH 6 ) 



slow 




Br 



+ HBr 



bromobenzene 
(C„H 5 Br) 



resistant to addition than cyclohexene, which means that its three conjugated 
double bonds constitute an unusually stable system for a triene. The stabiliza- 
tion of benzene is further indicated by a 37 kcal/mole lower heat of combus- 
tion than calculated for a molecule containing three ordinary double bonds. 
The heat of combustion of cyclohexene, on the other hand, is quite close to 



1 The general formula for a monosubstituted benzene is C 6 H 5 X. The group C 6 H 5 — is 
known as the phenyl group and the hydrocarbon C 6 H 5 CH=CH 2 is by the IUPAC system 
phenylethene, although commonly called styrene. 



sec 6.1 bonding in benzene 129 

that calculated using the standard table of bond energies (Table 2-1). The 



15 

+ _ o, ► 6 C0 2 + 3 H 2 A// e = —759.1 kcal 37 kcal 

%/* AW calc = -796.5 kcal difference 



17 



-» 6 C0 2 + 5 H 2 A W exp = -849.6 kcal 2 kcal 

AW cak = -851.5 kcal difference 



degree of stabilization of benzene compared to what might be expected for 
cyclohexene, about 35 kcal/mole, is a large quantity of energy. This is, of 
course, much larger than the 6 kcal/mole stabilization of 2,4-hexadiene; but 
even 6 kcal/mole can give rise to important chemical effects because energies 
are related to equilibrium constants (and, in a crude way, to rate constants) 
logarithmically (Sections 2- 5 A and 8-9). 

What is the origin of the stabilization associated with these conjugated 
systems and how does 1,4 addition occur? A qualitative answer to these 
questions is given in the next two sections. 



6-1 bonding in benzene 

Formation of an ordinary covalent bond between atoms A and B results in 
release of energy (stabilization) because the electrons that were localized on 
A and B can now interact with two nuclei instead of one : 

A- + B- * A:B 

The atomic orbitals on A* and B- occupied by the single electrons are de- 
signated by symbols (s, p, d, f) that are relics of the early days of atomic 
spectroscopy and serve now to indicate energy levels and orbital shapes. For 
example, s orbitals are spherical and an electron in an s orbital is of lower 
energy than an electron in a p orbital, which is dumbbell shaped (Figure 6T). 
Quantum theory tells us that only certain total energies are allowed of a 
system made up of a nucleus and an electron. The diffuse regions of space in 
which the electrons move are called orbitals, and the system has a specific 
energy when an orbital is occupied by one electron. An orbital can be des- 
cribed qualitatively by its approximate shape and quantitatively by rather 



Figure 6-1 Shapes of s and p orbitals. 






s atomic orbital p atomic orbital 



chap 6 bonding in conjugated unsaturated systems 130 

complex mathematical expressions that give the exact energies when the 
orbital is occupied by electrons. Because electrons have some of the character- 
istics of waves, the equations which define the electron distributions and their 
energies are called wave equations. While the equations describing one electron 
and one nucleus are not unduly complex, no rigorous solutions have so far 
been possible for molecules as simple as methane. 

Chemistry is almost entirely devoted to the study of molecules and, except 
for the noble gases, it is difficult to obtain a stable system for study of isolated 
atoms. Thus, elemental carbon exists not as an assemblage of isolated atoms 
but in the form of diamond or graphite, the atoms of which are covalently 
bonded together just as tightly as are the carbons in ethane and benzene. It is 
therefore preferable to direct attention not to orbitals for electrons in atoms 
but orbitals for electrons in molecules, molecular orbitals. There is a stable 
molecular orbital for each C— H bond in methane; they are all equivalent and 
point to the corners of a regular tetrahedron. 2 The eight valence electrons 
of these atoms assume the configuration of lowest energy regardless of the 
shape of the orbitals occupied by electrons on the isolated atoms. Only two 
electrons can occupy the same orbital 3 and there will then be two electrons 
in each of the C— H molecular orbitals. The repulsion between the four pairs 
of electrons makes the tetrahedral arrangement the one of lowest energy or 
greatest stability. The resulting bonds are equivalent and those like them in 
other saturated compounds are often called sp 3 bonds, on the basis of a 
mathematical analysis of the relations of the molecular orbitals to the orbitals 
occupied by electrons on atomic carbon. 

With carbon-carbon double bonds, a dilemma arises. Should we regard 
the two bonds as equivalent with both being bent or should we imagine that 
one of them — the a (sigma) bond — occupies the prime space along the bond 
axis and the other — the n (pi) bond — the space above and below the plane 
defined by the other bonds to the double-bonded carbons? We saw earlier 
(Section 2-6) that either of these models accounts for the geometry of ethene 
although the second of these is less straightforward and harder to visualize 
than the first. However, most theoretical treatments of conjugated systems 
make use of the idea of clouds of % electrons above and below the a bond. 
The popularity of this approach stems to a great extent from the fact that if 
we ascribe the unusual properties of benzene exclusively to the % electrons, we 



2 Strictly speaking, a molecular orbital describes the interaction of an electron with every 
nucleus in the molecule. In the case of molecules such as alkanes, however, the pair of 
electrons in a C— H bond appears to interact almost entirely with the carbon and hydrogen 
nuclei that make up the bond. To put it in other terms, the electrons in the orbitals of the 
C— H bonds of methane can be called " localized " bonding electrons. 

3 Why only two electrons to an orbital ? The reasons are not simple or very intuitively 
reasonable. That only two electrons can occupy a given orbital is a statement of the Pauli 
exclusion principle. The basic idea is that only two nonidentical electrons can occupy an 
orbital. How can electrons be nonidentical? By having different spins. For electrons there 
are only two different possibilities, corresponding to right handed or left handed. Two 
right-handed electrons or two left-handed electrons cannot occupy the same orbital — only 
a right-handed and left-handed pair, like the nonidentical pairs of animals allowed on 
Noah's ark. 



sec 6.1 bonding in benzene 131 




Figure 6-2 Diagram of the diradical -CHj — CH 2 - with the planes of the CH 2 
groups set at right angles to one another. Each carbon atom has a p orbital 
containing one electron. 



greatly reduce the number of electrons we have to deal with. We shall make 
use of this approach (which should be regarded as a matter of convenience 
and not as revealed truth) in the subsequent discussion. The three a bonds to 
carbon in an alkene, often designated sp 2 bonds, are taken to be spread at 
angles of 120° to one another. 

The n bonding in ethene can be thought to arise as follows. Imagine the 
diradical -Cr^ - CH 2 - with the carbon atoms joined only by a single bond. 
A reasonable electronic formulation would be the one in which the carbons 
are trigonal, that is to say, the three covalent bonds to each carbon lie in a 
plane at angles of 120° to one another (the maximum bond angle). The two 
extra electrons are considered to be placed in dumbbell-shaped p orbitals, 
one on each carbon atom. If the planes of the two CH 2 (methylene) groups are 
set at right angles to one another, the interaction between the electrons in the 
p orbitals is minimized (Figure 6-2) because the electrons are as far away from 
one another as they can be. If the methylene groups are now rotated with 
respect to one another until the carbons and hydrogens all lie in the same 
plane, the p orbitals will became parallel to one another (indeed, overlap with 
each other) and interaction between the electrons will be at a maximum. If 
the electrons have the same spin they cannot occupy a molecular orbital made 
up of the two p atomic orbitals; they repel each other and give an arrangement 
less stable than the one with the CH 2 groups at right angles. The setup with 
unpaired electrons does not give a stable 7r bond. On the other hand, if the 
spins are paired, they form a n bond — both electrons have the same energy, 
occupy the same region of space (n molecular orbital), and interact with both 
nuclei (see Figure 6-3). 

Benzene can be described by a similar orbital arrangement. Imagine a 
hexagon of carbon atoms joined by single bonds (<r bonds, localized along 
each bond axis) with a singly occupied p orbital on each. 7r-Bond formation 
between adjacent pairs of p orbitals as in ethene would give a triene (Figure 
6-4). The high degree of symmetry in this molecule, however, allows 7r-bond 
formation between each carbon and both its neighbors. The result is a system 
where the n electrons are associated with more than two nuclei and can be 
fairly called delocalized. Association of a pair of electrons with two nuclei 
results in stabilization with formation of a localized bond; association with 
more than two nuclei leads to still greater stabilization and delocalized bonds. 
The extra degree of stabilization that can be ascribed to delocalized n bonds is 



chap 6 bonding in conjugated unsaturated systems 132 



a 

CD 


H .*•*«& 
\ »- 

"LI WS^* 


,H 




H^v 


i H 

- - ;S S| , H 


/ 
/ 
/ 
/ 
/ 


H 


"H 


s 
\ 




fr ' 


...;\, H 



Figure 6-3 The energy levels that result from rotating the CH 2 groups of 
•CH 2 — CH 2 - into the same plane. The lowest energy form, above, is CH 2 =CH 2 , 
ethene, with a n bond. 



reflected in benzene's inertness to bromine addition and in its low heat of 
combustion. The stabilization energy, often called derealization or resonance 
energy, amounts to more than 30 kcal/mole (the difference between the heat of 
combustion calculated on the basis of "normal" localized bonds and the 
experimental heat of combustion). 

How should we draw the bonding arrangement in benzene ? The following 
notations are often used to indicate the symmetry of the molecule and the 
delocalized character of some of the bonds : 



It is difficult, however, to determine the number of % electrons by inspection 
of such structures. Indeed, with some derivatives of benzene one can be 
seriously misled by such notation (Section 20TB). A more revealing repre- 
sentation of the structure of benzene is achieved by drawing the following 
resonance structures: 



Figure 6-4 Electron-pairing arrangements for benzene. 



sec 6.2 conjugate addition 133 

The double-headed arrow bears no relation to the symbol <± used to describe 
chemical equilibrium. Rather, it indicates the two directions in which % bonds 
can be formed by showing the separate structures we would write if derealiza- 
tion were not considered and each pair of electrons was restricted to binding 
only two nuclei (see Figure 6-4). Because the orbital arrangement permits 
extensive derealization, we know that benzene is not the structure represented 
by either formula but is actually a hybrid which embodies some of the character 
expected of each and is more stable than either of the structures would be, if 
each were to correspond to a real substance. For convenience, we draw just 
one of these structures 



to indicate benzene whenever there is no need to represent the bonding more 
precisely. 

There are many advantages to writing the various possible structures of a 
delocalized molecule or ion. You can account for all the electrons in a system 
at a glance and determine whether a given structure is that of a cation, an 
anion, or a neutral molecule, and whether or not it is a radical. Furthermore, 
the description is in terms of structures usually with ordinary valences of 
carbon, oxygen, nitrogen, and so on. This way of describing molecules with 
delocalized electrons, the resonance method, as it is usually called, has been 
extraordinarily successful at predicting molecular geometries and behavior, 
notwithstanding its apparent artificiality in using more than one structural 
formula to describe a single compound and its occasional failures (Section 
6-7). 



6-2 conjugate addition 



How does conjugate addition (1,4 addition; seep. 127) occur? 1,3-Butadiene, 
a typical conjugated diene, exists in two important planar conformations, the 



H H 

\ / 
C-C 

// \ 
H 2 C CH 2 


H CH 

\ // 

c-c 

// \ 

H 2 C H 


s- cis 


s-trans 



s-cis and the s-trans forms. Although rotation occurs rather rapidly about the 
central bond, these are the two favored conformations because they permit 
some degree of 7r-bond formation across what is normally written as a single 
carbon-carbon bond. 

We have seen that addition of bromine to double bonds can occur by way 
of ionic intermediates (Section 4-4B). If we examine the bonding in the ion 
produced by the attack of bromine on a conjugated diene, we can see how 



chap 6 bonding in conjugated unsaturated systems 134 

conjugate addition takes place with 1,3-butadiene, and that an intermediate 
bromonium cation can be produced as follows: 

► CH 2 = CH-CH-CH 2 Br + Br e 



(In the earlier discussion, the possibility was raised that bridged bromonium 
ions are intermediates in addition of bromine to alkenes. Such species may 
also be formed here but do not vitiate the following argument.) Neither the 
double-bond n electrons nor the cationic charge will be localized as shown 
above. There will be n bonding between three carbons as indicated by the 
resonance structures [1]. In accord with these structures, the positive charge 
will be divided between the 2 and the 4 carbon. The charge will not be expected 
to be significant on the third carbon from the bromine because any resonance 
structure which puts the charge there will preclude ft-bond formation between 
adjacent carbons. We sometimes speak of the spreading of the charge as 
" derealization of the charge." It should, however, be clear from what we 
have said that derealization of charge is a consequence of the derealization 
of the electrons by 7r-bond formation between two different pairs of carbons ; 



CH 2 =CH-CH-CH 2 Br < ► CH,-CH=CH-CH 2 Br 

4 3 2 1 [|] 

that is, derealization of positive charge and electron derealization are two 
sides of the same coin. 

Addition of bromide ion can occur at either of the two partially positively 
charged carbon atoms in ion [l], 4 which accounts for the mixture of products 
that is, in fact, formed. If initial addition of the Br® ion were to take place at 



BrCH,-CH=CH-CH,Br 



[CH 2 =CH-CH-CH 2 Br < ► CH 2 -CH=CH-CH 2 Br] + Br e 



CH 2 = CH— CHBr — CH 2 Br 

the 2 position of the diene, the charge would be localized at the 1 position and 
the n electrons between atoms 3 and 4 so that no resonance stabilization 
of the carbonium ion intermediate (CH 2 =CH— CHBr— CH 2 ®) would be 
expected. On the other hand, the carbonium ion formed by addition of a Br® 
ion to a terminal position of the diene has a substantial degree of resonance 
stabilization. 

If we could prepare separate ions corresponding to the two localized 
structures that we have written to represent the hybrid, we would expect them 
to have similar bond energies and similar arrangements of the atoms in space. 

4 Note the singular form ion, not ions. In the resonance hybrid shown above we have used 
two structures, each with localized bonding, to represent as nearly as we can the true struc- 
ture of the single ion which is the reaction intermediate. 



sec 6.3 stabilization of conjugated dienes 135 



The reason for the similar geometries is that the cationic carbons of carbonium 
ions tend to have their three bonds planar (Section 2-5C), as do alkenic carbon 
atoms. Thus, delocalized 7r-bond formation in the hybrid ion [2] can be seen 
to occur in ar completely natural way for the two structures. This is an espe- 
cially important point to which we will return later. (The ion formed from the 
■y-cwjEonformation of the diene is shown here.) 



H 

t 
H 



CH,Br 



< 



CKUBr 



H 






CH,Br 



[2] 

A cationic carbon atom, such as those depicted in each of the structures 
making up the hybrid, possesses only a sextet of electrons (the six in the three 
a bonds) and hence has an empty p orbital available to interact with the 
adjacent pair of n electrons. These electrons thus spread over three atoms 
instead of two, providing a delocalized u bond. The consequence is a more 
stable system than would be represented by either resonance structure alone. 
In the next section we shall see that the electron system of 1,3-butadiene, al- 
though conjugated, is delocalized to a much smaller extent. 



6-3 stabilization of conjugated dienes 

The special character of conjugated dienes is manifest in their tendency to 
undergo 1,4 addition and in their lower heats of combustion. Conjugate 
addition has been accounted for on the basis of resonance stabilization (and 
preferential formation) of an intermediate cation which gives rise to the 
1,4 product. This reaction does not tell us whether or not resonance stabiliza- 
tion is important in the diene. However, its low heat of combustion can be 
attributed to resonance in the diene because here we are dealing with the over- 
all stabilization of the diene itself, not of a product to which it may be convert- 
ed. The stabilization of 1,3-butadiene, as determined by the difference between 
calculated and experimental heats of combustion, is only about 6 kcal as 
compared to more than 30 kcal for benzene. Why should this be so ? Cannot % 
electrons of the two adjacent double bonds bind together all of the carbons 
in the same way as in benzene? Such an orbital arrangement for the s-cis 
conformation is shown in [3]. 







■•y-^tfr 



chap 6 bonding in conjugated unsaturated systems 136 

The reason for the resonance stabilization being small is readily apparent 
when you compare the possible ways of getting n bonding in butadiene with 
those for benzene. First, we write the basic structure with four singly occupied 
p orbitals: 

H H 

\ . ./ 

c-c 

./ \. 

H — C C — H 

\ / 
H H 

Now we consider % bonds to be made by pairing the electrons two different 

H H H H 



.c- 



(■■•'■c^i? — H h — Al- 






ways [4]. The first way corresponds to normal double bonds between the 
1,2 and 3,4 carbons while the second corresponds to a double bond between 
the 2,3 carbons but essentially no binding between the 1,4 carbons, which are 
impossibly far apart to participate in effective binding, even though the 1,4 
electrons are paired. Any attempt to increase the stabilization of s-cis- 
butadiene by bringing the 1,4 atoms closer together, thus increasing the 1,4 
interaction, would run afoul of increasing angle strain and increasing inter- 
action between the inside 1,4 hydrogens. The 2,3 % bonding does seem to 
increase the stability of butadiene to some extent. That it does not do more 
can be ascribed to the fact that 2,3 % bonding is associated with 1,4 "non- 
bonding." 

The above discussion can be rephrased in terms of resonance structures as 
follows : 

H H H H 

V / \ / 

c— c c=c 

// \ / \ 

H— C C— H < ► H-C C— H 

\ / \ / 

H H H H 

The " 1 ,4 7t bonding " is represented in the second structure by a dotted line 
(sometimes called a formal bond) to emphasize that the structure does not 
represent cyclobutene, which is a stable isomer of butadiene and differs from 

H H H 

HC = CH >=/=c / 

H 2 C-CH 2 \/_ j/ 

/ \ 

cyclobutene H H 



sec 6.4 stabilization of cations and anions 137 

butadiene in chemical behavior and in the arrangement of its atoms in space. 
Unlike s-cis or ,y-/ra«.s-butadiene, cyclobutene does not have all its carbons 
and hydrogens in a single plane. Also unlike butadiene, the bonding between 
the two methylene (CH 2 ) carbons results from a cr-type C— C bond of 
essentially normal length. 

We can summarize the above discussion by noting that the resonance 
method considers the binding which might be produced by pairing electrons 
in different ways for a given geometrical arrangement of the atoms. The 
predictive power of the resonance method is derived from the fact that for a 
given arrangement of the atoms, important contributions will be made to the 
hybrid structure only by those ways of pairing the electrons which correspond 
to reasonably feasible ball-and-stick models. 

The % bonds in the hybrid structure for butadiene can be represented by a 
combination of heavy and light dotted lines. 



weak n binding 
- strong 7t binding 




No 1,4 bonding is shown because of the large distance between the atoms. 

It is important to recognize that each of the resonance structures for 
1,3-butadiene (or benzene) has the same number of pairs of electrons. No 
structures need be considered which have a different number of paired 

electrons, such as CH 2 — CH=CH— CH 2 , where now the 1,4 electrons are 
taken to be unpaired (both right handed or both left handed ; see Section 6-1). 
Such structures correspond to a diradical form of butadiene which is known 
to exist but has entirely different properties and is grossly less stable. It 
has a n bond only between two adjacent carbons. 

— ► CH,-CH-CH=CH, < ► CH, = CH-CH-CH 2 



no 7t bonds because the 
electrons are unpaired 



6-4 stabilization of cations and anions 

You might well ask why we do not consider ionic-type resonance structures 
for 1,3-butadiene similar to those written for the conjugate addition inter- 
mediate in the previous section. 



CH, = CH-CH-CH 2 < ► CH, = CH— CH-CH, 



etc. 



chap 6 bonding in conjugated unsaturated systems 138 

The answer is that in the ionic structures we have substituted ionic bonding 

© f. 

for n bonding; and just as CH 2 — CH 2 is far from the best structure for CH 2 = 
CH 2 , so the above ionic structures are less important than the wholly 71-bonded 
CH 2 =CH— CH=CH 2 structure for butadiene. To put it in another way, 
carbon is intermediate in electronegativity and has little tendency to attract 
electrons and become negative or give up electrons and become positive. 

Thus, the ionic (or dipolar) structures are less favorable than the shared 

© 
electron structures. The cationic intermediate CH 2 =CH— CH— CH 2 Br 

e 
<h>CH 2 — CH=CH 2 — CH 2 Br is different in that there is no way that the 
charge can disappear by changing the 71-bond arrangements. This cation is 
in fact just one representative of the important class of allyl cations, the 
parent of which is 

CH 2 -CH=CH 2 < > CH 2 = CH-CH 2 = [CH 2 ^CH=-CH 2 ]® 

The allyl cation is a relatively stable carbonium ion as carbonium ions go 
and we shall meet it again as a reaction intermediate in Chapter 8. 

The bicarbonate ion is an example of an anion stabilized by resonance. 
Sodium bicarbonate, NaHCO s , is a salt whose ionic components are sodium 
ion (Na ffi ) and bicarbonate ion (HC0 3 e ). If you draw a structural formula 
for bicarbonate ion so that the carbon and oxygen atoms have octets of elec- 
trons, you obtain 

O 
H-O-Cf Na 9 

V 

Each of the two oxygens on the right-hand side of this formula for the bi- 
carbonate ion is bonded only to the carbon atom. Their electronic environ- 
ments appear to be different : one oxygen bears a full negative charge and the 
other no charge at all. There is, however, another way to arrange the electrons 
which reverses the roles of the two oxygens and gives additional n bonding. 
The actual structure is therefore that of a resonance hybrid with half of the 
negative charge on each of the oxygens. 

O O e p+ e 

HOC < ► HOC or HOC 

O e O O* 

Since neither of these formulations is convenient to write, we usually indicate 
bicarbonate ion with the ambiguous formula H— O — C0 2 e or, more simply, 
HC0 3 e . Note, however, that only two of the three oxygen atoms in the ion 
can bear the negative charge. The third is bonded to hydrogen and cannot 
participate in n bonding to carbon without giving an unfavorable dipolar type 
of structure : 



9 / 
HO = C 



sec 6.5 vinyl halides and ethers 139 

The anion of a carboxylic acid as acetic acid CH / | is usually written 



K ) 

\ OH/ 



CH3CO2 to avoid the necessity of indicating the dispersal of the negative 
charge to both oxygen atoms, and the same is done with many inorganic ions 
such as N0 3 e , S0 4 2e , and N 3 e . It is important, however, to be aware that in 
any detailed consideration of the properties of these anions, more than one 
resonance structure must be taken into account. 



6-5 vinyl halides and ethers 



An anion that falls in the same classification as those mentioned in the pre- 
vious section is the allyl anion, which can be seen to be structurally analogous 
to the allyl cation and involves two energetically equivalent resonance struc- 
tures : 

e e 

CH 2 = CH-CH 2 < » CH 2 -CH = CH 2 == [CH 2 -CH-CH 2 ] e 

allyl anion 

For many stable compounds, such as unsaturated ethers and halides, we may 
write electronic structures formally similar to those of the allyl anion except 
that there is no net charge : 

CH 2 = CH— O— R CH 2 = CH — F: 

The problem is whether we should draw resonance structures corresponding 
to 71 bonding between carbon and the attached oxygen or halogen : 

f. ® 

CH 2 = CH — O— R < ► CH 2 — CH=0— R 



CH 2 = CH-F: <- 

This is a matter of controversy. Such structures lead to an increase of % 
bonding but with an associated separation of charge wherein electron-attract- 
ing nuclei such as oxygen and fluorine become positive, and weakly electron- 
attracting atoms become negative. There may be some degree of stabilization 
connected with resonance of this type but it is small at best. However, what is 
more certain is that attack of a positive reagent such as Br® on CH 2 =CH— 
OR will occur at the 2 carbon, not at the 1 carbon, because more v. bonding is 
possible in the resulting cationic intermediate : 

Br-CH 2 — CH — 6— R < > Br — CH 2 — CH=0— R 

CH 2 = CH — OR + Br® n bonding 

CH 2 -CH — 6— R 

I 
Br 



chap 6 bonding in conjugated unsaturated systems 140 

Generally speaking, we should expect that importance of resonance of the 

type CH 2 =CH— Z<-+CH 2 — CH=Z will increase as Z becomes less electro- 
negative and will, therefore, be much more important in CH 2 =CH— 

N(CH 3 ) 2 than in CH 2 =CH— F:, and all evidence indicates that this is indeed 

the case. 



6-6 rules Jor the resonance method 

The principles outlined in the previous sections can be reduced to a fairly 
precise prescription which is generally applicable for evaluating the properties 
of conjugated systems : 

1. All resonance structures must have identical locations of the atoms in 
space and the same number of paired electrons. Ionic structures should be 
considered if atoms are present that carry unshared pairs of electrons or 
unfilled octets, or are substantially different in electronegativity. 

2. Relative energies of the various structures are estimated by considering: 
(a) bond energies; (b) the degree of distortion from the geometrically favorable 
atomic positions (if the geometry of the actual molecule is known, then the 
spatial arrangements of all the resonance structures can be taken to conform 
with it) ; (c) the maximum number of valence electrons which each atom can 
accommodate in its outer shell — two for hydrogen and eight for first-row 
elements (this is particularly important). 

3. Electron stabilization is expected to be greatest when structures of 
lowest energy are equivalent. 

4. If there is only a single contributing structure of low energy, then, to a 
first approximation, the resonance hybrid may be expected to have properties 
like those predicted for the particular structure. 



6-7 molecular orbital method of H'uckel 

We have seen that the resonance method is a very useful way of describing 
the behavior of delocalized systems (in which an electron may interact with 
three or more nuclei) in terms of localized structures (in which each electron 
interacts with only two nuclei). The success of the resonance method in 
describing the chemistry of benzene, conjugated dienes, and so on has been 
indicated in earlier sections of this chapter. For benzene and the intermediate 
cation formed in the bromine-butadiene reaction, it was shown that bonding 
involving delocalized electrons could be conceived as arising from electron 
pairing in three or more p orbitals on adjacent atoms overlapping in the n 
manner. These representations were designed to try to overcome the principal 
esthetic defect of the resonance method in that it describes conjugated mole- 
cules or ions in terms of two or more structures. 

It should be recognized that the resonance method is by no means a unique 
approach to rationalizing the properties of conjugated molecules, and we now 



sec 6.7 molecular orbital method of Hiickel 141 

will consider briefly an alternative procedure, molecular orbital (MO) 
theory. Although this theory has many advantages it is less useful in purely 
qualitative structure analysis. MO theory was introduced in 1933 by the 
German chemist E. Hiickel as a method to calculate mathematically the 
relative energies of unsaturated molecules. In the original Hiickel treatment, 
the a bonds were ignored and only the energies of the n electrons considered. 
To do this, ethene and 1,3-butadiene were considered to be made up of 
CH 2 — CH 2 and CH 2 — CH— CH— CH 2 frameworks with singly occupied 
parallel p orbitals on each carbon atom, 



'(f 2 c 



Combination of these p atomic orbitals with due regard for sign (positive or 
negative) gives rise to two molecular orbitals for ethene and four molecular 
orbitals for the diene. (The number of molecular orbitals is always equal to 
the number of atomic orbitals in the combinations.) The energy of an electron 
in each of the molecular orbitals can be calculated, giving us a set of energy 
levels for the n electrons. 

The Hiickel MO method can be applied to a wide variety of unsaturated 
systems. 5 It permits rather simple (though lengthy) calculations to be made 
of the re-electron energies of localized and delocalized conjugated systems 
such as benzene. Furthermore, it allows us to see which energy levels con- 
tribute to bonding and which do not. If an electron in a molecular orbital has 
an energy higher than that of an electron in an isolated p orbital, the orbital 
will, in fact, be antibonding. We will see later that antibonding orbitals are 
especially important in formulating the excited states resulting from ab- 
sorption of visible or ultraviolet light (Section 26T). 

Huckel-type calculations of molecular orbital energies for the % electrons in 
ethene and in 1,3-butadiene reveal that they are arranged as in Figure 6-5. 
The "nonbonding level" is the energy of an electron in an isolated p orbital. 
In both these molecules, the low-lying levels are fully occupied with electrons 
having paired spins, as indicated by the symbol \\. 

A disadvantage of using Hiickel MO theory in an introductory study of 
organic chemistry is that both mathematical and artistic skills are required. 
Furthermore, while drawings in which p orbitals are shown combining toge- 
ther to form n molecular orbitals are useful in many cases (we have already 
used them for illustration earlier in the chapter), in others they can lead to 
false conclusions. The following example illustrates this point. 

Consider the 71-electron systems of 1,3-butadiene and trimethylenemethyl, 
(CH 2 ) 3 C, each of which is C 4 H 6 (Figure 6-6). If we arrange each system to 
give four molecular orbitals into which the four unsaturation electrons are 
placed, and draw dotted lines between the adjacent p-orbitals, we can see no 

5 J. D. Roberts, Notes on Molecular Orbital Calculations, Benjamin, New York, 1961. 



chap 6 bonding in conjugated unsaturated systems 142 











o 






u 


antibonding orbitals 










o 


increasing 








ene 


rgy 


© 


bonding orbitals 


© 
© 




ethane 




1,3-butadiene 



Figure 6-5 Schematic representation of MO energies of ethene and 1,3- 
butadiene. 



reason for any pronounced difference between the two possible arrangements. 
But inspection of possible electron-pairing schemes (as in the resonance 
method) leads immediately to the conclusion that butadiene is likely to be 
very different from trimethylene methyl. The latter will be a diradical if there 
is to be an average of one electron per carbon as can be seen from the possible 
forms that contribute to the resonance hybrid : 



CH, 



CH, = C 



\ 
CH, 



•CH, 



,CH 2 



CH, 



■CH, 



CH, 



CH, 



It should be emphasized that the detailed Hiickel MO calculation of energy 
levels leads to the same conclusion. The four energy levels (there are four p 
orbitals with which to construct molecular orbitals) are arranged as in Figure 
6-7. The two electrons at the nonbonding energy level occupy different orbitals 
and have unpaired spins because this arrangement minimizes the repulsion 
between them (Hund's rule). Trimethylenemethyl has been characterized 



Figure 6-6 Atomic orbital models of 1,3-butadiene and trimethylenemethyl. 



CH, 



CH 2 CH CH CH, 



CH, C *• 




\ CH, 



sec 6.7 molecular orbital method of Huckel 143 



increasing energy — 



o 

© 



antibonding 



- nonbonding 



bonding 



Figure 6-7 Molecular orbitals (energy levels) for trimethylenemethyl, 
(CH 2 ) 3 C. 



as a reaction intermediate and found to have the diradical character expected 
on the basis of the above treatments. 

The greatest triumph of the Huckel MO method is in its treatment of 
conjugated cyclic systems. We can show diagrammatically the result of a 
calculation of the energy levels in any regular, planar, cyclic conjugated system 
made up of trigonally bonded carbon atoms such as benzene by simply 
inscribing a polygon of the appropriate shape within a circle such that one 
apex of the polygon is directly downward. The center of the circle then 
represents the nonbonding level and the apexes represent molecular orbital 
energy levels (Figure 6-8). Note that in the odd-numbered cases, such as the 
three-, five-, and seven-membered rings, there is a trivalent carbon atom. If 
you consider the p orbital on this atom, neglecting for the moment any 
interaction with the neighboring n orbitals, you can see that it may contain 
0, 1 , or 2 electrons. If the orbital is empty, the species will be a carbonium 
ion; if it has one electron, the species will be a neutral radical; if it has two 
electrons, the species will be a carbanion. Using the five-membered ring as an 
example, the three entities are 



Figure 6-8 Energy levels of n molecular orbitals in various conjugated cyclic 
systems. 




3-membered 
ring 



V 

I 

H 



4-membered 
ring 



c— C 



5-membered 
ring 



6-membered 
ring 



,c-c s 



c- 
// 
-c 



w 






^r 



c^-% 



7-membered 
ring 



H 



H 



-H 



c=c x 

H-c C 

\\ // 

r C 

i 

H 



chap 6 bonding in conjugated unsaturated systems 144 



H H 






H H 



carbonium ion 



H 

rati ical 




'The resonance method, discussed earlier, suggests that all of these species 
should be substantially stabilized. Thus, we can write five equivalent structures 
for the carbonium ion. 



H H 



H H 



H 



H 



hA^h ~ hAAh 




I 



H H 



H 



H H 



H^C7-H < ' H-O-H 




There are five similarly equivalent resonance structures for the radical and 
for the carbanion and, from what has been said so far, there would seem to be 
little to choose between them as far as stability is concerned. In the same way, 
we should expect the radical and the ions of the seven-membered ring system 
to be highly stabilized. Again, we show the resonance forms of the carbonium 
ion only, although analogous structures can be written for the radical and 
carbanion : 









H 




























/ \ 
H H 








y 








H 




H 




H 


»w» 




H A H ■ 


H 


tf- 




H A/ H 


H~ 


/ \ 
H H 




H H 




/ \ 
H H 



summary 145 



____2_2_Q__o_ 

© © © © © © 
© © © 



nonbonding level 



C„H 6 C 7 H 7 



Figure 6-9 Filling of the orbitals for five-, six-, and seven-membered rings. 

It turns out that there is a considerable variation in the stabilities of these 
six systems — the cation, radical, and anion forms of the five- and seven- 
membered rings. The anion of the five-membered ring and the cation of the 
seven-membered ring are by far the most stable and we shall encounter these 
later (Section 20-7). It is sufficient to point out at this stage that the Huckel 
MO calculations correctly predicted (in advance) that these in fact were 
the more stable entities. The energy levels shown in Figure 6-8 might appear 
to have been obtained empirically but are, in fact, the consequence of the 
calculations. It will be seen that there are three bonding orbitals (orbitals of 
lower energy than a nonbonding arrangement in which the p orbitals do not 
interact at all) for the five-, six-, and seven-membered rings. The six-membered 
ring with six % electrons filling the three bonding orbitals is the very stable 
compound benzene. The five- and seven-membered rings can also accommo- 
date six % electrons by filling their bonding orbitals. The resulting electronic 
configurations give the anion of the five-membered ring and the cation of the 
seven-membered ring. Because the bonding in these species is maximized, 
they are expected to be particularly stable. The filling of the orbitals is shown 
in Figure 6-9. 

A similar analysis of rings of other sizes shows that the number of electrons 
which suffice to fill the bonding molecular orbitals is given by the formula 
An + 2, where n is an integer or zero. Thus, we expect planar cyclic systems 
containing 2, 6, 10, 14, . . . % electrons to be particularly stable. This has be- 
come known as the "Huckel An + 2 rule" and can be stated thus: "Regular 
monocyclic planar systems made up of trigonal carbon atoms will be most 
stable when they contain {An + 2) n electrons." 



summary 

Compounds with conjugated double bonds are normally more stable than 
their unconjugated isomers, and often undergo conjugate addition reactions. 
Benzene, C 6 H 6 , is an extreme example of stabilization by conjugation. The 
heat of combustion of benzene is more than 30 kcal per mole less than that 
calculated for a compound containing three double bonds, and it does not 
add bromine. The unusual degree of binding in a benzene molecule can be 
expressed in terms of two structural formulas, each with localized bonding 



chap 6 bonding in conjugated unsaturated systems 146 

(electrons interacting with only two nuclei). These correspond to the two 
important ways of pairing the electrons in the p orbitals perpendicular to the 



ring. Benzene is well represented as a resonance hybrid of these structures with 
delocalized bonding (electrons interacting with more than two nuclei). 

1,4-Addition to 1 ,3-butadiene occurs by way of a carbonium ion intermedi- 
ate which can be formulated as a resonance hybrid having the positive charge 
divided between two different carbon atoms : 

CH 2 = CH-CH=CH 2 + Br 2 > Br e + [CH 2 — CH=CH — CH 2 -Br 



< ► CH 2 =CH — CH— CH 2 — Br] 

The n bonding in this ion can be pictured as involving pairing of the electrons 
in three overlapping p orbitals — two from the double bond and one from the 
adjacent atom: 



CH,Br 




1,3-Butadiene, CH 2 =CH— CH=CH 2 , is only slightly stabilized by 
resonance because there is but one favorable low-energy structure. Other 
possible structures are less important because they correspond to much less 
efficient binding. In writing resonance structures, all those to be considered for 
a specific molecule must have the same number of unpaired electrons. Thus, for 
1,3-butadiene, we do not include structures such as • CH 2 — CH=CH— CH 2 • , 
where this structure is meant to be the one with the 1,4 electrons unpaired. 
Such a structure represents a different chemical entity from butadiene. 

Cations such as the allyl carbonium ion, the BrCH 2 derivative of which is 
the intermediate in the 1,4 addition of bromine to butadiene, and anions such 

o o e 

CH 2 =CH-CH 2 < > CH 2 -CH=CH 2 H — O-C < > H-O— C 

O e O 

allyl carbonium ion bicarbonate ion 

as the bicarbonate ion, are stabilized by resonance which, in these cases, 
leads to dispersal of charge. 

Vinyl halides and others have minor contributions to their hybrid structures 
from ionic electron-pairing schemes : 



exercises 147 

e 9 

CH 2 = CH— X < ► CH 2 -CH = X 

CH 2 = CH-0-R < ► CH 2 -CH=0-R 

The most important of the rules for application of the resonance method 
(Section 6-6) are that all of the resonance structures must have the same 
locations of the atoms and the same number of paired electrons. 

Although the resonance method generally gives very satisfactory qualitative 
predictions of the behavior of delocalized systems, the Hiickel molecular 
orbital (MO) procedure for calculating molecular orbital energy levels has 
found many important applications, particularly in cyclic systems. 

In the MO procedure, the p orbitals on each carbon atom in an unsaturated 
system are combined to give molecular orbitals for the n electrons. The molec- 
ular orbitals with energies less than the nonbonding level (which corresponds 
to isolated noninteracting p orbitals) are bonding orbitals, and those of higher 
energy are antibonding orbitals. In regular, planar, cyclic systems made up 
of trigonal carbon atoms, the number of n electrons which exactly fill the 
bonding orbitals is given by the expression An + 2 (where n is an integer or 
zero). 



exercises 

6-1 Discuss the meaning and merit of a statement such as "compound X res- 
onates between forms A and B " in terms of a specific example. 

6-2 Evaluate the relative stabilities of actual molecules of butadiene in each of 
the three forms shown below from the standpoint of steric effects and the 
resonance method. Give your reasoning. 




6-3 Evaluate the importance of resonance of the following types (it may be helpful 
to use ball-and-stick models) : 

H H 

H-_/V_H H^As_.H 

* — • iT < — - xX — 




b. 



H 

HC^ CH 

I II 



^ 



H H 



HCT CH 

HC^ CH 
% CH 



HC 

III 
HC 



H H 
HC 



HC 



# 



CH 
CH 



etc. 



chap 6 bonding in conjugated unsaturated systems 148 



c CH 2 = CH-O e < > CH 2 -CH = C) 

HC=CH HC CH 

A / \ \ 

d. HC X® < ► HC ^C 

^CH 2 7 ^CH 2 7 

H 2 C CH 2 H 2 C "CH 2 

CH ®CH 

e. // \ © < > / \ 

H 2 C CH 2 



H 2 C. 

I 

I 2 C n 2 v 



/ |>-CH 2 < ► | CH=CH 2 

H 2 C H 2 C 



6-4 Do you expect conjugate addition of bromine to occur with 

a. 1,3-cyclohexadiene c. 1,5-hexadiene 

b. l-penten-3-yne d. vinylbenzene 

6-5 Write the structure of the intermediate cation formed by the addition of 
Br® to each of the compounds in Exercise 6-4. 

6-6 Write resonance structures for the ions N 3 e and N0 2 e . 

6-7 Consider the importance of the following resonance for cyclobutene (give your 
reasoning in detail) : 

H 2 C CH 

> II III 



H 2 C-CH H 2 C CH 

Explain what differences in geometry one would expect for a hybrid of such 
structures as compared with conventional cyclobutene geometry. 

6-8 Write three structures for C 4 H 2 with tetravalent carbon and univalent 
hydrogen. Decide on the most favorable geometrical configuration and 
evaluate the resonance energy for this configuration. 

6-9 Propene reacts with chlorine at 300° to yield allyl chloride (3-chloropropene). 

a. Write two chain mechanisms for this chlorination, one involving attack 
of a chlorine atom on the hydrogen of the — CH 3 group and the other 
an attack of a chlorine atom at the double bond. 

b. Consider the relative feasibilities of the two different mechanisms on 
the basis either of bond energies or the stabilities of intermediate 
species. 

6-10 Use the heats of combustion of cyclooctatetraene (1095 kcal/mole) and of 
cyclooctane (1269 kcal/mole), and any required bond energies (Table 2-1), 
to calculate the heat of hydrogenation of cyclooctatetraene to cyclooctane in 
the vapor phase at 25°. 

6-11 Write the five Kekule-type resonance structures of phenanthrene, and show 
how these can account for the fact that phenanthrene, unlike benzene, adds 
bromine, but only across the 9,10 positions. 



exercises 149 




^ phenanthrene 



6-12 The compound trichlorocyclopropenium tetrachloroaluminate, C 3 C1 3 ®, 
AlCl 4 e exists as a stable salt at room temperature. Use the Huckel rule to 
account for the unusual stability of this carbonium salt. 

6-13 An ionic derivative of cyclooctatetraene, C 8 H 8 , has recently been prepared. 
On the basis of the Huckel rule, which of the following formulas do you 
expect it to have: C 8 H 8 ®, C s H 8 e , or C 8 H 8 2e ? 

6-14 In 1825, a yellow crystalline compound, named croconic acid, was prepared 
by the action of hot potassium hydroxide on carbon powder followed by 
acidification. This compound, whose formula was later shown to be H 2 C 5 5 
and to contain a five-membered carbon ring, is quite a strong acid and gives 
the ion C 5 5 2e on ionization. Suggest a structure for this acid and the ion 
and show how resonance stabilization of the ion accounts for the acid being 
strong. 



0; IB tl ft: "' ^ V -tW^ ' : ^&ipl^ij^ i 

isolation an<J identification 
•"■'■'? of organic compounds 



sec 7.1 isolation and purification 1S3 

A characteristic of chemistry that is not always shared with the other sciences 
is a concern for purity (in the sense of molecular homogeneity) in the materials 
under study. Much of the time of some organic chemists working in the lab- 
oratory is devoted to isolating compounds in pure form, and the culmination 
of a synthetic scheme is always the purification and confirmation of the iden- 
tity of the product. In this chapter we consider, first, some modern means of 
separating and purifying organic compounds and, second, instrumental 
methods which can be used to elucidate structures. There have been enormous 
advances in these regards over the past 20 years through introduction of new 
techniques. Structural or purification problems which once may have required 
years for solution, if they could be solved at all, can often now be solved in 
hours or days, but very frequently with complex and costly apparatus. 



7-1 isolation and purification 



Fractional distillation and crystallization are the time-honored methods for 
separating and purifying liquids and solids, respectively. Obviously, knowledge 
of the relationship between structure and boiling point will be helpful in 
conducting a distillation. Crystallization of solid compounds calls more for 
an understanding of solubilities as a function of temperature in the common 
solvents such as water, ethyl alcohol, benzene, and tetrachloromethane 
(carbon tetrachloride). The most suitable solvent for crystallizing or recrystal- 
lizing an impure compound is usually one in which the compound is slightly 
soluble at room temperature, for if the solubility is too high the recovery will 
be poor because too much of the compound will be left in solution. On the 
other hand, if the solubility at room temperature is too low, you can expect 
difficulty in trying to dissolve the compound completely in the solvent at 
higher temperatures. 

Liquid-liquid extraction techniques, using a separatory funnel, can some- 
times provide clean separations, particularly with compounds possessing 
a basic group such as amino ( — NH 2 ) or an acidic group such as carboxyl 




The most widely used technique for separating and purifying compounds 
on a small scale is chromatography. Chromatography can be defined as the 
separation of components of a mixture by differences in the way they be- 
come distributed (partitioned) between two phases. By this definition, chroma- 
tography includes the simple techniques described above, such as extraction in 
a separatory funnel (a liquid-liquid two-phase system), fractional distillation 
(a gas-liquid two-phase system), and crystallization (a solid-liquid two-phase 
system). These techniques require fairly large amounts of material and they 
may give unsatisfactory separations, which may be attributed to the small 
degree of separation possible in any given one-stage, or even several-stage, 
partitioning process. We now have available what might be called super- 
separation chromatographic methods, which involve multistage partitioning of 



chap 7 isolation and identification of organic compounds 154 

very small amounts of sample, a few milligrams or less, wherein extraordin- 
ary separations can often be achieved. To this end, the most frequently 
employed combinations of phases are gas-liquid and liquid-solid. 

Liquid-solid chromatography was originally developed for the separation 
of colored substances, hence the name chromatography (which stems from 
the Greek word chroma meaning color). In a typical examination, a colored 
substance suspected of containing colored impurities is dissolved in a suitable 
solvent and the solution allowed to pass through a column packed with some 
solid adsorbent (e.g., alumina), as shown in Figure 7-1. The " chromatogram " 
is then "developed" by adding a suitable solvent that washes the adsorbate 
down through the column. Hopefully, and this is the crux of the entire 
separation, the components are adsorbed unequally by the solid phase, and 
distinct bands or zones of color appear. The bands at the top of the column 
contain the most strongly adsorbed components; the bands at the bottom 
contain the least strongly held components. The zones may be separated me- 
chanically, or solvent can be added to wash, or elute, the zones separately from 
the column for further analysis. If all attempts to resolve a given substance 
chromatographically are unsuccessful, evidence (albeit negative) is thus pro- 
vided for the presence of a single pure chemical entity. Although the various 
zones can be observed by eye only if the substances are colored, a variety of 
other techniques can be used to detect the bands on the chromatogram and 
the method is by no means limited to colored substances. 



Figure 7-1 A simple chromatographic column for liquid-solid chromatog- 
raphy. 



im 



. solvent 



lUSlilMil I s °lid adsorbent with 
> bands of separated 
sample components 




eluate 



sec 7.1 isolation and purification 1SS 





pressure -regulated exit vapors i— detector , — packed column 
carrier gas supply i / / 

in/ 1 








( \ l_ll_J )) 

■ i oven 






\— sample injection port 





Figure 7-2 Schematic diagram of a vapor-phase chromatography apparatus. 
The detector is arranged to measure the difference in some property of the 
carrier gas alone versus the carrier gas plus effluent sample at the exit. Differ- 
ences in thermal conductivity are particularly easy to measure and give high 
detection sensitivities. 



More recently, gas-liquid phase chromatography, glpc (also called vapor- 
phase chromatography, vpc, or gas chromatography, gc) has added a new 
dimension for the analysis of volatile substances. In the usual form of vpc, 
a few microliters of a liquid to be analyzed are injected into a vaporizer 
and carried with a stream of gas (most frequently helium) into a long heated 
column, packed with some porous solid (such as crushed firebrick) impreg- 
nated with a nonvolatile liquid or oil. Gas-liquid partitioning occurs, and small 
differences between partitioning of the components can be magnified by the 
large number of repetitive partitions possible in a long column. Detection is 
usually achieved by measuring changes in thermal conductivity of the effluent 
gases. A schematic diagram of the apparatus and a typical separation are 
shown in Figures 7-2 and 7-3. While observation of a single peak in the vpc 



Figure 7-3 A vapor-phase chromatogram of a mixture of isomeric butyl 
alcohols, C 4 Hc,OH. 



; 


i 


CH, 


| 


CH,CH, 

\ 






CHj— G— « 


til ' 


^^^^^ft 


dete 


ctor 


CH, 


[. 




signal 




J 


^^^^^M 




M 










time 



chap 7 isolation and identification of organic compounds 1 56 

analysis of a material is, of course, only negative evidence for purity, the 
method is extraordinarily useful for detection of minute amounts of impuri- 
ties when these are separated from the main peak. The sensitivity of vpc 
makes it the method of choice for analysis of volatile pesticide residues in 
food products, automobile exhausts, smokestack effluents, trace components 
of alcoholic beverages, flavors, perfumes, and so on. 

Vapor-phase chromatography can be used effectively to purify materials 
as well as to detect impurities. To do this, the sample size and the size of the 
apparatus are generally increased, and the vapor of the pure component is 
condensed as it emerges from the column. 

Two other chromatographic methods, paper chromatography and thin- 
layer chromatography (tic), are described in Chapter 17. 



7-2 identification of organic compounds 

Distillation, crystallization, or chromatography often tells us something of the 
state of purity of a compound. A single chromatographic peak, a sharp boiling 
point, or a sharp melting point of a crystallized solid is an indication that a 
pure compound has been obtained. (Although melting and boiling points 
will be listed in this and other books as single temperatures, in practice one 
encounters a melting or boiling range of usually one degree or so.) 

If the substance being purified is believed to be a previously known com- 
pound, it is usually a simple matter to compare its physical properties (melting 
point, boiling point, refractive index) or spectroscopic properties with those 
reported in the chemical literature for the known compound. However, if 
the structure of the compound is not known, you must adopt a different 
approach. 

The traditional method, and indeed the only one really available until the 
second half of this century, is based on the molecular formula obtained from 
the molecular weight and the elemental analysis, in combination with 
qualitative chemical tests and degradative reactions which ultimately would 
lead to known compounds. The process was to a large extent highly intuitive 
but the conclusions drawn about structure usually proved correct. Although 
instrumental methods of analysis have revolutionized structure determina- 
tion in the past 25 years, it would be a mistake to conclude that a wide knowl- 
edge of the transformations of organic compounds that are possible is no 
longer necessary for the successful elucidation of molecular structures. 

The first step in the process of determining a structure is a quantitative 
analysis for the elements it contains. 



A. ELEMENTAL ANALYSIS 

The vast majority of organic substances are compounds of carbon with hydro- 
gen, oxygen, nitrogen, or the halogens. All of these elements can be determined 
directly by routine procedures. Carbon and hydrogen in combustible com- 
pounds can be measured by combustion of a weighed sample in a stream of 



sec 7.2 identification of organic compounds 157 



sample in 
platinum boal> 




furnace 



m^- 



-CuO pellets 




B excess 



MgC10 4 



soda-lime 



2^ 



Figure 7-4 Schematic representation of a combustion train for determination 
of carbon and hydrogen in combustible substances. 



oxygen (Figure 7-4) followed by gravimetric determination of the resulting 
water and carbon dioxide through absorption in anhydrous magnesium 
perchlorate and soda lime, respectively. Routine carbon-hydrogen analyses 
by combustion procedures utilize 3- to 5-mg samples and attain an accuracy 
of 0.1 to 0.2%. Complete elemental analyses permit computation of empirical 
formulas which are equal to, or are submultiples of, the molecular formulas. 
A knowledge of the molecular weight is needed to determine which of the 
multiples of the empirical formula corresponds to the molecular formula. For 
example, the empirical formula of the important sugar, glucose, is CH 2 0, 
but the molecular weight defines the molecular formula as (CH 2 0) 6 or 
C 6 H 12 6 . 

Molecular weights of gaseous organic compounds may be determined 
directly by vapor-density experiments. Liquid substances of moderate vola- 
tility may be vaporized, and their vapor densities determined. However, 
molecular weights of volatile compounds are not often determined in present- 
day organic research, because the molecular weight of a given substance 
may usually be estimated from the boiling point, to perhaps 25 % accuracy, 
by knowledge of the boiling points of related compounds. Molecular weights 
of high-boiling liquids and slightly volatile solids are generally determined 
by measuring freezing-point depressions or boiling-point elevations of solu- 
tions in suitable solvents. Solutions in a substance such as camphor, which 
gives a very large freezing-point depression (37.7° per mole of solute dissolved 
in 1000 g of camphor), are particularly suitable for small-scale operations and 
can, with ordinary equipment, permit reasonably accurate molecular-weight 
measurements by determination of melting points. 

There are many other more or less specialized techniques available, the 
choice depending on the problem at hand and accessibility of equipment. 
Thus, molecular weights of even slightly volatile substances may often be 
obtained by mass spectrometry, which is the method of choice, particularly 
since resolution is routinely achieved for masses as high as 600. The molec- 
ular weights of very high-molecular-weight compounds such as proteins and 
polymeric materials are frequently determined by end-group analysis, 
measurement of osmotic pressure, viscosity, light scattering, and sedimenta- 
tion. Some of these methods will be discussed in more detail in later chapters. 



chap 7 isolation and identification of organic compounds 158 
B. MASS SPECTROMETRY 

The application of mass spectrometry to organic molecules involves bom- 
bardment of a vaporized sample with a beam of medium-energy electrons in 
high vacuum and analysis of the charged particles and fragments produced. 

The elements of a mass spectrometer are shown in Figure 7-5. The positive 
ions produced by electron bombardment are accelerated by the negatively 
charged accelerating plates and are swept down to the curve of the analyzer 
tube where they are sorted as to their mass-to-charge (m/e) ratio by the 
analyzing magnet. With good resolution, only the ions of a single mass num- 
ber will pass through the slit and impinge on the collector, even when the 
mass numbers are in the neighborhood of several hundred or a thousand. 
The populations of the whole range of mass numbers of interest can be deter- 
mined by plotting the rate of ion collection as a function of the magnetic 
field of the analyzing magnet. 

The intense peak that is highest in mass number is of considerable impor- 
tance ; this corresponds to the positive ion formed from the parent molecule 
by the loss of just one electron and provides a highly accurate method for 
measuring molecular weights. 

In recent years, considerable success has been achieved in the correlation 
of the relative abundances of various-sized fragmentation products with the 
molecular structures of the parent molecules. This use of mass spectrometry 
is discussed in more detail in Chapter 29. 



Figure 7-5 Schematic diagram of a mass spectrometer. 



• electron gun 



accelerating plates (negative potential) 



sample 




ion collector- 



sec 7.3 absorption of electromagnetic radiation 159 

Super-resolution mass spectrometers are now available which permit 
distinction between ions of such molecules as C 43 H 50 N 4 O 6 (mol. wt. 718.373) 
and C 42 H 46 N 4 7 (mol. wt. 718.337). Such mass spectrometers can be used 
for elemental analysis. 



spectroscopy 

Virtually all parts of the spectrum of electromagnetic radiation, from X rays 
to radio waves, have found some practical application for the study of 
organic molecules. Visible light is an extremely narrow band in the spectrum 
characterized by the wavelength range 4000 to 7500 A; it occupies a position 
between ultraviolet and infrared radiation but, in terms of the way it interacts 
with organic molecules, it can be considered part of the ultraviolet region. 

The absorption of electromagnetic radiation by molecules results in an 
increase in their energy. Very short-wavelength radiation, such as X rays or 
gamma rays, supplies so much energy that it causes covalent bonds to rupture 
and thus leads to drastic chemical changes. Long-wavelength radiation such 
as radio waves, on the other hand, may be able to induce only more rapid 
rotation of a molecule. 

Most of the spectroscopic methods in use by organic chemists involve 
measuring the amount or wavelengths, or both, of radiant energy absorbed 
by molecules, and the principles and practice of some of these methods will 
be discussed. First, however, mention should be made of X-ray diffraction, 
because this technique has proved to be of special value in elucidating 
structures of organic molecules. 

When a beam of X rays strikes a crystal, the result is a diffraction pattern 
which is characteristic of the locations of the atoms in the crystal. The diffrac- 
tion pattern can usually be analyzed with the aid of a high-speed digital com- 
puter and the molecular structure may often be determined in a matter of 
days. (Diffraction does not itself involve absorption of radiation, although 
the crystal usually suffers some damage because of bond rupture caused by 
absorption of X rays.) Not all compounds can be examined in this way, 
because the compound must be crystalline and, preferably, contain at least 
one " heavy " atom, such as chlorine or bromine. 



7-3 absorption of electromagnetic radiation 

The energy, e, of a photon is related to its frequency of vibration, v, by Planck's 
constant, h: 

e = hx 

In general, you can say that a molecule will absorb incident radiation only if 
there is some higher energy state (with energy E 2 ) to which the molecule can 
be raised from its normal or ground state (with energy E t ) by the energy of 



chap 7 isolation and identification of organic compounds 160 

the photon. Because not all photons have the proper energy to take a molecule 
from the normal to a higher energy state, we say that the energy levels are 
quantized, and that a specific quantum of energy separates the two states : 



Energy AE l2 



! E, 

The difference in energy AE 12 (=E 2 — £"i) is related to the frequency (v sec -1 ) 
or wavelength (A cm) of the absorbed radiation by the equations 

he 

AE 12 = hv = — 

where h is Planck's constant and c is the velocity of light. We generally will 
be interested in the energy change in kilocalories (1 kcal = 10 3 cal) which 
would result from one mole of substance absorbing light and we can rewrite 
the above equation in the form 

286,000 

A£l2 = ^AT kcal 

where X is now in angstrom (A) units (1 A = 10~ 8 cm). As defined here, 
AE 12 corresponds to 1 einstein of radiation. 

The total energy (E) of a molecule (apart from nuclear and kinetic energy) 
can be expressed as the sum of three energy terms : 

-^ -'-'electronic ' -^vibrational ~t~ -^rotational 

The electronic energy levels correspond to the energies of the various molecular 
orbitals (Section 6-7) and are rather widely spaced. Ultraviolet light (and 
sometimes visible light) has sufficient energy to cause transitions between 
the electronic energy levels. Vibrational energies are also quantized but the 
vibrational energy levels are more closely spaced and the differences between 
them correspond to photons of infrared radiation. Finally, the energies of 
rotational states for a molecule or its parts are very closely spaced and ro- 
tational transitions result from absorption of microwaves (radar) or radio 
waves. 

A molecule that has absorbed a photon of ultraviolet light, for example, 
normally remains in the resulting excited electronic state for only a very 
brief period. The energy may be reemitted or it may be shuffled into vibration- 
al and rotational energy with a resulting increase in thermal energy of the 
system. In some cases, the excited molecules undergo chemical reactions and 
do not return to the original ground state. The fading of dyes is an example 
of this kind of behavior. The types of transition associated with absorption 
of radiation from the various regions of the electromagnetic spectrum are 
shown in Figure 7-6. 

Although microwave spectroscopy can provide valuable information about 
bond lengths and bond angles in simple molecules, it has not so far been of 



sec 7.4 infrared spectroscopy 161 



wavelength, X 1CT 10 1CT 6 10~ 2 10 2 10 6 

i 1 m — n 1 r 



X-RAYS 
7-RAYS 



] 



cm 



UV M IR RADAR RADIO 



type of transition (BOND RUPTURE) VIBRATIONAL 

ELECTRONIC ROTATIONAL 



Figure 7-6 The types of transitions in molecules brought about by absorption 
of radiation from various regions of the electromagnetic spectrum. 



general use for identification or structure proof of organic compounds. In- 
frared and ultraviolet spectroscopy are much more widely and routinely 
useful. The radio wave absorptions when used in conjunction with an applied 
magnetic field are also of great importance, as we shall see in a later section 
of this chapter. 



1-4 infrared spectroscopy 



Possibly the single most widely used tool for investigating organic structures 
and for organic chemical analyses is the infrared spectrometer. Spectra for 
infrared radiation over the wavelength region from 2 to 15 microns (1 micron 
= n = 10~ 4 cm) are of most interest. Recording infrared spectrophotom- 
eters with excellent resolution and reproducibility are commercially 
available and are widely used in organic research. The operating parts of a 
spectrometer are shown in Figure 7-7. The sample containers (cells) and opti- 
cal parts of infrared spectrophotometers are made of rock salt (NaCl) or 
similar materials, since glass is opaque to infrared radiation. Gaseous, liquid, 
or solid samples can be used. Solids are often run as finely ground suspensions 
(mulls) in various kinds of oils, or ground up with potassium bromide and 
compressed by a hydraulic press into wafers. Considerable differences are 
often observed between the spectra of a solid and its solutions. 

In Figures 7-8, 7-9, and 7-10 are shown the infrared spectra of an alkane, 
an alkene, and an alkyne. In accord with current practice, the infrared spectra 
given here are linear in wave numbers (v cm -1 ) that are related to radiation 
frequencies (v sec -1 ) so that v = v/c, where c is the velocity of light in cm/sec. 
A supplementary nonlinear wavelength scale in microns is shown, and to 
convert wave numbers v to A in microns we use the relation 1 = 10 4 /v. 

Considerable confusion exists as to the units used to express wavelength or 
frequency for various kinds of spectra. For electronic spectra, A is most 
commonly expressed in angstrom units (10~ 8 cm) or millimicrons (10~ 7 cm). 
For infrared spectra, absorbed radiation may be defined by its wavelength X 
in microns (10 ~ 4 cm) or by its frequency v in units of wave numbers (10 4 /A 
cm" 1 ). 

The infrared absorption bands between 3600 and 1250 cm -1 can be identi- 
fied in most cases with the changes in the vibration of a particular bond. 



chap 7 isolation and identification of organic compounds 162 



K 



iwi.ink-r 



i infrared detector 
(thermocouple) 




cell for solution 
of sample 



f!L 







infrared source 
(electrically 
heated filament) 



beam 
chopper 




-cell for solvent 



Figure 7-7 Schematic representation of a "double-beam" recording infrared 
spectrophotometer. The beam chopper permits radiation passing alternately 
through sample cell and solvent cell to reach the thermocouple. This 
procedure permits the difference in absorption by solute and solvent to be 
measured as an alternating electric current from the thermocouple. The 
alternating-current output is particularly desirable for electronic amplifica- 
tion. The usual commercial instruments operate on a "null" principle with 
the recorder pen linked mechanically to a "comb" (not shown here), which 
is placed across the solvent-cell beam and moved by a servomechanism to 
reduce or increase the solvent-cell-beam intensity. The servomechanism is 
actuated by the amplified thermocouple output to make the solvent-beam in- 
tensity equal to the solution-beam intensity — that is, to reduce the thermo- 
couple output to zero or the null point. The spectrum can be scanned 
through the various wavelengths by rotation of the prism in synchronization 
with the motion of the recorder drum. The use of a diffraction grating in 
place of the prism is becoming increasingly common. 



Below 1250 cm" 1 is the so-called fingerprint region, which is associated chiefly 
with complex vibrational changes in the molecule as a whole. 

The band at highest frequency in all three hydrocarbons is that which 
corresponds to changes in the C — H stretching energies. The higher the fre- 
quency of the radiation, the higher its energy, and this means that to raise the 
vibrational level of a C — H bond requires more energy than that for any 
other bond to carbon. We can regard the C— H bond to be like a spring be- 
tween the two atoms, with the light hydrogen atom vibrating with respect to 
the much heavier carbon atom. Because vibrational energies are strictly 
quantized, only certain values of the stretching frequency are allowed. 



sec 7.4 infrared spectroscopy 163 




Figure 7-8 The infrared spectrum of octane, CH 3 CH 2 CH 2 CH 2 CH 2 CH 2 CH 2 CH 3 . 



Figure 7-9 The infrared spectrum of 1-butene, CH 3 CH 2 CH=CH 2 . 







wavelength, IX 




3 4 5 5 


6 7 8 9 10 12 14 




1 ^JU-^. 11 1 1 III II 

: II \ i i / %' v\ / I 




^^^■HpiWpl lit 








■ ,\ (C.ll,) ■. A 


.2 


oveiB 
'1 ' 


E 


s^asw|Ji|lipsM^i^»li- |i% 


^ip|%feiillifr€tiiiilllilpps^ 




•■■V f-II 

/ suvu-h 
l--l'll 2 M 


H W beml ' | ! 




(CIU. r 1 1 , i 

^lllP^P^lll sNftPft^illiP 111! 






i i 


/ i i i 1 


1,1.1,1,1.1. 




3600 3200* 2800 2400 2000 2000 


J 8O0 1 600 1 400 1 200 1 000 800 

frequency, cm - 



chap 7 isolation and identification of organic compounds 164 









wavelength, jU 












3 4 


S 


5 6 7 8 

\ W Y*YV J \/ / 
1 " ' " [1. ^ 

~"N. ' * If It . 11 


9 

M 


10 

n V \ 

l| 


12 


14 

II | 
{II i 




I 1 , 


fjf 

L 


G 

.2 

I 

G 


1 




(.' = ( '; 
stretch !| 


V 


1 


1 




J. 


^ 






^^^^^^^^^^^^^^^« ^^^^^^^^^^^ 












h"-^ 


















f . C-H 






I^S/1 








1 




(BC-H S,R ' tdl 
» it i 


i 









i , 




[_ 




3600 3200 2800 2400 


2000 


2000 1800 1600 1400 1200 




1000 


800 








frequency, cm" 











Figure 7-10 The infrared spectrum of phenylacetylene, C 6 H 5 C=CH, in carbon 
tetrachloride solution. The sharp band near 1500 cm -1 is characteristic of a 
vibration of the benzene ring. 



At room temperature, more than 95 % of the C— H bonds of a sample of 
hydrocarbon are vibrating in their lowest (or zero-point) energy level. That is, 
the thermal energy of the system is sufficient to raise only a small fraction of 
the hydrogens to a higher vibrational level than that which would be found at 
absolute zero. The absorption bands near 3000 cm"" 1 , then, correspond to a 
change from the lowest vibrational state to the first excited vibrational state 
of the C— H bond. The slightly different locations of these bands for different 
kinds of C— H bond (compare Figures 7-8, 7-9, and 7-10) are of considerable 
use in structure identification. 

The bending frequencies for C— H bonds occur at lower wave numbers 
showing that a smaller quantum of energy is required for excitation of bending 
vibrations : 

V 
Eq— H > ^C-H • 



The absorption of energy by changing the stretching energies of carbon- 
carbon bonds occurs at lower frequencies than those of carbon-hydrogen 
bonds and are widely separated, depending on whether the atoms are joined 
by a single, double, or triple bond. The carbon-carbon single-bond stretching 
bands occur in the complex "fingerprint" region and are difficult to identify. 
Carbon-carbon double and triple bonds, however, absorb at higher frequen- 
cies and can be readily identified in an infrared spectrum. The analogy of 
bonds to springs can be usefully applied here. The stronger the forces holding 
two atoms together, the greater will be the energy required to raise the system 



sec 7.5 ultraviolet and visible spectroscopy (electronic spectroscopy) 16S 

to the next vibrational level. The total binding energy of a double bond is less 
than that of a triple bond, and this is shown in the infrared spectra by the fact 
that the C=C stretching band in 1-butene appears at 1700 cm -1 and the C=C 
band in phenylacetylene at 2100 cm -1 . 

The most common positions of the easily identified stretching frequencies 
in hydrocarbons are 

C— H C=C C=C 

-3000 cm- 1 -1650 cm- 1 -2200 cm" 1 

Table 7-1 lists the characteristic infrared frequencies for many kinds of com- 
pounds that we shall subsequently meet. 

, The spectrum of 1-butene in Figure 7-9 shows an absorption that can 
trap the unwary. Thus the band at 1830 cm -1 falls in the region where 
stretching vibrations are usually observed. However, this band actually arises 
from an overtone (harmonic) of the =CH 2 out-of-plane bending at 915 cm -1 . 
Such overtone absorptions come at just twice the frequency of the funda- 
mental frequency and whenever an absorption like this is observed, which 
does not seem to fit with the normal fundamental vibrations, the possibility 
of its being an overtone should be checked. 



7-5 ultraviolet and visible spectroscopy 
(electronic spectroscopy) 

The photons of visible light (4000-7500 A) are an order of magnitude 
higher in energy than infrared photons and the photons of ultraviolet light 
are of still higher energy. The type of spectroscopy that embraces these more 
energetic kinds of radiation is called electronic spectroscopy because the 
excitation that results from absorption of such photons normally involves 
the raising of one electron to a higher energy orbital. 

We expect that electronic excitation will occur with lower-energy photons 
in molecules whose ground-state electrons are only loosely bound to their 
nuclei. Such compounds may, in fact, absorb visible light and be colored. 
Molecules in which all the electrons are rather tightly bound — alkanes, for 
instance — can only be excited electronically by the high-energy ultraviolet 
radiation. The valence electrons in alkanes are in C— C and C— H a bonds, 
and raising one of these electrons to a higher energy, called a a -+ a* transi- 
tion, requires radiation of wavelength 1300 A or less. Such radiation is not 
easily accessible for routine analysis because neither air nor quartz (the best 
substance available for constructing cells and prisms) is transparent to such 
radiation. The practical lower limit for routine ultraviolet absorption spectros- 
copy is about 1900 A, and for this reason alkanes are transparent throughout 
the whole of the readily accessible ultraviolet and visible region. 

The experimental arrangement for determining such spectra is usually 
quite similar to that of infrared spectrometers shown in Figure 7-1. The 
principal differences lie in the use of a tungsten lamp (3200 to 8000 A) or 
hydrogen arc (1800 to 4000 A) as the light source; quartz prism and sample 
cells ; and a photoelectric cell, rather than a thermocouple, as the radiation 



chap 7 isolation and identification of organic compounds 166 
Table 7-1 Some characteristic infrared absorption frequencies 



bond 


type of compound 




frequency, 
cm -1 


intensity 


1 
— C-H 

1 


alkanes 




2850-2960 


strong 


1 
— C-D 

1 


alkanes 




-2200 


strong 


1 
=C-H 


alkenes and arenes 




3010-3100 


medium 


=C-H 


alkyras 




3300 


strong, 
sharp 


1 1 

— c-c- 

1 I 


alkanes 




600-1500" 


weak 


w 

r \ 


alkenes 




1620-1680 


variable 


-c=c- 


alkynes 




2100-2260 


variable 


— C=N 


nitriles 




2200-2300 


variable 


1 

— c— o- 

i 


1 1 
alcohols — C— OH, ethers — C— O- 
1 i 


1 

-c— , 

1 






J 


i i 


i 

p 

— c 
\ 
o- 


1000-1300 


strong 




-C — 

I . 





' In general, C — C single-bond stretching frequencies are not useful for identification. 



detector. In these spectrometers, the prism is placed ahead of the sample. 

In simple alkenes and alkynes the n electrons are, of course, less tightly 
bonded than are a electrons, but the excitation energy is still too great for 
significant absorption to occur in the easily accessible region of the ultraviolet 
spectrum. Conjugated dienes, however, absorb strongly at wavelengths above 
2000 A. This kind of absorption results from raising an electron from a normal 
rc-bonding state to a 71-antibonding state and is called a % -* n* transition. 

The more extended the conjugated system becomes, the smaller is the energy 
difference between the normal and excited states. Thus, the diphenylpolyenes, 
C 6 H 5 — (CH=CH)„— C 6 H 5 , absorb light at progressively longer wave- 
lengths as n increases ; this is apparent from the colors of these compounds, 
which range from colorless (» = 1) through yellow and orange (n = 2-7) to 
red (n = 8) as the wavelength increases from the ultraviolet well into the visible 
region of the electromagnetic spectrum. Lycopene, the red pigment in 
tomatoes, is a polyene with 11 conjugated double bonds (Section 29-3). 



sec 7.S ultraviolet and visible spectroscopy 167 
Table 7.1 Some characteristic infrared absorption frequencies (continued) 



bond 


type of compound 




frequency, 
cm -1 


intensity 


\ 

c=o 
/ 


O 

II 
aldehydes — C— H 




1720-1740 


strong 


x c=o 

/ 


O 

1 II 1 
ketones — C— C— C— 
1 1 




1705-1725 


strong 


x c=o 

/ 


O O 

/ // 
acids — C esters — C 

O-H O- 


-c— ' 

1 


1700-1750 


strong 


-O-H 


1 1 
alcohols — C— O— H, phenols =C- 


-O-H 


3590-3650 


variable, 
sharp 


-O-H 


hydrogen-bonded, alcohols and 
phenols — O — H-O 




3200-3400 


strong, 
broad 


-O-H 


hydrogen-bonded, acids — O — H- 


/ 
O 

\ 


2500-3000 


variable, 
broad 


-NH 2 


1 
amines — C — NH 2 




3300-3500 


medium 




1 




(double peak) 


1 


H 

1 1 1 




3300-3500 


medium 


— N— H 


amines — C— N — C— 
1 1 




(single peak) 



Electronic absorption bands are usually much broader than infrared bands 
for two principal reasons. First, at ordinary temperatures, molecules in 
either the ground or excited electronic states exist in a number of vibrational 
or rotational states, and the transitions which occur between the electronic 
states are brought about by quanta of slightly different energy. The result is 
to give an absorption band made up of a large number of lines which are too 
closely spaced to be separately distinguishable. The second reason arises from 
the fact that electronic spectra are usually taken of solutions, and the range 
of solute-solvent interactions gives a spread of energies to both the ground 
and excited electronic states. 

We have seen that conjugated molecules, such as 1,3-butadiene, have normal 
or ground states (Section 6-3) which are slightly more stable than would other- 
wise be expected. Because conjugated molecules also have electronic absorp- 
tion bands toward longer wavelengths than nonconjugated molecules 
(smaller energy differences between ground and excited states), the energies 



chap 7 isolation and identification of organic compounds 168 

of the excited states of such molecules must have a higher degree of resonance 
stabilization than the ground state. 

Can you depict excited electronic states by conventional structural for- 
mulas? Only in a rather unsatisfactory way, unfortunately. You can formulate 

the excited state of 1,3-butadiene associated with absorption of ultraviolet 

© e 

radiation by a dipolar structure such as CH 2 — CH=CH— CH 2 , which we 
have already considered in connection with the ground state of the molecule 
(Section 6-4). Thus, you can consider the ground state as being close to [1] 



CH 2 =CH-CH=CH 2 < ► CH 2 -CH=CH-CH 2 

[1] [2] 



[3] 



with minor contributions from [2], [3], and other energetically unfavorable 
structures. On the other hand, the excited state can be taken to correspond to 
[2] with major contributions from other polar forms such as [3], which can 
contribute in a major way because they should have comparable energies and 
a minor contribution from [I]. The important points are that the excited state 
has less total bonding than the ground state and, because it is expected to be 
more of a hybrid structure than the ground state, it will have a different geom- 
etry, particularly a shorter 2,3 carbon-carbon bond. In general, we will ex- 
pect that the greater the degree of conjugation and the more favorable the 
polar forms appear, the more the excited state will be stabilized and the 
longer will be the wavelength of the maximum absorption, A max . 



7-6 nuclear magnetic resonance spectroscopy 

Nuclear magnetic resonance (nmr) spectroscopy is very useful for identifi- 
cation and analysis of organic compounds. The principles of this form of 
spectroscopy are quite simple. The nuclei of some kinds of atoms act like 
tiny magnets and become lined up when placed in a magnetic field. In nmr 
spectroscopy, we measure the energy required to change the alignment of 
magnetic nuclei in a magnetic field. 

A schematic diagram of a very simple form of an nmr instrument is shown 
in Figure 7-11. When a substance such as ethyl alcohol, CH 3 — CH 2 — OH, 
(the hydrogens of which have nuclei that are magnetic) is placed in the center 
of the coil between the magnet pole faces, and the magnetic field is increased 
gradually, energy is absorbed by the sample at certain field strengths and the 
current flow in the coil is increased. The result is a spectrum such as the one 
shown in Figure 7-12. This spectrum is detailed enough to serve as a most 
useful fingerprint for ethyl alcohol but is also simple enough for the origin of 
each line to be accounted for, as we shall see. 



sec 7.6 nuclear magnetic resonance spectroscopy 169 











A 
















f 


1 






W/// 


m 








radiofrequency 
oscillator 






i 


i 


o\ 




W/ 


w 


AA 

sensitive 
ammeter 












magns 




ngth/7 








-tic field of stre 







Figure 7-11 Essential features of a simple nmr spectrometer. 



For what kinds of substances can we expect nuclear magnetic resonance 
absorption to occur? Magnetic properties are always found with nuclei of 
odd-numbered masses, *H, 13 C, 15 N, 17 0, 19 F, 31 P, and so on, and nuclei of 
even mass but odd number, 2 H, 10 B, 14 N, and so on. Nuclei such as 12 C, 
16 0, 32 S, and so on, with even mass and atomic numbers have no magnetic 
properties and do not give nuclear magnetic resonance signals. For various 
reasons, routine use of nmr spectra in organic chemistry is confined to 1 H, 



Figure 7-12 Nuclear magnetic resonance spectrum of ethyl alcohol (contain- 
ing a trace of hydrochloric acid). Chemical shifts are relative to tetramethyl- 
silane, (CH 3 ) 4 Si or TMS = 0-Q0 ppm. The stepped line is an integral of the 
areas under each of the resonance lines. 




chap 7 isolation and identification of organic compounds 170 

19 F, and 31 P. We shall be concerned here principally with nmr spectra of 
hydrogen ('H), often called pmr spectroscopy (proton magnetic resonance 
spectroscopy). 

Nuclear magnetic resonance spectra may be so simple as to have only a 
single absorption peak but can also be much more complex than the spectrum 
of Figure 7-12. On the one hand, complexity is helpful because it makes the 
spectra more individualistic and better suited as fingerprints for characteri- 
zation of organic molecules. However, complexity can hinder the use of nmr 
spectra for qualitative analysis and structure proofs. Fortunately, with the aid 
of isotopic substitution and a technique known as " double resonance," it 
is now possible to analyze completely spectra that show literally hundreds of 
lines. The ways of doing this are beyond the scope of this book. However, it is 
important to recognize that no matter how complex an nmr spectrum appears 
to be, it can be analyzed in terms of just three elements : chemical shifts, 
spin-spin splittings, and kinetic (reaction-rate) processes. 

The kind of nmr spectroscopy we shall discuss here is limited in its applica- 
tions, because it can only be carried on with liquids or solutions. Fortunately, 
the allowable range of solvents is large, from hydrocarbons to concentrated 
sulfuric acid, and for most compounds it is possible to find a suitable solvent. 

A. THE CHEMICAL SHIFT 

Ethyl alcohol, CH 3 — CH 2 — OH, has three kinds of hydrogens: methyl 
(CH 3 ), methylene (CH 2 ), and hydroxyl (OH). In a magnetic field, the nuclei 
(protons) of each of these kinds of hydrogens have slightly different magnetic 
environments as the result of the motions of their valence electrons and those 
of neighboring atoms in response to the magnetic field. The magnetic field 
strength at a particular nucleus is usually less than the strength of the applied 
external magnetic field, because the motions of the electrons result in a shield- 
ing effect (the so-called diamagnetic shielding effect). The important point is 
that the effects arising from the motions of the electrons will be different for 
each kind of hydrogen and, therefore, the resonance signal produced for each 
kind of hydrogen will come at different field strengths. A plot of signal against 
field strength (Figure 7T2) thus shows three principal groups of lines for ethyl 
alcohol. The areas under the curves as measured by the stepped line of the 
chart ("the integrals") correspond to the three varieties of hydrogen. 

Differences in the field strength at which signals are obtained for nuclei of 
the same kind, such as protons or 19 F, but located in different molecular 
environments, are called chemical shifts. 

Chemical shifts are always measured with reference to a standard. For 
protons in organic molecules, the customary standard is tetramethylsilane, 
(CH 3 ) 4 Si, which has the advantage of giving a strong, sharp nmr signal in a 
region where only a very few other kinds of protons absorb. Chemical shifts 
are usually expressed in hertz (Hz, or cycles per second) relative to tetramethyl- 
silane (TMS). These may seem like odd units for magnetic field strength but 
since resonance occurs at a radio frequency, the use of either frequency units 
(hertz, cycles per second) or magnetic field units (gauss) is appropriate. 

Most nmr spectrometers for routine use operate with radio frequency (rf) 



sec 7.6 nuclear magnetic resonance spectroscopy 171 



200 



150 



CAU 



\ l! fi\ 



yy\~ 



3.0 



CM. 






100 Hz 



■Mr 



L_/ 



fVV: 



2.0 



ppm 



Figure 7-13 Nuclear magnetic resonance spectrum of iodoethane, CH 3 CH 2 I, 
at 60 MHz relative to TMS, 0.00 ppm. 



oscillators set at 30, 60, 100, 220, or 300 megahertz (MHz). Since chemical 
shifts turn out to be strictly proportional to the spectrometer frequency, we 
expect lines 100 Hz apart at 60 MHz to be 167 Hz apart at 100 MHz. To 
facilitate comparisons between chemical shifts measured at different frequen- 
cies, shifts in hertz are often divided by the oscillator frequency and reported 
as ppm (parts per million). Thus, if a proton signal comes at 100 Hz at 60 
MHz downfield (toward higher frequencies), relative to tetramethylsilane, it 
can be designated as being ( + 100 Hz/60 x 10 6 Hz) x 10 6 = +1.67 ppm 
relative to tetramethylsilane. 1 At 100 MHz the line will then be 1.67 x 100 x 
10 6 x 10~ 6 = 167 Hz downfield from tetramethylsilane. A table of typical 
proton chemical shifts relative to TMS is given in Table 7-2. The values quoted 
for each type of proton may, in practice, show variations of 5 to 20 Hz. This is 
not unreasonable, because the chemical shift of a given proton is expected to 
depend somewhat on the nature of the particular molecule involved and also 
on the solvent, temperature, and concentration. 

B. SPIN-SPIN SPLITTING 

We have noted that organic molecules with protons on contiguous carbon 
atoms, such as ethyl derivatives CH 3 CH 2 X (X # H), show principal resonance 
signals for protons of different chemical shifts (see Figure 7T2). Each of these 
signals is actually a group of lines that results from " spin-spin splitting." 
Taking as a typical example the protons of iodoethane (Figure 7T3), the 



1 In the past, a ppm scale was more commonly used than at present, based on so-called 
"t values." The t scale has the TMS reference at +10, so that most proton signals fall in 
the region of to + 10 t. A t value can be converted to ppm, with TMS at 0.0, by subtracting 
it from 10. 



chap 7 isolation and identification of organic compounds 172 



Table 7*2 Typical proton chemical shift values 
(dilute chloroform solutions) 



type of proton" 


chemical shift 6 


type of proton" 


chemical shift* 


ppm 


Hz' 


ppm 


Hz c 


R-CH 3 


0.9 


54 


0=C-CH 3 

1 
R 


2.3 


126 


R-CH 2 -R 


1.3 


78 














R-CH 2 -C1 


3.7 


220 


R 3 CH 


2.0 


120 














R-CH 2 -Br 


3.5 


210 


R 2 C=CH 2 


~5.0 


300 














R-CH 2 -I 


3.2 


190 


R 2 C=CH 

1 


~5.3 


320 








i 
R 






RCH(-Cl) 2 d 


5.8 


350 


HC— CH 

/ ^ 

HC CH 

\ / 

HC=CH 






R-O-CH3 


3.8 


220 


7.3 


440 


(R-0-) 2 CH 2 d 


5.3 


320 






R-C-H 

II 
O 


9.7 


580 


R-C=C-H 


2.5 


150 






R 2 C=C-CH 3 

| 


~1.8 


108 


R-O-H 


~5 e 


300" 


R 






HC-CH 
t \ 
HC C-OH 

HC=CH 






HC-CH 

// \ 
HC C-CH 3 

\ / 
HC=CH 






~T 


420* 


2.3 


140 












R-C-OH 

II 
O 


~ll e 


660 e 













" The proton undergoing resonance absorption is shown in heavy type. The group R denotes a 
saturated hydrocarbon chain. 

* Relative to tetramethylsilane as 0.00 ppm. 

c Spectrometer frequency, 60 MHz (14,100 gauss magnetic field). 

d Note how the shift produced by two chlorines or two RO — groups is greater than, but by 
no means double, that produced by one chlorine or RO — group. 

e Sensitive to solvent, concentration, and temperature. 



chemical-shift difference between the methyl and methylene protons gives 
the two main groups of lines. These are split ("first-order" effect) into equally 
spaced sets of three and four lines by mutual magnetic interactions that are 
called " spin-spin interactions." Several of these lines are further discernibly 
split as the result of "second-order" spin-spin splitting. 

How do we know what we are dealing with when there are so many lines 
present ? First, the chemical shift is easily recognizable as such by the fact that 
the spacing between the main groups is directly proportional to the oscillator 



sec 7.6 nuclear magnetic resonance spectroscopy 173 

frequency v. If we double v, the spacing doubles. In contrast, the line spacings 
for the first-order splitting are independent of v, and for this reason the 
first-order splitting is easily recognized, also. Finally, the second-order 
splitting turns out to depend on v for rather complicated reasons ; it tends to 
disappear as v is increased. (See Figure 7-14.) 

It can be shown by isotopic substitution with heavy hydrogen (deuterium, 
D) that the three-four pattern of lines observed for spin-spin splitting with 
compounds having ethyl groups (XCH 2 CH 3 , see Figure 7-13) arises from 
magnetic interaction of each group of protons with the other. Deuterons have 
much smaller magnetic moments than protons, and substitution of one deu- 
teron on the methyl of an ethyl group (XCH 2 CH 2 D) reduces the resonance of 
the methylene group to a triplet (actually somewhat broadened because of the 



Figure 7-14 Comparison of the pmr spectra of 2-methyl-2-butanol at rf oscil- 
lator frequencies of 60, 100, and 220 MHz. The line at 16S Hz in the 60-MHz 
spectrum is due to the OH protons, and this is off-scale to the left in the 220- 
MHz spectrum. The large single line in the center of the spectra arises from 
the resonances of the six methyl hydrogens. 



20 MHz 



Jl_J 



100 MHz 



~J 



JV\J% 



60 MHz 



__jlJ^ 



*_ 



400 



200 



Hz 



chap 7 isolation and identification of organic compounds 174 

small magnetic effect of the deuteron) ; substitution of two deuterons (XCH 2 
CHD 2 ) produces a doublet resonance with the splitting caused by the re- 
maining proton. Three deuterons (XCH 2 CD 3 ) give a one-line XCH 2 spectrum 
(see Figure 7-15). Thus, for this particular simple case, the multiplicity of 
lines can be seen to be (n + 1) where n is the number of protons on contiguous 
carbons. That the methylene resonance of an ethyl group is not complicated 
beyond the observed quartet by interaction of the methylene protons with 
each other is, for our purposes here, best condensed into a simple catechism. 
Protons with the same chemical shift do not normally split one another's 
absorption lines. Thus, only single resonance lines are observed for H 2 , 
CH 4 , C 2 H 6 , (CH 3 ) 4 Si, and so on, and we say that the protons in such com- 
pounds are equivalent. 

In general, the magnitude of the spin-spin splitting effect of one proton on 
another proton (or group of equivalent protons) depends on the number and 
kind of intervening chemical bonds, and on the spatial relations between the 
groups. For nonequivalent protons on adjacent saturated carbon atoms, the 
so-called three-bond splitting is normally about 5-8 Hz. 



H-C-C-H 



5-8 
Hz 

For protons separated by more than three bonds, the coupling is usually too 
small for observation unless a double or triple bond intervenes. 

The ratios of the line intensities in spin-spin splitting patterns usually follow 
simple rules when the chemical shifts are large with respect to the splittings. 
A symmetrical doublet is produced by a single proton, a 1 : 2 : 1 triplet by 
two protons in a group, a 1:3:3:1 quartet by three protons in a group, 
1:4:6:4:1 quintet by four protons, and so on. The intensities follow the 
binomial coefficients. 

The spectrum of (CH 3 0) 2 CHCH 3 (Figure 7-16) provides an excellent 
example of how nmr shows the presence of contiguous protons. The symmet- 
rical doublet and 1:3:3:1 quartet are typical of interaction between a single 
proton and an adjacent group of three. The methyl protons of the CH 3 
groups are too far from the others to give demonstrable spin-spin splitting. 



C. USE OF NUCLEAR MAGNETIC RESONANCE SPECTROSCOPY 
IN QUALITATIVE ANALYSIS 

The solution of a typical qualitative analysis problem by nmr can be illustrated 
with the aidof the spectrum shown in Figure 7-17. Here, we see three principal 
groups of lines at 9.8, 2.4, and 1.0 ppm for a compound of formula C 3 H g O. 
The relative heights of the principal steps of the integrated spectrum show these 
groups to arise from one, two, and three hydrogens, respectively. The single 



sec 7.6 nuclear magnetic resonance spectroscopy 175 



X— CH 2 — CH, 1 M I 



X— CH 2 — CH 2 D ■ 



X— CH,— CHD, 



X— CH, — CD, 



CH 2 



i_J_L 



i i i 



CH 3 



Figure 7-15 Schematic spectra of deuteriated ethyl derivatives. The right-hand 
set of lines is always a triplet when observable because of the two protons of 
the X — CH 2 — group. 



Figure 7-16 Nuclear magnetic resonance spectrum of dimethyl acetal, 
(CH 3 0) 2 CHCH 3 , at 60 MHz relative to TMS, 0.00 ppm. 



250 



CM 



4.0 



200 



150 



100 



50 



iGCIl. 



CH 




f 



3.0 



2.0 



1.0 



Hz 



TMS 

* 



ppm 



chap 7 isolation and identification of organic compounds 176 



~i 1 r 

586 



150 



100 





50 



Hz 



CH,— C- 




TMS 



J IL 



9.77 



2.0 



1.0 



ppm 



Figure 7-17 Nuclear magnetic resonance spectrum and integral for compound 
of formula C 3 H 6 Q at 60 MHz relative to TMS. 



hydrogen at 9.8 ppm is seen from Table 7-2 to have a chemical shift compatible 
with either RCHO or RC0 2 H. The latter possibility is, of course, excluded 
because the compound has only one oxygen. The only structure that can be 
written for C 3 H 6 possessing an RCHO group is CH 3 CH 2 CHO (propion- 
aldehyde), and this structure is completely compatible with the other features 
of the spectrum. Thus, the CH 3 — resonance comes at 1.0 ppm (0.9 ppm 
predicted for CH 3 R) and the — CH 2 — resonance at 2.4 ppm (2.3 ppm pre- 
dicted for CH 3 COR). 

The spin-spin splitting pattern agrees with the assigned structure, there 
being the same 7 Hz spacings as in the characteristic three-four pattern of the 
CH 3 CH 2 ~ group shown by iodoethane (Figure 7-13). The doubling up 
(somewhat obscured by second-order splitting) of each of the — CH 2 — 
resonance lines is due to a small (~2 Hz) coupling between the — CHO and 
— CH 2 — protons. This interaction also causes the —CHO resonance to be 
split into a 1:2:1 triplet, as expected from the n + 1 rule. Three-bond 
couplings between —CHO and adjacent — CH 2 — protons appear to be 
generally much smaller than — CH 2 — CH 3 couplings. 



D. NUCLEAR MAGNETIC RESONANCE AND RATE PROCESSES 



An nmr spectrometer is unusual among instruments used to study molecules 
through absorption of electromagnetic radiation in that it acts like a camera 
with a relatively slow shutter speed. In fact, its "shutter speed" is quite 
commonly about the same as a conventional camera having exposure times 



sec 7.6 nuclear magnetic resonance spectroscopy 177 

of a 100th of a second or so. When we take an nmr spectrum of a molecule 
that is undergoing any rapid motion or reaction, the result is something like 
taking a picture of a turning spoked wheel with a box camera. If the wheel 
turns only once a minute, a photograph at a 100th of a second will show the 
individual spokes without much blurring. On the other hand, if the wheel 
turns 100 or more times a second, then a photograph does not show the in- 
dividual spokes at all, but only the average outline of the rim and hub as a 
border to the gray of the spokes. Pictures showing blurred individual spokes 
result only when the wheel is turning neither very rapidly nor very slowly in 
relation to the camera shutter speed. 

A vivid example of the use of nmr in the study of motions within molecules 
is given by the fluorine ( 19 F) resonance spectrum of l-chloro-2-fluoro-l,l,2,2- 
tetrabromoethane. This molecule exists in three staggered rotational confor- 
mations [4-6]. Of these, [5] and [6] will have identical nmr spectra because for 
each the fluorine on one carbon is equivalently located with respect to the 
halogens on the other carbon. However, the fluorine of [4] has a different 






environment, being located between Br and Br, and should have a different 
chemical shift from the fluorines of [5] and [6]. In principle, therefore, one 
would expect to observe separate nmr resonances for the isomers ; in fact, at 
122°, rotation around the C— C bond occurs so rapidly that only a single 
fluorine resonance line is obtained (see Figure 7-18). The observed line position 
is an average position, the location of which depends on the lengths of time 
the molecules exist separately as [4], [5], and [6]. The rate of rotation about 
the C— C bond becomes slower with decreasing temperature, and is so slow 
at —40° that you can observe the separate resonances of [4] and the isomers 
[5] and [6] as shown in Figure 7-18. Studies of the changes in line shape of the 
nmr spectrum of l-chloro-2-fluoro-l,l,2,2-tetrabromoethane with temperature 
provide a means of evaluating the amount of energy the molecules must have 
to allow rotation to take place. Because the halogens are large and do not 
move past each other easily, it takes fully 1 5 kcal/mole of thermal energy to 
interconvert [4], [5], and [6]. At 25°, the rate of interconversion is about 100 
times per second. 

The loss of identity of particular protons by rapid rate processes makes the 
nmr spectra of ethyl derivatives much simpler than they would otherwise be. 
Inspection of a ball-and-stick model of an ethyl derivative in the staggered 
conformation (see Figure 2-3) shows that one of the CH 3 hydrogens (marked 
here with *) should not have exactly the same chemical shift as the other two. 



chap 7 isolation and identification of organic compounds 178 






AiftMfHt'''- 



-,«^»JsW<n 



^v* i %Vi*,v«>,',\Mv< , /** KV >> rfJ * u * ; >■^Ay^ ,w, '* ,, ' 



Figure 7-18 The 19 F spectrum at 56.4 MHz of l-chloro-2-fluoro-l,l,2,2-tetra- 
bromoethane as a function of temperature. 



However, rotation about the single bond in this case is sufficiently fast (10~ 6 
sec) to average out the differences between the protons, and an average 
resonance line position is observed. 



H* 

H, /H 



The nmr spectra of cyclohexane and substituted cyclohexanes have been 
extensively studied in connection with the interconversions of the boat and 
chair forms (Section 3-4B). The proton nmr spectrum of cyclohexane itself 
is a single line at room temperature because of rapid inversion (~10 6 times 
per second) which averages to zero the chemical-shift differences of the equa- 
torial and axial protons (Figure 3-6). At — 100°, the pmr spectrum of cyclo- 



summary 179 

hexane is so complex as to be ilninterpretable because all of the axial protons 
have different chemical shifts from the equatorial protons and especially 
complex spin-spin interactions occur. The device of substituting deuterium 
for hydrogen (Section 7-6B) has particular utility here. Undecadeuteriocyclo- 
hexane, C 6 D n H, gives a single proton nmr line at room temperature, while 
at — 100° and 60 MHz two equally intense lines are observed separated by 
29 Hz. These lines arise from the axial and equatorial protons in [7] and [8], 
respectively. The rate of interconversion of [7] and [8] is on the average about 

D 

[7] [8] 



once per 10 seconds at — 100° so that it can be seen why separate resonance 
lines are obtained for the hydrogens in [7] and [8]. 



summary 

Organic compounds can be isolated and purified by distillation, crystalliza- 
tion, or chromatographic techniques. The latter include (a) multistage liquid- 
liquid extractions ; (b) column chromatography, in which a mixture is separated 
on a column of alumina or other solid absorbent by passage of appropriate 
solvents through the column; (c) vapor-phase chromatography, in which a 
gas stream carries the sample through a column packed with a nonvolatile 
liquid on a porous solid ; and (d) paper and thin-layer chromatography (which 
will be discussed later). 

A compound can often be identified by comparison of its physical and spec- 
troscopic properties with those of known compounds published in the litera- 
ture. If this is unsuccessful, elemental analysis and molecular weight deter- 
mination (by freezing-point depression, etc., or by mass spectrometry) can 
be used to establish the molecular formula and, following this, the structure 
may be deduced by studies of its chemical transformation products, X-ray 
analysis, or spectroscopic analysis. 

The three most important types of spectral analysis in routine use are 
infrared, electronic (ultraviolet and visible), and nuclear magnetic resonance 
spectroscopy. In each the absorption of a photon of electromagnetic radiation 
produces an excited state. 

Absorption of infrared radiation causes changes in the vibrational and 
rotational energy levels of molecules. The stretching vibrations of C— H 
bonds are characterized by absorptions near 3000 cm" l whereas their bending 
vibrations (and both kinds of vibration for other bonds to carbon) occur at 
lower frequencies. For carbon-carbon bonds, the greater the bond strength, 
the higher the frequency of vibration and the greater the energy required for 
excitation to a higher stretching vibrational state. Thus, C=C absorptions 



chap 7 isolation and identification of organic compounds 180 

occur near 2100 cm -1 , C=Cnear 1700 cm -1 , and C— Cat still lower frequen- 
cies. The region below 1250 cm -1 is called the "fingerprint" region because 
here occur complex vibrations characteristic of the molecule as a whole 
rather than of specific bonds. 

Absorption of ultraviolet or visible light causes electronic excitations. The 
usual practical short- wavelength limit for the ultraviolet region is about 1900 
A. Saturated compounds and those containing isolated C=C or C=C bonds 
are usually transparent down to this wavelength, but conjugated polyenes 
(and other unsaturated compounds) show strong absorption above 1900 A. 
In general, the greater the degree of conjugation, the lower will be the energy 
necessary for excitation and the longer will be the wavelength of absorption. 
Colored substances are those which absorb visible light and hence require 
intermediate energy photons for excitation. 

Nuclear magnetic resonance spectroscopy (nmr) involves absorption of 
very low-energy radio-frequency photons by atomic nuclei in an applied 
magnetic field. The magnetic nuclei characteristic of atoms such as 1 H, 19 F, 
and 31 P can be oriented so as to be lined up with or opposed to the applied 
field and hence can differ very slightly in energy. This energy difference will 
depend, among other things, on the atom's environment in the molecule. In 
practice, either the frequency of the radiation or magnetic field can be kept 
constant while the other is varied. The chemical shift for protons is given in 
hertz (cycles per second) or ppm and is relative to some standard, usually 
tetramethylsilane (TMS). In general, the presence of electronegative groups 
or multiple bonds can cause large proton shifts from the standard. 

When nonequivalent protons are attached to adjacent carbon atoms, there 
is a spin-spin splitting of 5-8 Hz. A single proton will cause a split of its 
neighbor's absorption into a symmetrical doublet, a pair of protons produces 
a 1 : 2 : 1 triplet, and so on. 

Rate processes, particularly interconversion of conformational isomers, can 
be examined by nmr because the time interval for absorption of the radio- 
frequency photons used in nmr is often comparable to the rates of confor- 
mational interconversion. 



exercises 

7-1 Suppose you are asked to purify a compound by recrystallization from one 
of these solvents: water, ethanol, benzene, or heptane. The solubilities of the 
compound in grams per 100 ml of solvent at 20° and 80°, respectively, in the 
four solvents are water (0.15, 0.27), ethanol (0.70, 8.2), benzene (16, 38), 
heptane (36, 41). Which solvent would you choose for the purpose? 

7-2 A liquid hydrocarbon, which does not decolorize bromine, has prominent 
bands near 1500 cm -1 and 3000 cm -1 in its infrared spectrum. Its mass 
spectrum shows a parent peak at 92 mass units. Suggest a likely structure 
for this compound. 



exercises 181 

7-3 A compound that is gaseous at room temperature was found on analysis to 
contain 88.9% carbon and 11.1 % hydrogen. Its infrared spectrum contained 
a rather weak band near 2200 cm^ 1 . There are only two known compounds 
that fit this description. What are their names ? 

7-4 Explain how a mass spectrometer, capable of distinguishing between ions 
with mje values differing by 1 part in 50,000, could be used to tell whether 
an ion of mass 29 is C 2 H 5 ffl or CHO®. 

7-5 Calculate the energy in kilocalories which corresponds to the absorption of 1 
einstein of light of 5893 A (sodium D line) by sodium vapor. Explain how this 
absorption of light by sodium vapor might have chemical utility. 

7-6 From the discussion in Section 7-5 about the structures of the ground and 
excited states of butadiene, see if you can rationalize why it is that the degree 
of 7t bonding between the 2,3 carbons is relatively larger in the excited state 
than in the ground state. 

7-7 Sketch out the nmr spectrum and integral expected at 60 MHz withTMS 
as standard for the following substances. Show the line positions in hertz, 
neglecting spin-spin couplings smaller than 1 to 2 Hz and all second-order 
effects. Note that chlorine, bromine, and iodine (but not fluorine) act as 
nonmagnetic nuclei. 

a. CH 3 C1 /. CHCl 2 CHBr 2 

b. CH 3 CH 2 C1 g. CH 3 (CH 2 ) 6 CH 3 

c. (CH 3 ) 2 CHC1 h. C1CH 2 CH 2 CH 2 I 

d. CH 3 CD 2 CH 2 C1 i. (C1CH 2 ) 3 CH 

e. (CH 3 ) 3 CC1 

7-8 Figure 7-19 shows nmr spectra and integrals at 60 MHz for three simple 
organic compounds. Write a structure for each substance that is in accord 
with both its molecular formula and nmr spectrum. Explain how you 
assign each of the lines in the nmr spectrum. 

7-9 Figure 7-20 shows the nmr spectrum of a compound, C 5 H S 2 . Which of the 
following structures fits the spectrum best? Explain. 

CH 3 CH=CHC0 2 CH 3 CH 2 =CHCH 2 C0 2 CH 3 

CH 2 =CHC0 2 CH 2 CH 3 HC0 2 CH 2 CH 2 CH =CH 2 

CH 2 =C(CH 3 )C0 2 CH 3 1 1 

CH 2 =C(OCH 3 )COCH 3 OCH 2 CH 2 CH 2 CH 2 C=0 

(CH 2 ) 2 CHC0 2 CH 3 

7-10 In reasonably concentrated solutions in water, acetic acid acts as a weak acid 
(less than 1 % dissociated). Acetic acid gives two nmr resonance lines at 2 
and 1 1 ppm, relative to TMS, while water gives a line at 5 ppm. Nonetheless, 
mixtures of acetic acid and water are found to give just two lines. The position 
of one of these lines depends on the ratio of acetic acid to water concentration, 
while the other one does not. Explain and show how you would expect the 
position of the concentration-dependent line to change over the range of 
acetic acid concentrations from to 100%. 



chap 7 isolation and identification of organic compounds 182 






5.0 


3.0 


1.0 ppm 




.1 

200 

C,H,.Bi 


I 1 
150 100 

> 


1 
50 1 


1 
D 11/ 


f 








_^-— ' 


1 . 1 


>~- 




„ -,-. ., , ^AAM-u.-, ,,J 


l t 


1 1 


1 


1 





4.0 



3.0 



2.0 



1.0 



ppm 





' T" 
200 






150 




-I 

100 llz 


CH.RiI 










-— 




' 1 I 


,- 


— 


__ 





■■ * ^ ' 




1 i f* 














u 














in 


Sjlljjl 








^^^^^^^^M 




i ill 1 

H:!;li, 










II 11 i 
A n 1 / 




Hi V V 

1 1 


v— ™ 






i 


, „ wtirn. 


\^_j\ 







4.0 3.0 2.0 ppm 

Figure 7-19 Nuclear magnetic resonance spectra and integrated spectra of 
some simple organic compounds at 60 MHz relative to TMS, 0.00 ppm. See 
Exercise 7-8. 



exercises 183 




ppm 



Figure 7-20 Spectrum of a compound C 5 H 8 2 at 60 MHz relative to TMS as 
standard. See Exercise 7-9. 



7-11 Sketch out the principal features you would expect for the infrared and nmr 
spectra of the following substances. (It will be helpful to review pages 20 
and 116 as well as Sections 7-4 and 7-6.) 

a. CH3C-CCH3 

b. CH 3 C=CH (expect a long-range nmr coupling of 3 Hz) 

c. CH 3 CH 2 C=CCH 2 CH 3 

d. HC=C-CH=CH-C=CH (cis and trans) 

e. (CH 2 ) 8 III 



IS 






SI 

Si 

111111 

ill 

IIP 



'Ibhhi 

Mil?. *iSKS'p' JH" 

afi- iispi- pp^ia 
ATM Wife Wife s&£ S$fc 






Ses5a&*fe' " " 



in 1 
WTO 

|t«;S! 

Mm 






a^MifM^af ion reactions 




ilft 



SB., -si** 















SUSS® 

IlltSH 
\mm 

mill 
[Slip 

I SKIP 






^«%. ■•id* 






«#S S ASS? 



i : r f.- ■•*?■. -is' ■■!&•;>. "-v.* -.v>j 






chap 8 nucleophilic displacement and elimination reactions 187 

There are relatively few basic types of organic reactions. Of these, substitution, 
addition, and elimination are of the greatest importance. So far, we have 
discussed substitution of halogen for hydrogen, addition reactions of alkenes, 
and addition reactions of cycloalkanes with strained rings. In this chapter, the 
principal topics are the substitution or displacement by nucleophilic reagents 
of groups attached to carbon, and the formation of carbon-carbon double 
bonds by elimination reactions. These reactions are often profoundly in- 
fluenced by seemingly minor variations in structure, reagents, solvent, and 
temperature. It is our purpose to show how these variations can be understood 
and, as far as possible, predicted in terms of the principles we have already 
discussed. Before proceeding, however, it will be helpful to consider the nomen- 
clature of many of the organic reactants and products involved. The reader 
who is already acquainted with these nomenclature systems, or wishes to 
study them at a later time, can pass directly to Section 8-7. 

8-1 organic derivatives of inorganic compounds 

We have so far classified organic compounds by their functional groups; 
alkenes possess double bonds, alkynes triple bonds, and so on. Another useful 
classification of organic compound considers them as substitution products 
of water, ammonia, hydrogen sulfide, nitrous or nitric acids, and so on, 
through replacement of one or more hydrogens with an organic group. 
Reference to Table 8- 1 shows how alcohols, ethers, carboxy lie acids, anhydrides, 
and esters may be regarded as derivatives of water; mercaptans and sulfides 
as hydrogen sulfide derivatives; amines and amides as ammonia derivatives; 
alkyl nitrates as derivatives of nitric acid; alkyl nitrites as derivatives of 
nitrous acid; and alkyl sulfates as derivatives of sulfuric acid. For the sake 
of completeness, we include alkyl halides, which we have already classed as 
substituted alkanes, but which may also be considered as derivatives of the 
hydrogen halides. 

8-2 alcohol nomenclature 

In naming an alcohol by the IUPAC system, the ending -ol is appended to 
the name of the parent hydrocarbon. The latter corresponds to the longest 
straight chain of carbon atoms that includes the carbon carrying the hydroxyl 
group; it also includes the double bond when the compound is unsaturated. 
Note also that the -ol function normally takes precedence over a double bond, 
halogen, and alkyl in determining the suffix of the name. With respect to 
numbering, the carbon carrying the hydroxyl group is taken as number one 
if it terminates the chain, or the lowest number thereafter if it is attached to 
the nonterminal carbon. 

OH CH 2 CH 3 

I I 

CH 3 -OH CH 3 CH 2 -OH C 6 H 5 CH 2 CHCH 2 CH 3 C1CH 2 CH = CCH 2 0H 

methanol ethanol 1 -phenyl- 4-chloro-2-ethyl- 

2-butanol 2-buten-l-ol 

The most commonly used system for naming alcohols (and halides) com- 
bines the name of the appropriate alkyl group with the word alcohol (or 



chap 8 imcleophilic displacement and elimination reactions 188 
Table 8-1 Organic compounds as derivatives of common inorganic compounds 



parent 
compound 






organic derivative 




class of compound 


example 




H-O-H 


R— O— H 




alcohol 


CH 3 OH 


methanol 




R-0-R< 




ether 


CH3OCH3 


dimethyl ether 




O 

II 

R-C-O- 






O 

|| 






-H 


carboxylic acid 


CH3COH 


acetic acid 




O 

II 

R-C-O- 






O 

|| 






-R' 


carboxylic ester 


CH3COCH3 


methyl acetate 




O 

II 

R-C-O- 


O 

II 
-C-R 


carboxylic 
anhydride 


O O 

II II 

CH3COCCH3 


acetic anhy- 
dride 


H-S-H 


R-S-H 




thiol 
(mercaptan) 


CH 3 SH 


methanethiol 
(methyl 
mercaptan) 




R-S-R' 




thioether 
(sulfide) 


CH3SCH3 


methylthio- 
methane (di- 
methyl 
sulfide) 




O 

II 
R-C-S- 


H 


thio acid 


O 

II 
CH3CSH 


thioacetic acid 


NH 3 


RNH 2 




prim, amine 


CH 3 NH 2 


methylamine 




R 2 NH 




sec. amine 


(CH 3 ) 2 NH 


dimethylamine 




R 3 N 




tert. amine 


(CH 3 ) 3 N 


trimethylamine 




O 

II 
R-C-NH 


[ 2 


acylamine (un- 
substituted 
amide) 


O 

II 
CH 3 -C-NH 2 


acetamide 




O 

II 






O 

II 






R— C— NHR 


(monosubsti- 
tuted amide) 


CH3-C-NHCH3 


N-methyl- 
acetamide 




O 
II 
R— C— NR 


2 


(disubstituted 
amide) 


O 

II 
CH 3 -C-N(CH 3 ) 2 


N,N-dimethyl- 
acetamide 


H-ON0 2 

(nitric acid) 


R-ON0 2 




alkyl nitrate 


CH 3 -ON0 2 


methyl nitrate 


H— ONO 
(nitrous acid) 


R-ONO 




alkyl nitrite 


CH3-ONO 


methyl nitrite 


H-N0 2 

(unknown) 


R-N0 2 




nitroalkane 


CH 3 -N0 2 


nitromethane 



sec 8.2 alcohol nomenclature 189 



Table 8-1 Organic compounds as derivatives of common inorganic 
compounds {continued) 



parent 
compound 




organic derivative 




class of compound exampl 


e 


H-NO (hypo- 
nitrous acid, 


R-NO 


nitrosoalkane CH 3 — NO 


nitroso- 
methane 


monomenc 
form) 








O 

II 
H— O— S— OH 

II 

O 
(sulfuric acid) 


O 
II 
R-O-S-OH 

II 
O 


o 

II 
alkyl hydrogen CH 3 — O-S— OH 

sulfate II 
O 


methyl hydrogen 
sulfate 




O 

11 

R-O-S-O-R 

II 

o 


O 

II 
dialkyl sulfate CH 3 -0— S-OCH 3 
II 
O 


dimethyl 
sulfate 


H-X 

(X=F, CI, 
Br, I) 


R-X 


alkyl halide CH 3 — F 
(haloalkane) 


methyl fluoride 
(fluoro- 
methane) 



halide). The system works well whenever the group name is simple and 
easily visualized : 



CH 3 
I 
CH 3 — C— OH 
I 
CH 3 

/-butyl alcohol 



CH 3 
I 
CH 3 — C— CH 2 — OH 

I 
H 

isobutyl alcohol 



CH 2 =CH-CH 2 -C1 



allyl chloride 



A prevalent but unofficial procedure names alcohols as substitution prod- 
ucts of carbinol, CH 3 OH, a synonym of methanol. Many alcohols that are 
cumbersome to name by the IUPAC system may have structures that are 
more easily visualized when named by the carbinol system. 



H 3 C H CH 3 

III 
CH 3 — C— C— C — CH 3 

III 
H 3 C OH CH 3 

di-r-butylcarbinol 

(2,2,4,4-tetramethyl-3- 

pentanol by 

IUPAC system) 



C 6 H 5 C CH 2 CH — CH 2 

OH 

allylphenylcarbinol 

( 1 -phenyl-3-buten- 1 -ol 

by IUPAC system) 



The carbinol system has been extended to include other derivatives. Alkyl 



chap 8 nucleophilic displacement and elimination reactions 190 



halides and alkylamines, for example, are frequently called carbinyl halides 
and carbinylamines. 



/ \-CH 2 Cl 
cyclohexylcarbinyl chloride 



C 6 H 5 CH CH 2 CH 3 

NH 2 
ethylphenylcarbinylamine 



8-3 ether nomenclature 

Symmetrical ethers (both R groups in R— O— R being the same) are named 
simply dialkyl, dialkenyl, or diaryl ethers, as the case may be. The prefix di- 
to denote disubstitution is sometimes omitted as superfluous, but most 
current opinion regards this form of redundancy desirable to help prevent 
errors. Clearly, when an ether is unsymmetrical, the names of both R groups 
must be included: 



CH 3 CH 2 -0-CH 2 CH 3 

diethyl ether 



CH 3 -0-CH = CH 2 

methyl vinyl ether 



diphenyl ether 



8-4 carboxjlic acid nomenclature 

According to the IUPAC system, carboxylic acids with saturated alkyl 
chains are called alkanoic acids. The suffix -ok is added to the name of the 
longest continuous-chain hydrocarbon in the molecule that includes thecarbon 
of the carboxyl (— C0 2 H) group. Note that the carboxyl function normally 
takes precedence over the hydroxyl function and that the carboxyl carbon 
is C-l. 



CI 

I 



2-chloropentanoic acid 



CH 3 OH 

I I 

CH 3 CHCH= CHCHC0 2 H 

2-hydroxy-5-methyl-3-hexenoic acid 



The simple alkanoic acids have long been known by descriptive but 
unsystematic names that correspond variously to their properties, odors, or 
natural origin. It seems likely that for the following acids these trivial names 
will not be completely superseded by the more systematic names (shown in 
parentheses) for some time to come. 



HC0 2 H 

CH 3 C0 2 H 

CH 3 CH 2 C0 2 H 

CH 3 CH 2 CH 2 C0 2 H 

H 3 C 
\ 
CHC0 2 H 



formic acid 
acetic acid 
propionic acid 
butyric acid 



(methanoic acid) 
(ethanoic acid) 
(propanoic acid) 
(butanoic acid) 



(L. formica, ant) 
(L. acetum, vinegar) 
(proto + Gr. pion, fat) 
(L. butyrum, butter) 



isobutyric acid (2-methylpropanoic acid) 



H 3 C 



sec 8.6 single- or multiple-word names 191 

Esters of carboxylic acids carry the suffix -oate in place of -oic (or -ate in 
place of -ic for acids with descriptive names). 

O O CI o 

II II I II 

CH 3 CH 2 COCH 3 C 6 H 5 CH 2 COCH 2 CH 3 CH 2 = CHCHCOCH 2 CH 3 

methyl propionate ethyl phenylacetate ethyl 2-chloro-3-butenoate 

(methyl propanoate) 

Alkanoic acids are sometimes called aliphatic acids (or fatty acids) to 
distinguish them from those containing a benzene ring and which are called 
aromatic acids. The term aliphatic is often used for noncyclic systems in 
general. 

8-5 the use of greek letters to denote 
substituent positions 

Considerable use is made of the Greek letters a, fl, y, and so on to designate 
successive positions along a hydrocarbon chain. The carbon directly attached 
to the principal functional group is denoted as a, the second as /?, and so on. 

CH 3 CH 3 

I I 

CH 2 = CH-C-OH Br 2 CH-C-C0 2 H 
I I 

CH 3 Br 

a.a-dimethylallyl alcohol a,/J,/J-tribromoisobutyric acid 

(2-methyl-3-buten-2-ol) (2,3,3-tribromo-2-methyl- 

propanoic acid) 

Note that the a position of an acid is at the number 2 carbon atom. In general, 
the use of these names is to be deplored, but since it is widespread, cognizance 
of the system is important. 

8-6 single- or multiple-word names 

A troublesome point in naming chemical compounds concerns the cir- 
cumstances that govern whether a compound is written as a single word (e.g., 
methylamine, trimethylcarbinol) or as two or more words (e.g., methyl 
alcohol, methyl ethyl ether). When a compound is named as a derivative of 
substances such as methane, ammonia, acetic acid, or carbinol because of the 
substitution of hydrogen for some other atom or group, its name is written 



(C 6 H 5 ) 3 CH CH 3 - 


-N- 
H 


-C 2 H 5 


triphenylmethane methylethylamine 


CH 2 = CH-CH 2 -C(CH 3 ) 2 




CH 2 = CH-CH 2 Q 


OH 






dimethylallylcarbinol 




vinylacetic acid 


CH 3 MgI 




C 6 H 5 Li 


methylmagnesium iodide 


phenyllithium 



chap 8 nucleophilic displacement and elimination reactions 192 

as a single word. This is correct because we do not speak of" a methane " but 
of the compound "methane." It follows that a derivative such as (C 6 H 5 ) 3 CH 
is called triphenylmethane and not triphenyl methane. We do speak, however, 
of an alcohol, an ether, a halide, acid, ester, sulfide, or ketone, for these words 
correspond to types of compounds rather than particular compounds. There- 
fore additional words are required to fully identify particular alcohols, 
ethers, halides, and so on. Several examples are given. 



ethyl iodide 



O 

II 

CH3COH 

acetic acid 



(CH 3 ) 2 CHOH 

isopropyl alcohol 



O 

II 

CH3COCH3 

methyl acetate 



O 

II 
C( 

methyl ethyl ketone 



dimethyl sulfide 



CH 3 OCH 2 CH 3 

methyl ethyl ether 



O 

II 

(CH 3 C) 2 

acetic anhydride 



nucleophilic displacement reactions 



8-7 general considerations 

Broadly defined, a displacement reaction involves the replacement of one 
functional group (X) by another (Y) : 



RX + Y 



RY + X 



We are here concerned with nucleophilic displacement reactions of alkyl 
derivatives; these are ionic ox polar reactions involving the attack by a nucleo- 
phile (i.e., an electron-pair donating reagent) at carbon. A typical example is 
the reaction of hydroxide ion with methyl bromide to displace bromide 
ion. The electron pair of the C— O bond to be formed can be regarded as 



H:0: 9 CH, 



:Br: 



CH 3 :0:H 



donated by the hydroxide ion, whereas the electron pair of the C— Br bond 
to be broken departs with the leaving bromide ion. The name for this type 
of reaction is abbreviated S N , S for substitution and N for nucleophilic. 

A number of nucleophilic reagents commonly encountered in S N reactions 
are listed in Table 8-2 along with the names of the products obtained when 
they react with methyl chloride. The nucleophile may be an anion, Y: e , or a 



sec 8.7 general considerations 193 
Table 8'2 Typical S N displacement reactions of alkyl halides, RX 



1. 


R 


:X + Y:e 


->■ R:Y + X=e 




nucleophilic agent 


product 




product name, R = CH 3 


useful solvents 


Cl e 


RC1 




methyl chloride 


acetone, ethanol 


Br e 


RBr 




methyl bromide 


acetone, ethanol 


I e 


RI 




methyl iodide 


acetone, ethanol 


e OH 


ROH 




methyl alcohol 


water, dioxane-water 


e OCH 3 


ROCH 3 




dimethyl ether 


methyl alcohol 


e SCH 3 


RSCH 3 




dimethyl sulfide 


ethyl alcohol 


P P 
// // 
CH 3 — C RO — C 

V CH 3 




methyl acetate 


acetic acid, ethanol 


e :C=N 


RCN 




acetonitrile 


acetone, dimethyl 
sulfoxide 


HC=C: e 


RC^CH 




propyne 


liquid ammonia 


e :CH(C0 2 C 2 H 5 ) 2 


RCH(C0 2 C 2 H 5 ) 2 


diethyl methylmalonate 


ethyl alcohol 


e :NH 2 


RNH 2 




methylamine 


liquid ammonia 


e © e 
:N=N=N: 


RN 3 




methyl azide 


acetone 


O 

II 


O 

II 








II 
o 


II 

II 




) 


N-methylphthalimide 


N,N-dimethylforma- 
mide 


N0 2 e 


RN0 2 




nitromethane 


N,N-dimethylforma- 
mide 


2. 


R 


:X + Y: 


© e 
-s- R:Y + X: 




nucleophilic agent 


product 




product name, R = CH 3 


useful solvents 


(CH 3 ) 3 N: 


RN(CH 3 ), 


e 
X 


tetramethylammonium 
chloride 


ether, benzene 


(C 6 H 5 ) 3 P: 


RP(C 6 H 5 ) 


e 
3 X 


triphenylmethylphos- 
phonium chloride 


ether, benzene 


(CH 3 ) 2 S: 


RS(CH 3 ) 2 


e 
X 


trimethylsulfonium 
chloride 


ether, benzene 



chap 8 nucleophilic displacement and elimination reactions 194 
Table 8-2 Typical S N displacement reactions of alkyl halides, RX {continued) 



3. 


R: 


:X + H.Y: ~> 


e 
R.Y 


:;H + X:e -> 


R 


Y: +H 


:X 


nucleophilic 


agent 


product 




product name, 


R = 


= CH 3 


useful solvents 


H 2 




ROH, 




methyl alcohol 






water, dioxane-water 


CH 3 OH 




ROCH 3 

O 

II 

ROCCH3 




dimethyl ether 






methyl alcohol 


CH 3 C0 2 H 






methyl acetate 






acetic acid 


NH 3 




RNH 2 




methylamine 






ammonia, methanol 



neutral molecule, Y: or HY:, and the operation of each is illustrated in the 
following general equations for a compound RX : 



R— X + Y: e 
R— X + Y: 
R — X + HY: 



R— Y + X: fc 



R — Y 



X: fc 



-+ RYH + X: e 



-♦ RY: + HX 



The wide range of products listed in Table 8-2 shows the synthetic utility of 
S N reactions. Displacement can result in the formation of bonds between 
carbon and chlorine, bromine, iodine, oxygen, sulfur, carbon, nitrogen, and 
phosphorus. 

Nucleophilic displacements are by no means confined to alkyl halides. 
Other alkyl derivatives include alcohols, ethers, esters, and "onium ions." 1 
Some illustrative reactions of several different alkyl compounds with various 
nucleophiles are assembled in Table 8-3. 

As we shall see in a later section, the mechanism of an S N reaction and the 
reactivity of a given alkyl compound RX toward a nucleophile Y depend 
upon the nature of R, X, and Y, and upon the nature of the solvent. For 
reaction to occur at a reasonable rate, it is very important to select a solvent 
that will dissolve both the alkyl compound and the nucleophilic reagent. 
Furthermore, the nucleophile may require considerable assistance from the 
solvent in breaking the slightly polar C— X bond. However, the highly polar 



Examples of -onium cations follow: 



R 4 N® 
Tetraalkylammonium 



Tetraalkylphosphonium 



R 3 0® 
Trialkyloxonium 



R3C 95 
Trialkylcarbonium 



R 3 S® 
Trialkylsulfonium 



R _ N = N : 
Alkyldiazonium 



R 2 I® 
Dialkyliodonium 



sec 8.8 mechanisms of S N displacements 195 

nucleophilic agents most used (e.g., NaBr, NaCN, H 2 0) are seldom soluble 
in the solvents that best dissolve slightly polar organic compounds. In 
practice, relatively polar solvents, or solvent mixtures, such as acetone, 
aqueous acetone, ethanol, aqueous dioxane, and so on are found to provide 
the best compromise for reactions between alkyl compounds and salt-like 
nucleophilic reagents. A number of useful solvents for typical S N reactions are 
listed in Table 8-2. 



8-8 mechanisms of S N displacements 

Two mechanisms may be written for the reaction of methyl chloride with 
hydroxide ion in aqueous solution that differ in the timing of bond breaking 
in relation to bond making. In the ensuing discussion, we are not implying 
that these are the only possible mechanisms that one could conceive for this 
reaction. They appear to be the most plausible ones and by confining our 
attention to two possibilities we can greatly simplify the discussion. In the 
first mechanism, A, the reaction is written as taking place in two steps, the 
first of which involves a slow and reversible dissociation of methyl chloride 



Table 8-3 S N displacement reactions of various types of compounds, RX 



type of compound, RX 



alkyl chloride 
alkyl bromide 
alkyl iodide 

dialkyl sulfate 



benzenesulfonate 
ester 



acetate ester 

alcohol 

ether 

ammonium ion 

iodonium ion 

diazonium ion 



R-Cl + I e ; 
R-Br + I e ; 
R-I + CH 3 6 



RI + Cl e 
RI + Br e 
->■ ROCH 3 + 1 € 



R-OS0 2 OR + CH3O - 

O 

R-OShTj> + H 2 

o 



o 

II 

R-OCCH3 + H 2 
R-OH + HBr — 
R-OR' + HBr — 



-► ROCH 3 + e OS0 2 OR 



O 

II /=\ 
-> ROH + HOS-d J> 

O 



o 

li 

-► ROH + HOCCH3 



-> RBr + H 2 



R-NR 3 + HO e 
R-I-R' + OH e 

R-N=N + H 2 



-* RBr + R'OH 
—* ROH + NR 3 
— + ROH + R'l 



-> ROH + H ffi + N 2 



chap 8 nucleophilic displacement and elimination reactions 196 

to methyl cation and chloride ion. The second step involves a fast reaction 
between methyl cation and hydroxide ion (or water) to yield methanol. 

Mechanism A: 

slow 



CH,-C1 



e fast 



CH/ + OH e ' CH3OH 

© OH 9 

(or CH 3 e + H 2 CH 3 OH 2 » CH 3 OH + H 2 0) 

In the second mechanism, B, the reaction proceeds in a single step. Attack 
of hydroxide ion at carbon occurs simultaneously with the loss of chloride 
ion; that is, the carbon-oxygen bond is formed at the same time that the 
carbon-chlorine bond is broken. 



Mechanism B: 
HO: e CH,:Cl: 



slow 



Of the two mechanisms, A requires that the reaction rate be determined 
solely by the rate of the first step (cf. earlier discussion, Section 2-5B). This 
means that the rate at which methanol is formed (measured in moles per unit 
volume per unit time) will depend on the concentration of methyl chloride, 
and not on the hydroxide ion concentration, because hydroxide ion is not 
utilized except in a fast secondary reaction. In contrast, mechanism B requires 
the rate to depend on the concentrations of both reagents since the slow step 
involves collisions between hydroxide ions and methyl chloride molecules. 
More precisely, the reaction rate (v) may be expressed in terms of Equation 
8-1 for mechanism A and Equation 8-2 for mechanism B. 

v = k[CH 3 C\] (8-1) 

z; = MCH 3 Cl][OH e ] (8-2) 

Customarily, v is expressed in moles of product formed per liter of solution 
per unit of time (most frequently in seconds). The concentration terms 
[CH 3 C1] and [OH e ] are in units of moles per liter, and the proportionality 
constant k (called the specific-rate constant) has the dimensions of sec" 1 for 
mechanism A and mole -1 x liters x sec -1 for mechanism B. 

It is useful to speak of both the order of a reaction with respect to a specific 
reactant and the overall order of a reaction. The order of a reaction with 
respect to a given reactant is the power to which the concentration of that 
particular reagent must be raised to have direct proportionality between 
concentration and reaction rate. According to Equation 8-2 the rate of the 
methyl chloride-hydroxide ion reaction is first order with respect to both 
reagents. In Equation 8T the rate is first order in methyl chloride; the order 
with respect to hydroxide ion may be said to be zero since [OH e ]° = 1. 
The overall order of reaction is the sum of the orders of the respective re- 
actants. Thus, Equations 8-1 and 8-2 express the rates of first-order and 
second-order reactions, respectively. 



sec 8.9 energetics of S N 1 and S N 2 reactions 197 

We have, then, a kinetic method for distinguishing between the two possible 
mechanisms, A and B, that we are considering. Experimentally, the rate of 
formation of methyl alcohol is found to be proportional to the concentrations 
of both methyl chloride and hydroxide ion. The reaction rate is second order 
overall and is expressed correctly by Equation 8-2. From this we infer that the 
mechanism of the reaction is the single-step bimolecular process B. Reactions 
having this type of mechanism are generally classified as bimolecular nucleo- 
philic substitutions, often designated S N 2 (S for substitution, N for nucleo- 
philic, and 2 for bimolecular). 

On the other hand, the rate of the reaction of ?-butyl chloride with aqueous 
base depends only on the concentration of the chloride. The concentration of 
the base (and its strength) is irrelevant, as long as there is enough to neutralize 
the acid produced in the reaction. (Bicarbonate ion serves just as well as 
hydroxide ion for this purpose.) 

(CH 3 ) 3 CC1 s '° w > (CH 3 ) 3 C e + Cl e 

(CH 3 ) 3 C® + H 2 fast > (CH 3 ) 3 COH 2 B: > (CH 3 ) 3 CQH+ BH ffi 

(or (CH 3 ) 3 C ffl + OH e fast > (CH 3 ) 3 COH) 



Mechanisms such as this are designated S N 1 reactions because they are 
nucleophilic substitutions in which the first-order kinetics suggest that the 
rate-controlling step is unimolecular. That is, the rate is proportional to the 
concentration of alkyl halide and not to the concentration of the base. 

Many S N reactions are carried out using the solvent as the nucleophilic 
agent. They are called solvolysis reactions; specific solvents such as water, 
ethanol, acetic acid, and formic acid produce hydrolysis, ethanolysis, ace- 
tolysis, and formolysis reactions, respectively. The rates of all solvolysis 
reactions are necessarily first order since the solvent is in such great excess that 
its concentration does not change effectively during reaction, and hence its 
contribution to the rate does not change. But this does not mean that reaction 
is necessarily proceeding by an S N 1 mechanism, particularly in solvents such 
as water, alcohols, or amines, which are expected to be reasonably good 
nucleophilic agents. 



8-9 energetics of S N 1 and S N 2 reactions 

We have seen that the S N 2 reaction of methyl chloride with hydroxide is a 
one-step process in which no intermediate compound is formed. As the 
hydroxide ion begins to bond to carbon, the carbon-chlorine bond begins to 
break and one can formulate the course of the reaction as in Equation 8-3. 

u ^e H H H 

HCT \ 8e I Se ± / 

.-CI > [H-O-C-Cir > HO-C + CI (8-3) 



^C-Cl > [H--0~C~CI] + ► HO-C 

H "i "I V-H 



chap 8 nucleophilic displacement and elimination reactions 198 

The transition state in Equation 8-3 which is designated by t is not a 
molecule in the ordinary sense. It is simply a stage in the transition from 
reactants to products. Its importance derives from the fact that it represents 
the point of highest energy along the lowest-energy (most favorable) reaction 
path. There are many routes by which one could imagine the atoms of the 
reactants being rearranged so as to give the products, but most of them would 
involve transition states of exorbitantly high energy. (See mechanism A, 
Section 8-8, that involves formation of a primary carbonium ion, for one 
such route.) The course of any reaction — its mechanism — is that for which the 
highest energy state lying on the reaction path is lower than the lowest energy 
state on all other paths. Figure 8-1 shows the energy profile of the methyl 
chloride-hydroxide ion reaction. 

The transition state is at the top of an energy barrier that must be overcome 
for reaction to occur but, in geographical terms, it should be considered as 
the highest point along the path for the most favorable pass between 
reactants and products. The transition state is not a peak between reactants 
and products. 

The rate of a reaction will be expected to be determined by the height of the 
energy barrier to be overcome — that is, the energy difference between reactants 
and transition state. The energy of the final products is not relevant to the rate, 
although AG for the overall reaction must be negative for the equilibrium 
constant to be greater than unity (Section 2-5A). 

In the so-called "theory of absolute reaction rates," the reaction rate is 
determined by the free-energy difference between reactants and transition 



Figure 8-1 Energy profile for the reaction of a molecule of methyl chloride 
and a hydroxide ion. The reaction coordinate indicates the extent of reaction 
and might be taken as the difference in the CI to C and C to OH distances, 
thus, in principle, covering the range from — oo to + oo. 









H 

HO c a 




>> 

so 

a 










ch 3 ci - 


HOH e \ 




1 CH 3 OH + Cl e 






reaction coordinate 



sec 8.9 energetics of S N 1 and S N 2 reactions 199 



r S® 5e~| 

Lf-Bu CI J +0H 




f-BuOH + CI ' 



reaction coordinate 



Figure 8-2 Energy profile for reaction of t-butyl chloride, (CH 3 ) 3 C1, with 
hydroxide ion. 



state (AG*). This quantity, like AG, can be equated to a heat term and an 
entropy term: 

AG = AH-TAS 
AG t = AH t -TAS t 

The heat of activation, AH f , is usually the dominant term and can be thought 
of as representing the thermal energy a colliding pair of reactants must 
possess (in excess of the average) to reach the transition state. As the temper- 
ature is raised, more and more collisions will be between reactants having 
sufficient thermal energy to attain the transition state and the reaction rate 
will increase. Experimental values for AH f are, in fact, determined by mea- 
suring the effect of temperature on the reaction rate. 

The entropy of activation, AS*, is related to the difference in the degree of 
vibrational, rotational, and translational freedom of the reactants and the 
transition state. (Compare this with the description given previously for AS, 
in Section 2- 5 A.) The more freedom that the transition state possesses relative 
to the reactants, the more positive will be AS 1 and the greater the reaction rate. 
Conversely, the less freedom in the transition state relative to reactants, the 
more negative AS* will be and the slower the reaction rate. In simple terms, 
a transition state that is loose, with substantial freedom in the locations for 
the constituent atoms, will be favored over one in which a high degree of 
organization is required. 

The energy profile for the S N 1 reaction of ?-butyl chloride with hydroxide 
ion is given in Figure 8-2. This profile shows two transition states. The first 
leads to the formation of a discrete intermediate, (CH 3 ) 3 C®; the second, with 
a low barrier, corresponds to the very rapid reaction of the intermediate 



chap 8 nucleophilic displacement and elimination reactions 200 

carbonium ion with hydroxide ion (or water) to form the products. The overall 
reaction is believed to be unimolecular because the rate is independent of the 
concentration of hydroxide ion. Hydroxide ion is involved only in a fast 
following step, not in the initial slow ionization (see Section 8-8). Hydroxide 
ion is shown as an ingredient of the first transition state in the diagram only 
for the purpose of preserving the correct stoichiometry. It takes no part in the 
reaction until the critical ionization stage has been passed. 

In Section 4-4E we discussed reaction rates in terms of energies of possible 
intermediates that might be formed. A more rigorous procedure is to consider 
not AG for formation of the intermediate but AG* for formation of the tran- 
sition state. In the case of the ionization of /-butyl chloride, this does not make 
much difference because the energy of the /-butyl cation formed as the inter- 
mediate is not likely to be greatly different from that of the transition state 
that precedes it. How can we expect this ? Largely on the basis of intuition 
(tempered by experience) that an alkyl carbonium ion is likely to react very 
rapidly with chloride ion to form /-butyl chloride. The reaction between the 
more basic OH e and t-Bu® will be expected to have a still smaller AG % , and 
this is reflected in Figure 8-2 by the smaller height of the second barrier as 
compared to the first. The anion, SbF 6 e , of the super acid, HSbF 6 , (Section 
1-2C), has such a small tendency to react with t-Bu® that it is possible to pre- 
pare t-Bu® SbF 6 e and determine the nmr spectrum of the cation. 



8-10 stereochemistry of S N 2 displacements 

If we pause to consider the S N 2 reaction of methyl chloride with hydroxide 
ion in more detail, we can think of two simple ways in which the reaction 



Figure 8-3 Back-side (inverting) and front-side (noninverting) attack of 
hydroxide ion on methyl chloride, as visualized with ball-and-stick models. 



[ f 









sec 8.1 1 structural and solvent effects in S N reactions 201 

could be effected; these differ in the direction of approach of the reagents, 
one to the other (see Figure 8-3). The hydroxide ion might attack methyl 
chloride directly at the site where the chlorine is attached (i.e., front-side 
approach). Alternatively, hydroxide might approach the molecule from the 
rear to cause expulsion of chloride ion from the front (i.e., back-side approach). 

There is no simple way of proving which of these paths is followed in this 
particular case, but the arguments in favor of the back-side approach are 
very strong. First, the transition state for this approach will have the oxygen 
and chlorine atoms well separated; that is, the charge in the transition state 
will be dispersed. The front-side approach would not disperse the negative 
charge to nearly the same extent and hence would be a less favorable arrange- 
ment. Second, in the case of cyclic compounds, the two types of displacement 
predict different products. For example, an S N 2 reaction between cis-3- 
chloro-1-methylcyclopentane and hydroxide ion would give the cis alcohol 
by front-side attack but the trans alcohol by back-side attack (Figure 8-4). 
The actual product is the trans alcohol, from which we infer that reaction 
occurs by back-side displacement. 

Third, in the case of certain kinds of stereoisomers of open-chain com- 
pounds, back-side displacement has been conclusively proven. This is dis- 
cussed later in the book (Section 1 4-9). 



8-11 structural and solvent effects in S N reactions 

We shall consider first the relation between the structures of alkyl derivatives 
and their reaction rates toward a given nucleophile. This will be followed by 
a discussion of the relative reactivities of various nucleophiles toward a 
given alkyl derivative. Finally, we shall comment on the role of the solvent 
in S N reactions. 



Figure 8-4 Back-side and front-side displacement paths for reaction of 
cis- 3-chloro-l-methylcyclopentane with hydroxide ion. 



back-side 

— 1 

displacement 




+ OH 



front- side 
displacement 




OH 



H 



trans alcohol 



H,C 




OH 



cis alcohol 
(not formed) 



chap 8 nucleophilic displacement and elimination reactions 202 
A. STRUCTURE OF THE ALKYL GROUP, R 

The rates of S N 2 displacement reactions of simple alkyl derivatives, RX, 
follow the order primary R > secondary R > tertiary R. In practical syntheses 
involving S N 2 reactions, the primary compounds generally work very well, 
secondary isomers are fair, and the tertiary isomers are completely impractical. 
Steric hindrance appears to be particularly important in determining S N 2- 
reaction rates, and the slowness of tertiary halides is best accounted for by 
steric hindrance to the back-side approach of an attacking nucleophile by the 
alkyl groups on the a carbon. Neopentyl halides, which are primary halides, 
are very unreactive in S N 2 reactions, and scale models indicate this to be 
the result of their steric hindrance by the methyl groups on the /? carbon. 



CH 3 -C-CH 2 Br 
I 
CH 3 

neopentyl bromide 
(slow in S N 2-type reactions) 

In complete contrast to S N 2 reactions, the rates of S N 1 reactions of alkyl 
derivatives follow the order tertiary R > secondary R > primary R. 

Steric hindrance is relatively unimportant in S N 1 reactions because attack 
by the nucleophile is not involved in the rate-determining step. In fact, steric 
acceleration is possible in the solvolysis of highly branched alkyl halides 
through relief of steric compression by formation of a planar cation : 

CH 3 |ch 3 CH 3 ch 3 

I I _ ve I / 

CH 3 — C — C— X > CH 3 -C — C® 

lil I \ 

CH 3 CH 3 CH 3 \^CH 3 

■*. . steric relief of 

crowding strain 

The reactivity sequence tertiary > secondary > primary is to be expected 
since we know that electron-deficient centers are stabilized more by alkyl 
groups than by hydrogen. The reason for this is that alkyl groups are less 
electron attracting than hydrogen. 



B. THE LEAVING GROUP, X 

The reactivity of a given alkyl derivative, RX, in either S N 1 or S N 2 reactions 
is determined in part by the nature of the leaving group, X. In general, there 
is a reasonable correlation between the reactivity of RX and the acid strength 
of H— X, the X groups that correspond to the strongest acids being the best 
leaving groups. Thus, since H— F is a relatively weak acid and H— I is a very 
strong acid, the usual order of reactivity of alkyl halides is R— I > R— Br 
> R— CI >R— F. Also, the greater ease of breaking a C— OS0 2 C 6 H 5 bond 
than a C— CI bond in S N 2 reactions on carbon correlates with the greater 
acid strength of HOS0 2 C 6 H 5 in relation to HC1. A further factor influencing 



sec 8.11 structural and solvent effects in S N reactions 203 

the rate of nucleophilic displacements is the polarizabilities of the attacking 
and leaving groups. A highly polarizable atom is one whose electron cloud 
can be easily deformed by an electric field, such as will be produced by ions in 
solutions. Polarizability increases as one goes down a group in the Periodic 
Table, and this means that iodide is not only more easily displaced than the 
other halogens but is itself a more reactive nucleophile. In a similar way, 
sulfur compounds react faster than the analogous oxygen compounds. 

Alcohols are particularly unreactive in S N reactions, unless a strong acid is 
present as a catalyst. The reason is that the OH e group is a very poor leaving 
group. The acid functions by donating a proton to the oxygen of the alcohol, 
transforming the hydroxyl function into a better leaving group (H 2 in 
place of OH e ). Reactions of ethers and esters are acid catalyzed for the same 
reasons : 

ROH + Br e — #—> RBr + q H 

H 
R:5:H + H* ,. R:0:H ffi 

: H 
RJ:0:H e + Br e > RBr + H 2 S N 2 

H 
Ri:6:H > R® + H 2 Br6 > RBr S„l 



Heavy-metal salts, particularly those of silver, mercury, and copper, catalyze 
S N 1 reactions of alkyl halides in much the same way as acids catalyze the S N 
reactions of alcohols. The heavy-metal ion functions by complexing with the 
unshared electrons of the halide, making the leaving group a metal halide 
rather than a halide ion. This acceleration of the rates of halide reactions is 
the basis for a qualitative test for alkyl halides with silver nitrate in ethanor 
solution. Silver halide precipitates at a rate that depends upon the structure of 
the alkyl group, tertiary > secondary > primary. Tertiary halides usually 
react immediately at room temperature, whereas primary halides require 
heating. 

r:X: t^^ [R:X:-Ag] ffi -^U R® _^_> RY + H® 

(-AgX) 

There is additional evidence for the formation of complexes between 
organic halides and silver ion: where the formation of carbonium ions is 
slow enough to permit determination of water solubility, the solubility of the 
halide is found to be increased by the presence of silver ion. 



C. THE NUCLEOPHILIC REAGENT 

The S N 2 reactivity of a particular reagent towards an alkyl derivative can be 
defined as its nucleophilicity, which is its ability to donate an electron pair 
to carbon. The nucleophilicity of a reagent does not always parallel its basicity, 
measured by its ability to donate an electron pair to a proton. The lack of 
parallelism can be seen from Table 84, which indicates the range of reactivities 



chap 8 nucleophilic displacement and elimination reactions 204 

Table 8-4 Reactivities of various nucleophiles toward methyl bromide 
in water at 50° 



nucleophile 


approximate 
reaction half-time, hr° 


rate relative 
to water 


Kb 


H 2 


1,100* 


0) 


,0-16 


CH 3 C0 2 & 


2.1 


5.2 x 10 2 


10-" 


Cl e 


1 


1.1 x 10 3 


~io~ 20 


Br e 


0.17 


7.8 x 10 3 


<io- 20 


N 3 e 


0.11 


1.0 x 10* 


10-" 


HO e 


0.07 


1.6 x 10 4 


10° 


C 6 H 5 NH 2 


0.04 


3.1 x 10 4 


10 -io 


SCN e 


0.02 


5.9 x 10 4 


io- 14 


I e 


0.01 


1.1 x 10 5 


<io- 22 



" Time in hours required for half of methyl bromide to react at constant (1 M) concentration of 
nucleophile. 

b Calculated from data for pure water, assuming water to be 55 M, 

of various nucleophilic agents (toward methyl bromide in water) and their 
corresponding basicities. Clearly, a strong base is a good nucleophile (e.g., 
OH e ), but a very weak base may also be a good nucleophile (e.g., I s ) if it is 
highly polarizable. 

D. THE NATURE OF THE SOLVENT 

The rates of most S N 1 reactions are very sensitive to solvent changes. This is 

reasonable because the ionizing power of a solvent is crucial to the ease of 

e e 
formation of the highly ionic transition state R —X from RX. 

Actually, two factors are relevant in regard to the ionizing ability of sol- 
vents. First, a high dielectric constant increases ionizing power by making it 
easier to separate ions, the force between charged particles depending inversely 
upon the dielectric constant of the medium. On this count, water with a 
dielectric constant of 80 should be much more effective than a hydrocarbon 
with a dielectric constant of 2. A related and probably more important factor 
is the ability of the solvent to solvate the separated ions. Cations are most 
effectively solvated by compounds of elements in the first row of the periodic 
table that have unshared electron pairs. Examples are ammonia, water, 
alcohols, carboxylic acids, sulfur dioxide, and dimethyl sulfoxide, (CH 3 ) 2 SO. 
Anions are solvated most efficiently by solvents having hydrogen attached to a 
strongly electronegative element Y so that the H— Y bond is strongly polarized. 
With such solvents, hydrogen bonds between the solvent and the leaving 
group assist ionization in much the same way that silver ion catalyzes ioniz- 
ation of alkyl halides (Section 8*1 IB): 

Se a© ?. se se 
Y-H-:Ci:-H-Y 



H 




H 


\ 


© 


/ 


:<>:- 


--Na- 


-:<>: 


/ 
H 




\ 
H 



solvation of a cation solvation of an anion 

by a solvent with unshared by a hydrogen-bonding 

electron pairs solvent 



sec 8.11 structural and solvent effects in S N reactions 205 

Water appears to strike the best compromise with regard to the structural 
features that make up ionizing power, that is, dielectric constant and solvating 
ability, and we expect f-butyl chloride to hydrolyze more readily in water- 
alcohol mixtures than in ether-alcohol mixtures. An ether can only solvate 
cations effectively whereas water can solvate both anions and cations. 
(However, the water solubility of alkyl halides is too low for pure water to be 
a suitable medium for these reactions.) 

For S N 2 reactions, effects of changing solvents might be expected to be 

smaller because the reactants and the transition state each possess a full unit 

s& so 

of negative charge: HO e + RX -> (HO— R~ X). No charges have been 
created but the charge in the transition state is less concentrated than in 
the reactants. Accordingly, a poor solvating solvent should raise the energy 
of the reactants more than it raises that of the transition state (a large diffuse 
ion is solvated less than a small concentrated one) and hence speed up the reac- 
tion. This hypothesis is not easily tested with solvents such as hexane or carbon 
tetrachloride because they do not dissolve metal hydroxides. We can, however, 
look for solvents with high dielectric constants but which lack hydrogen- 
bonding ability to solvate anions well. There are a number of such solvents 
and the most important of these, together with their dielectric constants, are 
listed here. 

H 2 C-CH 2 O 

o / \ / 

II H 2 C X /CH 2 H-C 

CH3-S-CH3 o / %Q N(CH3)2 

dimethyl sulfoxide tetramethylene sulfone dimethylformamide 

(DMSO) e = 48 (sulfolane) s = 44 (DMF) £ = 38 

O 

(CH 3 ) 2 N-P-N(CH 3 ) 2 

N(CH 3 ) 2 

hexamethylphosphoramide 
(HMP)e = 30 

These solvents, called polar aprotic solvents, have a remarkable effect on 
the rates of many S N 2 reactions. For example, the S N 2 reaction of methyl 
iodide with chloride ion, 

Cl e + CH3I ► [CI-CH3-I] ► CICH3 + 1° 

is a million times faster in dimethylformamide than in water. Of the four 
solvents listed above, DMSO and HMP are usually the most effective in 
accelerating S N 2 reactions. 



elimination reactions 

The reverse of addition to alkene double bonds is elimination. Generally, 
an alkyl derivative will, under appropriate conditions, eliminate HX, where 



chap 8 nucleophilic displacement and elimination reactions 206 

X is commonly a halide, hydroxyl, ester, or onium function, and a hydrogen 
is located on the carbon adjacent to that bearing the X function : 

II \ / 

— C — C— ► C=C + HX 

II / \ 

H X 

O 
X = C1, Br, 1, — O-C-CHj, -SR 2> — NR 3 ,— OH 2 

Substitution and elimination usually proceed concurrently for alkyl 
derivatives and, in synthetic work, it is important to be able to have as much 
control as possible over the proportions of the possible products. As we shall 
see, substitution and elimination have rather closely related mechanisms, a 
fact which makes achievement of control much more difficult than if the 
mechanisms were sufficiently diverse to give very different responses to changes 
in experimental conditions. 



8-12 the E2 reaction 

Consider the reaction of ethyl chloride with sodium hydroxide : 

CH 3 CH 2 OH + Cl e S N 2 
CH 3 CH 2 C1 + OH e 

CH 2 = CH 2 + H 2 + Cl e E2 

Elimination to give ethene competes with substitution to give ethanol. 
Furthermore, the rate of elimination, like the rate of substitution, has been 
found to be proportional to both the concentration of ethyl chloride and the 
concentration of hydroxide ion; thus, elimination is here a bimolecular process, 
appropriately abbreviated as E2. As to its mechanism, the attacking base, 
OH e , removes a proton from the ft carbon simultaneously with the formation 
of the double bond and the loss of chloride ion from the a carbon : 

, . O. 

CH 2 j-CH 2 :Cl= ► H 2 + CH 2 = CH 2 + Cl e 

H:0: 

Structural influences on E2 reactions have been studied extensively. The 
ease of elimination follows the order tertiary R > secondary R > primary R. 
The contrast with S N 2 reactions is strong here because E2 reactions are only 
slightly influenced by steric hindrance and can take place easily with tertiary 



sec 8.13 the El reaction 207 



CH 3 

I 

o , , - CH3-C-OH 

CH 3 Sn2^/-* 3 I 

I e -— -"^ CH 3 

-C— CI + OH 

I . ^\P CH, 



CH, ^---^ ___ / 



1 i 



CH 2 = C 



halides, unlike S N 2 reactions. Rather strong bases are generally required to 

bring about the E2 reaction. The effectiveness of bases parallels base strength, 

e e e e 

and the order NH 2 > OC 2 H 5 > OH > 2 CCH 3 is observed for E2 re- 
actions. This fact is important in planning practical syntheses because the 

E2 reaction tends to predominate with strongly basic, slightly polarizable 

e e 

reagents such as amide ion, NH 2 , or ethoxide ion, OC 2 H 5 . On the other 
hand, S N 2 reactions tend to be favored with weakly basic reagents such as 
iodide ion or acetate ion. Elimination is favored over substitution at elevated 
temperatures. 



8-13 the El reaction 

Many secondary and tertiary halides undergo El type of elimination in com- 
petitition with the S N 1 reaction in neutral or acidic solutions. For example, 
when /-butyl chloride solvolyzes in 80 % aqueous ethanol at 25°, it gives 83 % 
/-butyl alcohol by substitution and 17% 2-methylpropene by elimination: 

CH 3 

I 
CH 3 -C=CH 2 17% 



I 
CI 



El 
80%C 2 H 5 OH 

" ^S N l CH 3 



CH 3 -C-CH 3 83% 
I 
OH 



The ratio of substitution and elimination remains constant throughout the 
reaction, which means that each process has the same kinetic order with 
respect to the concentration of /-butyl halide. Usually, but not always, the 
S N 1 and El reactions have a common rate-determining step — namely, slow 
ionization of the halide. The solvent then has the choice of attacking the 
intermediate carbonium ion at carbon to effect substitution, or at a jS hydrogen 
to effect elimination. 

CH 3 

I e 

CH CH, CH 3 -C=CH 2 + H 3 El 

I 3 I 3 H z O -» 

CH 3 -C-CH 3 -~^ CH 3 -C-CH 3 -J2!L>^ c „ 



*3 Y 3 Slow ' ^"3 ^ ^i 

CI 



2H 2 



I 
OH 



chap 8 nucleophilic displacement and elimination reactions 208 

Structural influences on the El reaction are similar to those for the S N 1 
reactions and, for RX, the rate orders are X = I > Br > CI > F and tertiary 
R > secondary R > primary R. With halides such as f-pentyl chloride, which 
can give different alkenes depending upon the direction of elimination, the 
El reaction tends to favor the most stable, that is the most highly substituted, 
alkene. 

CH 3 



CI 



H 2 1/ 
./slow 



CH, 



CH 3 CH 2 = C— CH 2 -CH 



^^^ 20% + H ,O e +cf 



fast 

CH 

+ I " 

e CH,-C = CH-CH 



CI 



Another feature of El reactions (and also of S N 1 reactions) is the tendency 
of the initially formed carbonium ion to rearrange if by so doing a more stable 
ion results. For example, the very slow S N 1 formolysis of neopentyl iodide 
leads predominantly to 2-methyl-2-butene. Here, ionization results in 



CH 3 


CH 3 


[ 
3-C— CH 2 I 

1 


— ie 1 si 


HC0 2 H ' | 


CH 3 


CH 3 


► CH 3 


-C=CH-CH 3 




CH 3 



-* CH 3 -C — CH 2 — CH 3 
I 
CH, 



migration of a methyl group with its bonding pair of electrons from the ft to 
the a carbon transforming an unstable primary carbonium ion to a relatively 
stable tertiary cation. Elimination of a proton completes the reaction. 

Rearrangements involving shifts of hydrogen (as H: e ) occur with compar- 
able ease if a more stable carbonium ion can be formed thereby. 

H 3 C H CH 3 CH 3 

CH 3 -C-C-CH 3 — ^-* CH 3 -C-CH-CH 3 ► CH 3 -C-CH 2 CH 3 

|| " |_> e © 

H Br H 

CH 3 CH 3 

I I 

CH 3 — C-CH 2 CH 3 CH 3 -C = CHCH 3 

OH 

Rearrangements of this type are also discussed in Chapter 10. 




summary 209 



summary 



Many organic compounds can be considered derivatives of the inorganic com- 
pounds water (alcohols, ethers, carboxylic acids, anhydrides), hydrogen 
sulfide (thiols, thioethers, thioacids), ammonia (amines and amides), nitric 
and nitrous acids (alkyl nitrates and nitrites), sulfuric acid (alkyl sulfates), and 
hydrogen halide (alkyl halides). 

CH 3 
I 

Three methods of naming alcohols can be illustrated using CH 3 — CH— OH 
as an example : 2-propanol (IUPAC name), isopropyl alcohol, and dimethyl- 
carbinol. Ethers, ROR, take their names from the two R groups. Carboxylic 
acids, RC0 2 H, are named either as alkanoic acids (IUPAC system) in which 
the longest chain provides the name and -oic acid is added or, in the case of 
the short-chain acids, with the common names formic, acetic, propionic, or 
butyric (Q to C 4 ). The carboxyl carbon is taken as C-l in the IUPAC 
system and the adjacent carbon is a in the common system. Thus, CH 3 — 
CI 
I 

CH— CH 2 C0 2 H can be named either 3-chlorobutanoic acid or j6-chloro- 
butyric acid. 

Nucleophilic displacement reactions of the following types occur: 

RX + Y 9 ► RY + X s 

RX + Y > RY« + X 9 

RX + HY > RY + HX ( a ^ vol f h is re f cti °" 

if HY is the solvent) 

RZ® + Y ► RY® + Z 

RZ ffi + Y e ► RY + Z 

Common nucleophiles are : 

Y e = Cl e , Br e , I s , HO e , RO e , RC0 2 e , CN e , R e , NH 2 e , N 3 e , N0 2 e 

Y = R 3 N, R 3 P, R 2 S 
HY = H 2 0, ROH, RC0 2 H, NH 3 

Easily displaced groups are : 

X e = Cl e , Br e , I e , RS0 3 e , RC0 2 e 
Z = R 3 N, RI, N 2 

Nucleophilic displacements can occur by either S N 1 or S N 2 mechanisms. 

o 
state 



S„l RX ► [R da3 -X de ] ► R®+ X e two-step reaction, rate a [RX] 

transition 



Y© 

♦ RY 



fast 

S N 2 RX + Y e ► [Y 8e -R-X se ] ► RY + X e one-step reaction, 

transition rate a [RX][Y] 

state 

The course of such reactions can be plotted on an energy-profile diagram with 



chap 8 nucleophilic displacement and elimination reactions 210 

energy as a function of the reaction coordinate. The activation parameters 
AH 1 and AS 1 that govern the rate of the reaction are analogous to the terms 
AH and AS that determine the equilibrium position of a system. The S N 2 
route is favored for primary and the S N 1 route for tertiary alkyl groups. 

In general, strong bases make good nucleophiles and stable molecules or 
ions make good leaving groups. Weak bases and less stable leaving groups 
may also react rapidly if their polarizabilities are high. 

S N 1 reactions are greatly accelerated by highly polar solvents which promote 
ionization. S N 2 reactions that involve attack by an anion are greatly accelerated 
by polar aprotic solvents, such as dimethyl sulfoxide. The energy of the 
nucleophile is raised more than that of the transition state by these solvents, 
which solvate anions poorly. 

Elimination processes are analogous to nucleophilic displacements except 
that the base Y e removes a proton from the adjacent carbon. The same 

RCH 2 CH 2 X + Y e ► RCH=CH 2 + 

variations in structure and charge of X and Y are possible with elimination 
as with displacement and the two kinds of reaction compete with one another. 
Furthermore, there are two mechanisms, El and E2, analogous to S N 1 and 
S N 2. The E2 reaction is favored over S N 2 by the use of (a) powerful bases of 
low polarizability, (b) high temperatures, and (c) tertiary alkyl substrates. 

The El and S N 1 reactions usually have a common first step and so the factors 
that govern their rates are the same. Rearrangement of the intermediate 
cation that is produced in both processes sometimes occurs. 



exercises 

8-1 Name each of the following by an accepted system: 

CH 3 

I 
a. CH 3 -C-CH 2 -CH 2 OH d. BrCH 2 CH 2 OCH=CH 2 

CH 3 CH 3 

'■Ch H 

OH 



CH 3 



c. 



CH 3 -CH-CH-C0 2 CH 3 /■ <f \-C-CH 3 

Br Br NO 



8-2 Write structural formulas for each of the following: 

a. dimethylvinylamine 

b. allyl trimethylacetate 

c. N-methyl-N-ethylformamide 

d. formic acetic anhydride 

e. a-phenylethanol 
/. isoamyl nitrite 



exercises 211 

i-3 Name each of the following alcohols by the IUPAC method: 

a. s-butyl alcohol c. trirnethylcarbinol 

b. allyl alcohol d. isobutyl alcohol 

i-4 Name each of the following carboxylic acids by the IUPAC method: 

a. CH 3 CH 2 CHC0 2 H c. (C 6 H 5 ) 2 CHCH 2 C0 2 H 

I 
CH 3 

b. CH 2 =CHCH 2 CH 2 CH 2 C0 2 H d. C1 3 CCH 2 CH 2 CC1 2 C0 2 H 

•5 Complete the following equations and provide a suitable name for each of the 
organic products. If additional products are likely to be formed, provide 
structures and names for them also. 



H 2 Q 

H 2 

► 

C 2 H 5 OH 



a. (CH 3 ) 2 CHCH 2 Br + KOH 

b. C 6 H 5 CH 2 CH 2 Br+NH 4 I 

c. (CH 3 ) 2 CHCl+NaOC 2 H 5 

d. C 2 H 5 Br + NH 3 H2 ° > 

(~"H OH 

e. (C 6 H 5 ) 3 CCH 2 I + CH 3 OH — ^ ► 

8-6 Write structural formulas for the principal organic products formed by the 
action of each of the following reagents on (C 6 H 5 ) 2 CHC1: sodium cyanide, 
ammonia, potassium ethoxide, sodium acetate, methanol. 

8-7 The numbers 1 and 2 in the symbols S N 1 and S N 2 designate the " molecularity " 
of the rate-controlling step; that is, the number of molecular species that are 
believed to react to form the transition state. This often corresponds to the 
kinetics of the reactions, S N 1 displacements often being first order and S N 2 
displacements often being second order. Under what conditions would the 
molecularity and the observed kinetics not correspond ? 

8-8 How would you expect geometry of the transition state to be related to the 
entropy of activation ? 

8-9 The reaction of alcohols with hydrobromic acid to give alkyl bromides is an 
equilibrium reaction. Alkyl bromides are usually formed from alcohols and 
concentrated hydrobromic acid in good yields, whereas alkyl bromides 
hydrolyze almost completely in neutral water solution. Estimate the change in 
equilibrium ratio of alkyl bromide to alcohol in changing from a solution 
with 10 M bromide ion buffered at pH 7 to 10 M hydrobromic acid. 

8-10 The S N 1 reactions of many RX derivatives that form moderately stable 
carbonium ions are substantially retarded by added X e ions. However, such 
retardation is diminished, at given X e concentrations, by adding another 
nucleophile such as N 3 e . Explain. 

The relative reactivity of water and N 3 e toward methyl bromide is seen from 
Table 8-4 to be 1 : 10,000. Would you expect the relative reactivity of these 
substances toward the f-butyl cation to be larger, smaller, or about the same? 

Why? 



chap 8 nucleophilic displacement and elimination reactions 212 

8-11 Classify the following solvents according to effectiveness expected for solva- 
tion of cations and anions : 

a. acetone d. chloroform 

b. carbon tetrachloride e. trimethylamine, (CH 3 ) 3 N 

c. anhydrous hydrogen fluoride / trimethylamine oxide, 

e e 
(CH 3 ) 3 N-0 

8-12 An alternative mechanism for E2 elimination is the following: 

CH 3 CH 2 Cl+OH e ^z± CH 2 CH 2 C1 + H 2 S '° W » CH 2 =CH 2 +Cr 

a. Would this mechanism lead to first-order kinetics with respect to the 
concentrations of OH e and ethyl chloride? Explain. 

b. This mechanism has been excluded for several halides by carrying out 
the reaction in deuteriated solvents such as D 2 and C2H5OD. Explain 
how such experiments could be relevant to the reaction mechanism. 

a e 
8-13 a. Why is potassium ?-butoxide, KOC(CH 3 ) 3 , an excellent base for 

promoting elimination reactions of alkyl halides, whereas ethylamine, 
CH3CH2NH2, is relatively poor for the same purpose? 
b. Potassium 7-butoxide is many powers of ten more effective as an elim- 
inating agent in dimethyl sulfoxide than in ?-butyl alcohol. Explain. 

8-14 The reaction of ?-butyl chloride with water is strongly accelerated by sodium 
hydroxide. How would the ratio of elimination to substitution products be 
affected thereby? 

8-15 Write equations and mechanisms for all the products that might reasonably 
be expected from the reaction of s-butyl chloride with a solution of potassium 
hydroxide in ethanol. 

8-16 Why is apocamphyl chloride practically inert toward hydroxide ion? 
H 3 C. 




•17 Show how the following conversions may be achieved (specify reagents and 
conditions; note that several steps may be needed). Write a mechanism for 
each reaction you use. (Note that some of the steps required are described in 
earlier chapters.) 

CH 3 CH 3 

I ■ ! 

a. CH 3 -C-CH 2 -CH 3 ► CH 3 -C-CH-CH 3 

I II 

Br HO Br 



exercises 213 



\ 

CH 2 CH 3 



c. CH3-C-CH3 



CH 3 -C = CH-CH 3 
CH, 



CI I 



CH 3 CH 3 



d. CH 3 — C— CH 3 — 

I 1 

CI H 

8-18 Explain how (CH 3 ) 2 CDCHBrCH 3 might be used to determine whether 
trimethylethylene is formed directly from the bromide in an El reaction, or by 
rearrangement and elimination as shown in Section 8-13. 

8-19 Predict the products of the following reactions: 

a. CH 3 CH 2 CBr(CH 3 )CH 2 CH 3 s " 2 ° t > 

b. (CH 3 ) 3 CCH(CH 3 )CI ^f^' 

H 2 C 

c. H 2 C V X \:H-CH 2 Br " 2 ° » 

i \ / 2 S N 1, El 

H 2 C 

8-20 Write structural formulas for each of the following substances: 

a. diisobutyl ether 

b. 2-methyl-3-buten-2-ol 

c. dineopentylcarbinol 

d. <x,/3-dibromopropionic acid 

e. ethyl vinyl ether 

/ 9-(2,6,6-trimethyl-l-cyclohexenyl)-3,7-dimethyl-2,4,6,8-nonatetraen- 
l-ol 

8-21 Name each of the following by the IUPAC system and, where applicable , by 
the carbinol (or substituted acid) system: 

CH 3 
a. HC = C-CH 2 OH d. CH 3 -C-CH 2 C0 2 H 

CH 3 CH 3 



b. CH 3 -C-CH-CH 



CH 3 
H 3 C OH e. CH 3 -CH-CH-CH-C0 2 H 

H 3 C /=\OH CI OH 



H 3 C \==/ H 



chap 8 nucleophilic displacement and elimination reactions 214 

8-22 Indicate how you would synthesize each of the following substances from the 
given organic starting materials and any other necessary organic or inorganic 
reagents. Specify reagents and conditions. 

a. 2-butyne from ethyne 

b. 3-chloropropyl acetate from 3-chloro-l-propene 

c. methyl ethyl ether from ethanol 

d. methyl r-butyl ether from 2-methylpropene 

e. l-iodo-2-chloropropane from propene 

8-23 Which one of the following pairs of compounds would you expect to react 
more readily with (A) potassium iodide in acetone, (B) concentrated sodium 
hydroxide in ethanol, and (C) silver nitrate in aqueous ethanol? Write 
equations for all the reactions involved and give your reasoning with respect 
to the predicted orders of reactivity. 

a. methyl chloride and isobutyl chloride with A, B, and C 

b. methyl chloride and Nbutyl chloride with A, B, and C 

c. 7-butyl chloride and l-fluoro-2-chloro-2-methylpropane with B and C 

d. allyl and allylcarbinyl chlorides with A, B, and C 



1 chapter 9 

[alkyl Jialides and oi^ikh 

metallic compounds 



chap 9 alkyl halides and organometallic compounds 217 

Many simple halogen derivatives of hydrocarbons have been met in earlier 
chapters. Their nomenclature was described in Section 3-1, the mechanism 
of their formation by substitution in alkanes in Sections 2-5A, 2-5B, and 
3-3B, and the details of their reactions with nucleophiles (S N 1, S N 2) and 
bases (El, E2) in the previous chapter. 



9-1 physical properties 



Methyl iodide (iodomethane), CH 3 I, boils at 42° and is the only mono- 
halomethane that is not a gas at room temperature and atmospheric pressure. 
Ethyl bromide (bromoethane), CH 3 CH 2 Br, bp 38°, is the first monobromo- 
alkane in the series tobea liquid and thetwochloropropanes,CH 3 CH 2 CH 2 Cl 
(1-chloropropane), bp 47°, and CH 3 CHC1CH 3 (2-chloropropane), bp 37°, 
are the first monochloroalkanes in the series to be liquids. 

With the exception of the fluoroalkanes, which are discussed in a later sec- 
tion, the boiling points of haloalkanes tend to be near those of the alkanes of 
the same molecular weight. 

All alkyl halides have extremely low water solubilities. 

All iodo-, bromo-, and polychloro-substituted alkanes are denser than 
water. 



9-2 spectra 



The infrared spectra of alkyl halides have relatively few bands that are as- 
sociated directly with the C— X bond. However, C— F bonds give rise to 
very intense absorption bands in the region of 1350 to 1000 cm -1 ; C— CI 
bonds absorb strongly in the region of 800 to 600 cm -1 ; whereas C— Br and 
C — I bonds absorb at still lower frequencies. 

Carbon tetrachloride, CC1 4 , and chloroform, CHC1 3 , are commonly used 
solvents for infrared work, and their spectra are shown in Figure 9-1. Carbon 
tetrachloride contains only one kind of bond and this makes its spectrum 
simpler than that of chloroform. Furthermore, the high degree of symmetry 
in carbon tetrachloride contributes to the simplicity of its spectrum. Absorp- 
tion of a quantum of radiation can only occur if accompanied by a change in 
the polarity of the molecule. In many cases, changes in the vibrational energies 
of highly symmetrical molecules may not result in a change in polarity and 
thus not correspond to observable absorptions in the infrared spectrum. You 
should be able to deduce from this discussion why the absorption correspond- 
ing to changes in the C— H vibrational energy but not the C=C vibrational 
energy can be observed in the infrared spectrum of ethyne, HC=CH. 

The nmr spectra of a number of halogen derivatives of hydrocarbons were 
described in Chapter 7. 

The ultraviolet spectra of monohaloalkanes are unremarkable. Neither 
fluoroalkanes nor chloroalkanes show significant absorption in the accessible 
part of the spectrum. Bromo- and iodoalkanes have weak absorption maxima 
between 2000 A and 2500 A. Conjugation of a halogen atom with a double 



chap 9 alkyl halides and organometallic compounds 218 



4 5 



wavelength, n 
6 7 



9 10 12 14 




3600 3200 


2400 2000 2000 












1800 1600 1400 
frequency, cm" 1 


1200 


1000 


800 




3600 3200 2400 2000 2000 1800 1600 1400 1200 1000 800 

frequency, cm -1 



Figure 9*1 Infrared spectra of carbon tetrachloride, CC1 4 (upper), and chloro- 
form, CHC1 3 (lower); neat liquids, 0.1-mm thickness (look for overtones). 



bond, however, causes significant absorption bands to appear, as does an 
accumulation of iodine on a single carbon, CHI 3 being yellow and CI 4 red. 



9-3 preparation of alkyl halides 

A number of ways of forming a carbon-halogen bond have been outlined 
previously. These are illustrated for the production of the isomeric compounds 
1-bromopropane and 2-bromopropane. 



sec 9.5 vinyl halides 219 

a. Halogenation of alkanes is not usually a satisfactory preparative 



CH 3 CH 2 CH 3 + Br 2 -^ HBr + ^h^ ^ ""Ti 

1 CH 3 CH 2 CH 2 Br (minor product) 



method for bromides unless a tertiary C— H is to be substituted. With 
chlorine, serious mixtures of mono- and polysubstitution products may be 
formed. 
b. Reaction of alcohols with hydrogen halides is satisfactory for most 

CH 3 CH 2 CH 2 OH + HBr ► CH 3 CH 2 CH 2 Br + H 2 

CH,CHOHCH, + HBr ► OHLCHBrCH, + H,0 



bromides and iodides; primary alcohols (RCH 2 OH) react only slowly with 
HC1 unless ZnCl 2 is added (Section 10-5). 
c. Addition of hydrogen halides to alkenes proceeds as follows : 

CH 3 CH=CH 2 + HX > CH 3 CHXCH 3 (X = F,C1, 1) 



CH 3 CH=CH 2 + HBr . 



peroxides^. CHCHCHBr 



9-4 reactions of alkyl halides 



a. Displacement (S N 1 , S N 2). The displacement reactions of alkyl halides with 
nucleophiles were listed in Table 8-2. They can be summarized as follows: 

R— X + Y s ► R— Y + X e Y e = Cl e , Br e , OH e , RO e , RS e , CH 3 C0 2 e , 

CN e , Re, NH 2 e , N 3 e , N0 2 9 

R— X + HY ► R— Y + HX HY = H 2 0, ROH, RC0 2 H, NH 3 

b. Elimination (El, E2). If there is a hydrogen atom on the carbon atom 
adjacent to the C— X group, elimination of HX will compete with the 
displacement reaction : 

RCH 2 CH 2 X + Y e ► RCH=CH 2 + HY + X e 

c. Formation of organometallic compounds from alkyl halides and 
metals is discussed in Section 9-9B. 



9-5 vinyl halides 



The most readily available vinyl halide is vinyl chloride, which can be prepared 
by a number of routes: 



chap 9 alkyl halides and organo metallic compounds 220 



CI 



CH^CH 


+ HC1 

\ 




CH 2 =CH 2 + CI 

/high 
,/ temp. 




CH 2 = 


= CHC1 






ow yii 




\0H 9 
E2 \ 




* CH 2 - CH 2 




CH 3 -CHC1 2 


■h 




CI CI 







The most feasible commercial preparation (though not convenient on a 
laboratory scale) is probably by way of high-temperature chlorination of 
ethene. 

The outstanding chemical characteristic of vinyl halides is their general 
inertness in S N 1 and S N 2 reactions. Thus vinyl chloride, on long heating with 
solutions of silver nitrate in ethanol, gives no silver chloride, fails to react 
with potassium iodide by the S N 2 mechanism, and with sodium hydroxide 
only gives ethene by a slow E2 reaction. The haloalkynes, such as RC=C— CI, 
are not very reactive in S N 1 and S N 2 reactions. 

The phenyl halides, C 6 H 5 X, are like the vinyl and ethynyl halides in being 
unreactive in both S N 1 and S N 2 reactions. The chemistry of these compounds 
is discussed in Chapter 21. 



9-6 allyl halides 



Allyl chloride is made on a commercial scale by the chlorination of propene 
at 400° (1 ,2-dichloropropane is a minor product under the reaction conditions, 
although at room temperature it is essentially the only product obtained). 

CH 2 = CH-CH 3 + Cl 2 400 ° » CH 2 = CH — CH 2 C1 + HC1 

Allyl chloride is an intermediate in the commercial synthesis of glycerol 
(1,2,3-propanetriol) from propene. 

CH 2 =CHCH 3 C ' 2 > CH 2 =CHCH 2 C1 H2 ° > CH 2 =CHCH 2 OH 

HOC1 „„ „„ „„ , „„ „„ „„ H 2 CH CH _ CH 



I I I I I I II 

OH CI OH CI OH OH OH OH OH 

glycerol 

A general method for preparing allylic halides is by addition of halogen 
acids to conjugated dienes, which usually gives a mixture of 1,2 and 1,4 
addition products (see Section 6-2). 

In contrast to the vinyl halides, which are characteristically inert, the allyl 



sec 9.6 allyl halides 221 



CI 

CH 3 -CH-CH=CH 2 + CH 3 -CH=CH-CH 2 C1 

3-chIoro-l-butene l-chloro-2-butene 

(a-methylallyl (y-methylallyl chloride) 

chloride) 

halides are very reactive — in fact, much more reactive than corresponding 
saturated compounds, particularly in S N 1 reactions. Other allylic derivatives 
besides the halides also tend to be unusually reactive in displacement and 
substitution reactions, the double bond providing an activating effect on 
breaking the bond to the functional group. A triple bond has a comparable 
effect and, for example, it is found that the chlorine of 3-chloro-l-propyne is 
quite labile. 



HC = C-CH,C1 



3-chloro- 1 -propyne 
(propargyl chloride) 



The considerable S N 1 reactivity of allyl chloride compared with ra-propyl 
chloride can be explained by reference to the electronic energies of the 
intermediate carbonium ions and starting halides, as shown in Figure 9-2. 
As we have seen previously (Section 6-2), two equivalent electron-pairing 
schemes may be written for the allyl cation, which suggest a stabilized hybrid 
structure with substantially delocalized electrons : 

(one low-energy 

electron-pairing 

scheme) 



I© a i & j- © 1 

CH 2 = CH-CH 2 < ► CH 2 -CH=CH 2 ~ CH 2 -CH-CH 2 J +Cl e 

(two low-energy (hybrid structure) 

electron-pairing 
schemes) 

No such stabilized hybrid structure can be written for the n-propyl cation. 



e 
+ CI 



(one low-energy (one low-energy 

electron-pairing electron-pairing 

scheme) scheme) 



Thus, we see that less energy is required to form the allyl cation from allyl 
chloride than to form the «-propyl cation from n-propyl chloride. (The ease 



chap 9 alkyl halides and organometallic compounds 222 



- CH,CH 2 CH, ffl 



-[CH 2 -CH=CH, 



AH, 



AH., 



CH 2 = CH— CH 2 C1 



propyl 



-CH 3 CH,CH,C1 



Figure 9-2 The high S N 1 reactivity of allyl chloride compared with n-propyl 
chloride is here related to the low energy of allyl-cation formation. 



of reaction is actually determined by the energy differences between the 
starting halides and the transition states. However, the transition states must 
be rather close in energy to the carbonium ions, and it is convenient to deal 
with the latter species when developing the argument; see Section 8-9.) 
Figure 9-2 shows the energetics of the ionization reaction. 



9-7 poljhalogen compounds 

Polychlorination of methane affords the di-, tri-, and tetrachloromethanes 



CH 2 C1 2 

dichloromethane 

(methylene chloride) 

bp40° 



CHC1 3 

trichioromethane 

(chloroform) 

bp61° 



ecu 

tetrachloromethane 

carbon tetrachloride) 

bp77° 



cheaply and efficiently. These substances have excellent solvent properties 
for nonpolar and slightly polar substances. Chloroform was once widely 
used as an inhalation anesthetic but has a deleterious effect on the heart and 
is slowly oxidized by atmospheric oxygen to highly toxic phosgene (COCl 2 ). 
Commercial chloroform contains about 1 % ethanol to destroy any phosgene 
formed by oxidation. 

Carbon tetrachloride is very commonly employed as a cleaning solvent, 
although its high toxicity entails some hazard in indiscriminate use. Carbon 
tetrachloride was once widely used as a fire extinguishing fluid for petroleum 
fires, although its tendency to phosgene formation makes it undesirable for 
confined areas. The common laboratory practice of removing traces of water 
from solvents with metallic sodium should never be applied to halogenated 
compounds. Carbon tetrachloride-sodium mixtures can detonate and are 
shock sensitive. 



sec 9.7 polyhalogen compounds 223 

Trichloroethylene (" Triclene," bp 87°) is a widely used dry-cleaning solvent. 
It may be prepared from either ethene or ethyne. 



Ca(OH) 2 
HC^CH + 2C1 2 ► CHC1 2 -CHC1 2 — et-\ H CI 

\ / 

c=c 

300° y / \ 



CH 2 = CH 2 



3 CI 



2 (-3HC1) 



CI 



CI 



Methylene chloride reacts with hydroxide ion by an S N 2 mechanism very 
much less readily than does methyl chloride. The chloromethanol formed 
then undergoes a rapid E2 elimination to give formaldehyde, a substance that 
exists in water largely as dihydroxymethane (formaldehyde hydrate). 



CH,C1, 



e OH 

slow 
S N 2 



OtH 
CH 2 
CI 



l<c, 



OH e 

fast 

E2 



-► H 2 C = 



H 2 Q 

fast 



OH 
/ 
CH 2 
\ 
OH 



Carbon tetrachloride is even less reactive than methylene chloride. One might 
expect chloroform to be intermediate in reactivity between methylene chloride 
and carbon tetrachloride, but chloroform is surprisingly reactive toward 
hydroxide ion and ultimately gives carbon monoxide, formate, and chlo- 
ride ions. We may then infer that a different reaction mechanism is involved. 
Apparently, a strong base, such as hydroxide ion, attacks the chloroform 
molecule much more rapidly at hydrogen than at carbon. There is strong 
evidence to show that the carbanion so formed, Cl 3 C: e , can eliminate 
chloride ion to give a highly reactive intermediate of bivalent carbon, :CC1 2 , 
called dichloromethylene, a carbene (Section 2-5C). This intermediate has 
only six valence electrons around carbon (two covalent bonds), and although 
it is electrically neutral it is powerfully electrophilic, and rapidly attacks the 
solvent to give the final products. 



:CC1, 



e 
H + OH 



e 
C1,C: 



slow 



H 2 



fast 
2 H 2 



CUC: 



■> :CC1 2 + cr 



♦ CO + 2 HCI 



? 

fast * HC v + 2 HC1 
O e 



Note the analogy between this mechanism for the hydrolysis of chloroform 
and the elimination mechanism of Exercise 8T2. Both reactions involve a 
carbanion intermediate, but subsequent elimination from a /? carbon leads to 
an alkene, and from an a carbon to a carbene. Carbene formation is the 
result of 1,1 or a elimination. 

The electrophilic nature of dichloromethylene, :CC1 2 , and other carbenes, 
including the simplest carbene (:C^ called methylene), can be profitably 



chap 9 alkyl halides and organometallic compounds 224 

used in synthetic reactions. Alkene double bonds can provide electrons, 
and carbenes react with an alkene by cis addition to the double bond to 
give cyclopropane derivatives, by what can be characterized as a cis-\,\ 
cycloaddition to the double bond. Activated carbenes, such as are formed 




/CI 



+ :CC1 2 ► [ j6c 




from the light-induced decomposition of diazomethane (CH 2 N 2 ), even 
react with the electrons of a carbon-hydrogen bond to "insert" the carbon 
of the carbene between carbon and hydrogen. This transforms C— H to 
C-CH 3 . 

I I 

— C:H + =CH 2 ► — C— CH 2 -H 

The high activity of the carbene formed from diazomethane and light is 
because the :CH 2 is formed in an excited electronic state. The absorption 
of the photon not only cleaves the carbon-nitrogen bond but leaves the frag- 
ments in high-energy states. The :CH 2 generated this way is one of the most 
reactive reagents known in organic chemistry. More selective carbene-type 
reactions are possible by elimination of zinc iodide from iodomethylzinc 
iodide, ICH 2 ZnI, which leads only to cyclopropane formation with simple 
alkenes. 



+ CH 2 I, + Zn (CU) > I |;CH, + Znl 2 



9-8 Jluor incited alkanes 

A. FLUOROCHLOROMETHANES 

Replacement of either one or two of the chlorines of carbon tetrachloride 
by fluorine can be readily achieved with the aid of antimony triftuoride 
containing some antimony pentachloride. The reaction stops after two 
chlorines have been replaced. The antimony trifluoride may be regenerated 
continuously from the antimony chloride by addition of anhydrous hydrogen 
fluoride. 

3 CC1 4 + SbF 3 SbC ' 5 > 3 CFCI3 + SbCl 3 
bp25° 

3 CC1 4 + 2 SbF 3 -^-> 3 CF 2 CI 2 + 2 SbCI 3 

bp -30° 

Both products have considerable utility as refrigerants, particularly for 
household refrigerators and air-conditioning units, under the trade name 
Freon. Difiuorodichloromethane (Freon 12) is also employed as a propellant 



sec 9.8 fluorinated alkanes 225 

in aerosol bombs, shaving-cream dispensers, and other such containers. It is 
nontoxic, odorless, and noninflammable, and will not react with hot con- 
centrated mineral acids or metallic sodium. This lack of reactivity is quite 
generally characteristic of the difluoromethylene group, provided the fluorines 
are not located on an unsaturated carbon. Attachment of fluorine to a carbon 
atom carrying one or more chlorines tends greatly to reduce the reactivity of 
the chlorines toward almost all types of reagents. 



B. FLUOROCARBONS 

Plastics and lubricating compounds of unusual chemical and thermal stability 
are required for many applications in the atomic energy and space programs. 
As one example, extraordinary chemical resistance is needed for the pumping 
apparatus used for separating U 235 from U 238 by diffusion of very corrosive 
uranium hexafluoride through porous barriers. The use of substances made 
of only carbon and fluorine (fluorocarbons) for lubricants, gaskets, pro- 
tective coatings, and so on, for such equipment is suggested by the chemical 
resistance of the — CF 2 — group, and considerable effort has been spent on 
methods of preparing compounds such as -f-CF 2 -)- n . 

Direct fluorination is highly exothermic and exceedingly difficult to control, 
but an indirect hydrocarbon-fluorination process, using cobalt trifluoride as 
a fluorinating intermediate, works quite well. 

The radical-catalyzed polymerization of tetrafluoroethene produces the 
polymer called Teflon. 

«CF 2 =CF 2 R ' > -fCF 2 -CF 2 ^ 

Teflon is a solid, very chemically inert substance, which is stable to around 
300°. It makes excellent electrical insulation and gasket materials. It also has 
self-lubricating properties, which are exploited in the preparation of low- 
adhesion surfaces and light-duty bearing surfaces. 

Tetrafluoroethene can be made on a commercial scale by the following 
route : 

3 CHC1, + 2 SbF 3 SbC ' 5 > 3 CHC1F 2 + 2 SbCl 3 

2 CHC1F 2 7 °°T 90 J CF 2 =CF 2 + 2 HC1 
2 90% yield z l 

Radical polymerization of chlorotrifluoroethene gives a useful polymer 
(Kel-F) that is similar to polytetrafluoroethene (Teflon). 

An excellent elastomer of high chemical resistance (Viton) can be made by 
copolymerizing hexafluoropropene with 1,1-difluoroethene. The product is 
stable to 300° and is not attacked by red fuming nitric acid. 



C. PROPERTIES OF FLUOROCARBONS 

The fluorocarbons have extraordinarily low boiling points relative to the 
hydrocarbons of comparable molecular weights and, as seen in Figure 9-3, 



chap 9 alkyl halides and organometallic compounds 226 



7nn 


, 
















100 



100 














ulkanes^* ■*"* __ „-— 








4-*£Z ,000 \)uou\ l \k-jiw>> 








•** 


*r 
















// 





















-200' 



Figure 9-3 Boiling points of straight-chain fluorocarbons (C„F 2n+2 ) and 
hydrocarbons (C„H 2 „ + 2 ). 



the boiling points of fluorocarbons with the same number of carbons and 
about 3.5 times the molecular weight are nearly the same or even lower than 
those of the corresponding alkanes. Octafluorocyclobutane boils 17° lower 
than cyclobutane, despite a molecular weight more than three times as great. 
The high chemical stability, nontoxicity, and low boiling point of octafiuoro- 
cyclobutane make it of wide potential use as a propellant in the pressure 



H 2 C- 

I 



CH 2 

I 



bp + 12° 
mol. wt. = 56 



F 2 C— CF 2 

I I 
F 2 C-CF 2 

bp -5° 
mol. wt.=200 



packaging of food. Fluorocarbons are very insoluble in most polar solvents 
and are only slightly soluble in alkanes in the kerosene range. The higher- 
molecular-weight fluorocarbons are not even miscible in all proportions with 
their lower-molecular-weight homologs. 

The physiological properties of organofluorine compounds vary exception- 
ally widely. Dichlorodifluoromethane and the saturated fluorocarbons appear 
to be completely nontoxic. On the other hand, perfluoro-2-methylpropene 
is exceedingly toxic, more so than phosgene (COCl 2 ), which was used as a 
toxicant in World War I. Many phosphorus-containing organic compounds are 
highly toxic if they also have a P— F group. The so-called "nerve gases" are 
of this type. Sodium fluoroacetate (CH 2 FC0 2 Na) and 2-fluoroethanol are 
toxic fluorine derivatives of oxygen-containing organic substances. The 
fluoroacetate salt is sold commercially as a rodenticide. Interestingly, sodium 
trifluoroacetate is nontoxic. 



9-9 



oraanome 



r 9< 



talli 



ic compoun 



poi 



ids 



Research on the chemistry of organometallic compounds has progressed 
rapidly in recent years. A number of magnesium, aluminum, and lithium 



sec 9.9 organometallic compounds 227 

organometallics are now commercially available, and are used on a large 
scale despite their being extremely reactive to water, oxygen, and almost all 
organic solvents other than hydrocarbons or ethers. This high degree of 
reactivity is one reason for the interest in organometallic chemistry, because 
compounds with high reactivity generally enter into a wide variety of reactions 
and are therefore of value in synthetic work. 

Organometallic compounds are most simply defined as substances posses- 
sing carbon-metal bonds. This definition excludes substances such as sodium 
acetate and sodium methoxide, since these are best regarded as having oxygen- 
metal bonds. Among the common metallic elements that form important 
organic derivatives are lithium, sodium, potassium, magnesium, aluminum, 
cadmium, iron, and mercury. 

Less typically metallic elements (the metalloids) — boron, silicon, germa- 
nium, selenium, arsenic, and so on — also form organic derivatives, some of 
which are quite important, but these fall between true metallic and nonmetallic 
organic compounds. They are best considered separately and will not be 
included in the present discussion. 



A. GENERAL PROPERTIES OF ORGANOMETALLIC COMPOUNDS 

The physical and chemical properties of organometallic compounds vary over 
an extraordinarily wide range and can be well correlated with the degree of 
ionic character of the carbon-metal bonds present. This varies from sub- 
stantially ionic, in the case of sodium acetylide, CH=C: e Na®, to essentially 
covalent as in tetraethyllead, (C 2 H 5 ) 4 Pb. The more electropositive the 
metal, the more ionic is the carbon-metal bond, with carbon at the negative 
end of the dipole. 

Sel se 
— C: Metal 

I 

The reactivity of organometallic compounds increases with the ionic 
character of the carbon-metal bond. It is not then surprising that organo- 
sodium and organopotassium compounds are among the most reactive 
organometallics. They are spontaneously inflammable in air, react violently 
with water and carbon dioxide, and, as might be expected from their saltlike 
character, are nonvolatile and do not readily dissolve in nonpolar solvents. 
In contrast, the more covalent compounds such as organomercurials [e.g., 
(CH 3 ) 2 Hg] are far less reactive; they are relatively stable in air, much more 
volatile, and will dissolve in nonpolar solvents. 

For many organometallic compounds the metal atom does not formally 
have a full shell of valence electrons. Thus, the usual formulation of trimethyl- 
aluminum will have the aluminum with six electrons in its outer valence shell. 
There is a tendency for such compounds to form relatively loose dimers, or 
more complex structures, to give the metal more nearly complete shells in a 
manner discussed earlier for BH 3 which forms B 2 H 6 (Section 4-4B). 



chap 9 alkyl halides and organometallic compounds 228 
H 3 



CH 3 H 3 C x ,C. /CH3 

.Al. /( ,- A1 x 

CH 3 " " CH 3 H,C V CH 3 

H 3 

(monomer) (dimer) 

trimethylaluminum 



B. PREPARATION OF ORGANOMETALLIC COMPOUNDS 

Metals with Organic Halides. The reaction of a metal with an organic halide 
is a convenient method for preparation of organometallics derived from rea- 
sonably active metals such as lithium, magnesium, and zinc. Dialkyl ethers, 
particularly diethyl ether, provide an inert, slightly polar medium with 
unshared electron pairs on oxygen in which organometallic compounds are 
usually soluble. Care is necessary to exclude moisture, oxygen, and carbon 
dioxide, which would otherwise react with the organometallic compound, 
and this is usually done by using an inert atmosphere of nitrogen or helium. 

(CH 3 CH 2 ) 2 
CH 3 Br + 2 Li ► CH 3 Li + LiBr 

methyllithium 

CH 3 CH 2 Br + Mg HiiHkM^ CH 3 CH 2 MgBr 

ethylmagnesium 
bromide 

The reactivity order of thevarioushalides is I > Br > CI > > F. Alkyl fluorides 
do not react with lithium or magnesium. Concerning the metal, zinc reacts 
well with bromides and iodides, whereas mercury is satisfactory only if 
amalgamated with sodium. 

2 CH 3 I + Hg(Na) >• (CH 3 ) 2 Hg + 2 Nal 

Sodium presents a special problem because of the high reactivity of organo- 
sodium compounds toward ether and organic halides. Both lithium and 
sodium alkyls attack diethyl ether but, whereas the lithium compounds usually 
react slowly, the sodium compounds react so rapidly as to make diethyl ether 
impractical as a solvent for the preparation of most organosodium com- 
pounds. Hydrocarbon solvents are usually necessary. Even so, special pre- 
parative techniques are necessary to avoid having organosodium compounds 
react with the organic halide as fast as formed to give hydrocarbons by either 
S N 2 displacement or E2 elimination, depending on whether the sodium de- 
rivative attacks carbon or a fi hydrogen of the halide. 



Sn2 displacement: 



CH 3 CH 2 :'Na ffi + CH 3 CH, 
E2 elimination 



•Br ► CH 3 CH 2 CH 2 CH 3 + Na ffi Br 6 



CH 3 CH,:Na e + H-CHXH,:Br ► CH 3 CH 3 + CH 2 =CH 2 + Na® Br e 



sec 9.9 organometallic compounds 229 

Displacement reactions of this kind brought about by sodium and organic 
halides (often called Wurtz coupling reactions) are only of limited synthetic 
importance. 

Organometallic Compounds with Metallic Halides. The less reactive 
organometallic compounds are best prepared from organomagnesium halides 
(Grignard reagents) and metallic halides. 

CH 3 MgCl + HgCl 2 < — > CH 3 HgCl + MgCl 2 

2CH 3 MgCl + HgCl 2 < — ' (CH 3 ) 2 Hg + 2 MgCl 2 

These reactions, which are reversible, actually go so as to have the most 
electropositive metal ending up combined with halogen. On this basis, sodium 
chloride can be predicted confidently not to react with dimethylmercury to 
yield methylsodium and mercuric chloride. 

Organometallic Compounds and Acidic Hydrocarbons. A few organo- 
metallics are most conveniently prepared by the reaction of an alkylmetal 
derivative with an acidic hydrocarbon such as an alkyne or cyclopentadiene. 

CH 3 MgBr + CH 3 C=CH ► CH 4 + CH 3 C = CMgBr 

Such reactions may be regarded as reactions of the salt of a weak acid 
(methane, K K < 10~ 40 ) with a stronger acid (propyne, K A ~ 10~ 22 ). 

The more reactive organometallic compounds are seldom isolated from the 
solutions in which they are prepared. These solutions are not themselves 
generally stored for any length of time but are used directly in subsequent 
reactions. However, ether solutions of certain organomagnesium halides 
(phenyl-, methyl-, and ethylmagnesium halides) are obtainable commercially; 
also, «-butyllithium is available dissolved in mineral oil and in paraffin wax. 
Manipulation of any organometallic compounds should always be carried 
out with caution, owing to their extreme reactivity, and, in many cases, their 
considerable toxicity (particularly organic compounds of mercury, lead, and 
zinc). 



C. ORGANOMAGNESIUM COMPOUNDS (GRIGNARD REAGENTS) 

The most important organometallic compounds for synthetic purposes are 
the organomagnesium halides, or Grignard reagents. They are so named 
after Victor Grignard, who discovered them and developed their use as 
synthetic reagents, for which he received a Nobel Prize in 1912. As already 
mentioned, these substances are customarily prepared in dry ether solution 
from magnesium turnings and an organic halide. Chlorides often react 

ether _ 

CH 3 I + Mg > CH 3 MgI 

95% yield 

sluggishly. In addition, they may give an unwelcome precipitate of magnesium 



chap 9 alkyl halides and organometallic compounds 230 

chloride which, unlike magnesium bromide and iodide, is only very slightly 

soluble in ether. Very few organomagnesium fluorides are known. 

Organomagnesium compounds, such as methylmagnesium iodide, are not 

e ®> 
well expressed by formulas such as CH 3 MgI or CH 3 MgI because they 

appear to possess polar rather than purely covalent or ionic carbon-magnesium 

bonds. However, the reactions of Grignard reagents may often be conveniently 

considered as involving the carbanion, R e . 

se 8© © 

R-Mg-X « -* R: e + MgX 

The state of the MgX bond has not been specified here because it is not 
usually significant to the course of the reaction. The MgX bond may well have 

as much or more polar character than the RMg bond but our policy in this 

ae s® <5 e s& 

book will be to write structures such as R— MgX or RMg— X only when we 
feel that this will contribute something to understanding the reaction. Thus, 
CH 3 MgI is usually to be understood as the composition of a substance, 
rather than a depiction of a structure, in the same way as we use the formulas 
NaClandH 2 S0 4 . 

Grignard reagents as prepared in ether solution are very highly associated 
with the solvent. Not all the ether can be removed, even under reduced 
pressure at moderate temperatures, and the solid contains one or more 
moles of ether for every mole of organomagnesium compound. The ether 
molecules appear to be coordinated through the unshared electron pairs of 
oxygen to magnesium. 

Reaction with Active Hydrogen Compounds. Grignard reagents react with 
acids, even very weak acids such as water, alcohols, alkynes, and primary 
and secondary amines. These reactions may be regarded as involving the 
neutralization of a strong base (R: e of RMgX). The products are hydro- 
carbon, RH, and a magnesium salt : 

se §© © © 

CH 3 -MgI + CH 3 CH 2 OH <■ CH 4 + CH 3 CH 2 Mgl 

This type of reaction occasionally provides a useful way of replacing a halogen 
bound to carbon by hydrogen as in a published synthesis of cyclobutane from 
cyclo butyl bromide: 

H 2 C-CH-Br _^ H 2 C-CH-MgBr ^ 



H 2 C-CH 2 

Reaction with Oxygen, Sulfur, and Halogens. Grignard reagents react 
with oxygen, sulfur, and halogens to form substances containing C— O, 
C— S, and C— X bonds, respectively. These reactions are not usually impor- 

RMgX + 2 > R-O-O-MgX RMgX > 2ROMgX ^2Jl!t 2 ROH 

H 2 0, H ffi 
8 RMgX + S 8 > 8 RSMgX ► 8 RSH 

RMgX + I 2 ► Rl + MgXI 



sec 9.9 organometallic compounds 231 

tant for synthetic work since the products ROH, RSH, and RX can usually 
be obtained more conveniently and directly from alkyl halides by S N 1 and 
S N 2 displacement reactions, as described in Chapter 8. However, when both 
S N 1 and S N 2 reactions are slow or otherwise impractical, as for neopentyl 
derivatives, the Grignard reactions can be very useful. 

CH 3 CH 3 CH 3 

' Ma I I, I 

CH 3 -C-CH 2 C1 *-> CH 3 -C-CH 2 MgCl —* CH 3 -C-CH 2 I 

CH 3 CH 3 CH 3 

neopentyl chloride neopentyl iodide 

Also, oxygenation of a Grignard reagent at low temperatures provides an 
excellent method for the synthesis of hydroperoxides. To prevent formation 

RMgX + 2 ~ 7 ° » ROOMgX Hi6 > ROOH 

of excessive amounts of the alcohol, inverse addition is desirable (i.e., a solution 
of Grignard reagent is added to ether through which oxygen is bubbled rather 
than have the oxygen bubble through a solution of the Grignard reagent). 

Additions to Carbonyl Groups. The most important synthetic use of 
Grignard reagents is for formation of new carbon-carbon bonds by addition 
to multiple bonds, particularly carbonyl bonds. (Carbon-carbon double and 
triple bonds, being nonpolar, are inert to Grignard reagents.) In each case, 
magnesium is transferred from carbon to a more electronegative element. An 
example is the addition of methylmagnesium iodide to formaldehyde. The 

CH 3 :MgI + H 2 C = O v ► CH 3 :CH 2 -0 Mgl " 2 ° » CH 3 CH 2 OH 

new carbon-carbon bond 

yields of addition products are generally high in these reactions and, with 
suitable variations of the carbonyl compound, a wide range of compounds can 
be built up from substances containing fewer carbon atoms per molecule. 
The products formed from a number of types of carbonyl compounds with 
Grignard reagents are listed in Table 9T. (The nomenclature of carbonyl 
compounds is considered in Section 11-1.) 

The products are complex magnesium salts from which the desired organic 
product is freed by acid hydrolysis : 

ROMgX + HOH ► ROH + HOMgX 

HC1 



■> H 2 + MgXCl 

If the product is sensitive to strong acids, the hydrolysis may be conveniently 
carried out with a saturated solution of ammonium chloride ; basic magnesium 
salts precipitate while the organic product remains in ether solution. 
The reaction of carbon dioxide with Grignard reagents gives initially 



\ 



chap 9 alkyi halides and organometallic compounds 232 



Table 94 Products from the reaction of Grignard reagents as RMgX 
with carbonyl compounds 



product 



hydrolysis product customary yield 



formaldehyde 



aldehyde 



ketone 



carbon dioxide 



carboxylic acid 



carboxylic ester 



acid chloride 



,c=o 



R' 



c=o 



>-0 



CO, 



R' 

V 



,c=o 



HO 



R' 

\ 



,c=o 



R"0 



R' 

V 



c=o 



/ 

CI 



R' 

N,N-dimethyJ C = 

carboxamide ,„„ -, J 
(LH 3 ) 2 in 



RCH 2 OMgX prim, alcohol RCH 2 OH 



R' 
I 
R-C-OMgX 
1 
H 

R' 
1 
R-C-OMgX 
I 
R" 

RCQ 2 MgX 



R' 
I 
sec. alcohol R— CHOH 

R' 

I 
ten. alcohol R— C— OH 
I 
R" 

carboxylic acid RC0 2 H 



R'C0 2 MgX + RH carboxylic acid R'CO z H 



R' 
I 
R-C-OMgX 
I 
R 

R' 
I 
R-C-OMgX 
I 
R 



R' 
I 
R— C— OMgX ketone 

I 
N(CH 3 ) 2 



R' 

\ 



good 



good 



good to poor 



good 



good 



R' 

I 
ten. alcohol R— C— OH good to poor 

I 
R 

R' 
I 
ten. alcohol R— C— OH good to poor 

I 
R 



r — O good to poor 



RC0 2 MgX. This substance is a halomagnesium salt of carboxylic acid and 
RMgX + co 2 - 



->• R — C 



O 

// 



o 

— - — > R-C 
\ \ 

OMgX OH 



acidification produces the carboxylic acid itself. 



O 



Acid chlorides such as acetyl chloride, CH 3 C usually combine with 

CI 
two moles of Grignard reagent to give a tertiary alcohol. Presumably, the 
first step is addition to the carbonyl bond : 



sec 9.9 organometallic compounds 233 



O OMgX 

CH 3 C + RMgX ► CHj-C-Cl 



CI 



R 



C OM g X O 

CH 3 -C-^C1 ► CH 3 -C-R + MgXCl 

R 

O R 

II I 

CH3-C-R + RMgX ► CH 3 -C— OMgX 



The reaction of acid chloride with RMgX is impractical for the synthesis 
of ketones because RMgX usually adds rapidly to the ketone as it is formed. 
However, the use of the less reactive organocadmium reagent, RCdX, usually 
gives good yields of ketone. 

Reaction of esters with Grignard reagents is similar to the reaction of acid 
chlorides and is very useful for synthesis of tertiary alcohols with two identical 
groups attached to the carbonyl carbon: 

T 

CH 3 -C + 2 RMgX <• CH 3 -C-OMgX + MgX(OCH 3 ) 

OCH 3 r 

Many additions of Grignard reagents to carbonyl compounds proceed in 
nearly quantitative yields, while others give no addition product whatsoever. 
Trouble is most likely to be encountered in the synthesis of tertiary alcohols 
with bulky alkyl groups, because the R group of the Grignard reagent will 
be hindered from reaching the carbonyl carbon of the ketone and side re- 
actions may compete more effectively. 

Addition to Carbon-Nitrogen Triple Bonds. Nitrogen is a more electro- 
negative element than carbon and the nitrile group is polarized in the sense 

88 se 
— C=N. Accordingly, Grignard reagents add to nitrile groups in much the 

same way as they add to carbonyl groups : 

R' 

e a \ e a 

R— C=N + R — MgX ► C = N — MgX 

R 

Hydrolysis of the adducts leads to ketimines, which are unstable under the 
reaction conditions and rapidly hydrolyze to ketones: 



R' 

\ e e H®, H 2 

C = N— MgX > 

R 



R' 

\ 

C=NH 
/ 
R 

ketimine 



R' 
i^2 C = + NH 4 
R 



chap 9 alkyl halides and organometaliic compounds 234 

Small-Ring Cyclic Ethers. Grignard reagents react with most small-ring 
cyclic ethers by S N 2 displacement. The angle strain in three- and four- 
membered rings facilitates ring opening, whereas the strainless five- and six- 
membered cyclic ethers are not attacked by Grignard reagents. 

RMgX + H 2 C-CH 2 ► RCH 2 -CH 2 OMgX 

O 

ethylene oxide 



RMgX + H 2 C O ► RCH 2 CH 2 CH 2 OMgX 

C 
H 2 

trimethylene oxide 



D. ORGANOSODIUM AND ORGANOLITHIUM COMPOUNDS 

Alkylsodium and alkyllithium derivatives behave in much the same way as 
organomagnesium compounds, but with increased reactivity. As mentioned 
previously, they are particularly sensitive to air and moisture, and react with 
ethers, alkyl halides, active hydrogen compounds, and multiple carbon- 
carbon, carbon-oxygen, and carbon-nitrogen bonds. In additions to carbonyl 
groups they give fewer side reactions than Grignard reagents and permit 
syntheses of very highly branched tertiary alcohols. Triisopropylcarbinol, 
which has considerable steric hindrance between its methyl groups, can be 
made from diisopropyl ketone and isopropyllithium, but not with the cor- 
responding Grignard reagent. 

H 3 C O C h 3 h 3 C /H 3 C \ 

\ II / \ / \ \ 

CH-C-CH + CH-Li > I CH J C-OH 

H 3 C CH 3 H 3 C VHjC 7 / 3 

triisopropylcarbinol 

E. SOME COMMERCIAL APPLICATIONS OF ORGANOMETALLIC 
COMPOUNDS 

Tetraethyllead, bp 202°, is the most important organometaliic compound in 
commercial use. It greatly improves the antiknock rating of gasoline in 
concentrations on the order of 1 to 3 ml per gallon (Section 3-3). 1,2-Dibro- 
moethane is added to leaded gasoline to convert the lead oxide formed in 
combustion to volatile lead bromide and thus diminish deposit formation. 
Most tetraethyllead is made by the reaction of a lead-sodium alloy with ethyl 
chloride. The excess lead is reconverted to the sodium alloy. Tetramethyllead 
shows some advantage over tetraethyllead in high-performance engines. 

4C 2 H 5 C1 + 4PbNa >• (C 2 H 5 ) 4 Pb -I- 4 NaCl + 3 Pb 

Some alkylmercuric halides, such as ethylmercuric chloride, have fungicidal 
properties and are used to preserve seeds and grains. This practice, however, 
may be having a deleterious effect on ducks and other kinds of wildlife. 



sec 9.9 organometallic compounds 235 
F. FERROCENE 

An exceptionally stable organometallic compound of unusual structure was 
discovered in 1952. An orange solid, mp 174°, containing iron, it was given 
the name ferrocene. It can be prepared by converting cyclopentadiene to its 
anion and treating this with a ferrous salt. 




> Fe + 2 NaCl 



Bonding in the ferrocene molecule results from sharing of the n electrons 
of the two rings with iron. The carbons are in parallel planes about 3.4 A 
apart with the iron between, hence the name " sandwich compound." This 
compound is not simply an ionic salt, Fe 2 ®(C 5 H 5 e ) 2 , because it is insoluble 
in water, soluble in most organic solvents, and is not affected by boiling with 
dilute acid or base. In contrast, the truly ionic sodium salt of cyclopentadiene 
reacts rapidly with water or acids and is insoluble in most organic solvents. 

Analogous sandwich compounds (metallocenes) can be formed with many 
other transition metals, such as nickel, cobalt, and manganese. Other sandwich 
compounds are known with different ring sizes — six-, seven-, and eight- 
membered rings with metals as diverse as chromium and uranium. 

A few metals form more or less stable complexes with alkenes, which have 
metal-carbon bonds. Silver ion, as in solutions of silver nitrate, complexes 
some alkenes and alkadienes strongly enough to make them soluble in water 
and/or form crystalline silver nitrate complexes. Platinum, rhodium, and 
palladium, which in the metallic state are good hydrogenation catalysts, form 
some quite stable alkene complexes. A specially interesting example is stable 
71-cyclopentadienyldiethenerhodium, 1 a "half-sandwich" compound. 




£5 



c c 

H 2 H 2 

The organic groups of such compounds do not usually show much nucleo- 
philic character. Indeed, some platinum-ethene complexes are stable in strong 
hydrochloric acid. 



1 The designation 77-cyclopentadienyl means that the C 5 H 5 group is bound to the metal 
as in ferrocene. 



chap 9 alkyl halides and organometallic compounds 236 

summary 

Alkyl halides (haloalkanes) have low water solubilities and except for alkyl 
fluorides their boiling points are close to those of the alkanes of similar 
molecular weight. Other than their nmr spectra, their only striking spectral 
characteristics are strong infrared bands at 1350-1000 cm -1 (C— F stretch) 
and at 800-600 cm"" 1 (C— CI stretch). 
Alkyl halides can be prepared from alkanes, alkenes, and alcohols. 

RCH 2 CH 3 

ny " HX 

-> RCHXCH 3 < RCHOHCH3 (X = C1, Br, I) 



RCH 2 CH 2 Br 

The reactions of alkyl halides include displacement and elimination reac- 
tions. (See summary in Chapter 8.) 

Vinyl halides (RCH=CHX) are much less reactive than alkyl halides in 
nucleophilic displacement reactions. Allyl halides (RCH=CHCH 2 X) are 
much more reactive, because of the ease of forming the resonance-stabilized 

cation RCH=CH-CH 2 +-> RCH-CH=CH 2 . 

Polyhalogen compounds are useful solvents. Di- and trihalomethanes are 
rather unreactive in displacement reactions with strong bases. Chloroform, 
however, undergoes a ready elimination with base to give a carbene, :CC1 2 . 
This can add to the double bond of alkenes. Activated carbenes (as from 
diazomethane photolysis) also undergo insertion reactions at C— H bonds. 

^C -c. \ \ 

II + :CZ 2 ► LCZ 2 „C-H + :CZ 2 * ► ^C-CHZ 2 

/C^ -C' / / 

Fluoroalkanes have rather different properties than the other haloalkanes; 
they are very much more volatile, their polymers have exceptional thermal 
and chemical stability (Teflon, Viton), and their toxicities vary widely. 

Alkyl halides can be converted to organometallic compounds whose prop- 
erties vary from the highly ionic and reactive (R e Na®) to the highly covalent 
and unreactive (R 4 Pb, R 2 Hg). Midway are organomagnesium compounds 
(Grignard reagents), the most important of the organometallics. Grignard 

RX + Mg ► RMgX (in dry ether) 

reagents react with virtually all organic compounds except alkanes, alkenes, 
and ethers. A summary of important Grignard reactions follows and the final 
step (required in all but the first example) is addition of an active hydrogen 
compound, such as water, to destroy a halomagnesium salt. 



:gx + h 2 o — 


► RH + MgOHX 


+ 


o 2 — 


— 


ROH 


+ 


R'CHO 


— ► 


— ♦ R'RCHOH 


+ 


O 
II 
R'-C- 


R' - 


R 

-♦ — ♦ R' 2 COH 


+ 


O 
II 
RC-OCH3 


-—►—»—» R'R 2 COH 


+ 


co 2 — > 


- 


RC0 2 H 


+ 


RC=N 


— 


— — » R 2 C = 


+ 


V^ 


— 


RCH 2 CH 2 OH 



exercises 237 



ROH, RC0 2 H, NH 3 , and deriva- 
tives, and RC=CH react similarly 

(S g and I 2 react similarly) 

(HCHO reacts similarly and 
gives a primary alcohol) 



(via R'RC=0; RCOC1 reacts 
similarly) 



(O 



reacts similarly ) 



Organocadmium compounds, RCdX, are less reactive than Grignard 
reagents. They react with acid chlorides but not with ketones. 

RCdX + R'COCl > R'RC=0 + CdXCl 

The stable organoiron compound ferrocene, Fe(C 5 H 5 ) 2 , has a sandwich 
structure. Other transition metals form similar compounds. 

exercises 

9-1 Show how 2-bromoheptane can be prepared starting from (a) 2-heptanol, 
(b) 1-heptanol, (c) 1-heptyne. 

9-2 When 3-ethyl-3-chloropentane reacts at room temperature with aqueous 
sodium carbonate solution a mixture of two compounds, one an alcohol 
and one an alkene, is obtained. 

a. Write the equations for these two reactions and name each of the 
products. 

b. Suggest changes in reaction conditions that would favor formation of 
the alkene. 

(You may wish to review Sections 8-12 and 8-13.) 

9-3 a. Write resonance structures for the transition states of S N 2 substitution for 
allyl and n-propyl chlorides with hydroxide ion and show how these can 
account for the greater reactivity of the allyl compound. 
b. Would you expect that electron-donating or electron-withdrawing groups 
substituted at the y carbon of allyl chloride would increase the S N 2 
reactivity of allyl chloride ? 

9-4 The rate of formation of the CH 2 -addition product from iodomethylzinc 
iodide and cyclohexene is first order in each participant. Suggest a mechanism 
that is in accord with this fact. 



chap 9 alkyl halides and organometallic compounds 238 

9-5 What products would you expect from the reaction of bromoform, CHBr 3 , 
with potassium /-butoxide in ?-butyl alcohol in the presence of (a) trans-2- 
butene, (b) cw-2-butene? 

9-6 Would you expect the same products if, instead of addition to the carbonyl 
group, the acyl halides were to undergo a simple S N 2 displacement of halogen 
with the Grignard reagent acting to furnish R: e ? Explain why simple dis- 
placement is unlikely to be the correct mechanism from the fact that acid 
fluorides react with Grignard reagents faster than acid chlorides, which in 
turn react faster than acid bromides. 

9-7 What products would you expect to be formed in an attempt to synthesize 
hexamethylethane from ^-butyl chloride and sodium ? Write equations for the 
reactions involved. 

9 • 8 Write structures for the products of the folio wing reactions involving Grignard 
reagents. Show the structures of both the intermediate substances and the 
substances obtained after hydrolysis with dilute acid. Unless otherwise 
specified, assume that sufficient Grignard reagent is used to cause those reac- 
tions to go to completion which occur readily at room temperatures. 

a. C 6 H 5 MgBr + C 6 H 5 CHO 

b. CH 3 MgI + CH 3 CH 2 C0 2 C 2 H 5 

c. (CH 3 ) 3 CMgCl + C0 2 

d. CH 3 CH 2 MgBr + C1C0 2 C 2 H 5 

e. CH 3 MgI + CH 3 COCH 2 CH 2 C0 2 C 2 H 5 

(1 mole) (1 mole) 

O 

II 
/ C 6 H 5 MgBr + CH 3 0-C-OCH 3 

o o 

II II 

g. (CH 3 ) 3 CCH 2 MgBr + CH3C-O-CCH3 

9-9 Show how each of the following substances can be prepared by a reaction 
involving a Grignard reagent : 

H 2 C 

a. H 2 C^ \)H— CH 2 OH (two ways) 

H 2 C 

b. CH 2 =CH-C(CH 3 ) 2 OH 

c. CH 3 CH 2 CH(OH)CH 3 (two ways) 

d. (CH 3 CH 2 ) 3 COH (three ways) 

H,C S 

e. | ^CHOH 
H 2 C 

9-10 Complete the following equations: 

a. C 6 H 5 CH 2 CH 2 MgBr + (CH 3 )2S0 4 * 



exercises 239 



b. C 2 H 5 MgBr + CH 3 C=C-CH 2 Br < 

c. CH 2 =CH-CH 2 Li + CH 2 =CH-CH 2 C1 

d. CH 3 CH 2 CH 2 MgBr + ClCH 2 OCH 3 



9-11 Each of the following equations represents a " possible " Grignard synthesis. 
Consider each equation and decide whether or not you think the reaction 
will go satisfactorily. Give your reasoning and, for those reactions that are 
unsatisfactory, give the expected product or write "No Reaction" where 
applicable. 

a. methylmagnesium iodide + butyryl chloride ► * n-propyl 

methyl ketone 

b. methylmagnesium iodide + CH 3 CH=N—CH 3 ► ► 

CH 3 
I 
CH 3 CH 2 -N-CH 3 

Mg CH 2 = 

c. 2-bromoethyl acetate * Grignard reagent ► ► 

ether 

3-hydroxypropyl acetate 

d. allylmagnesium chloride + ethyl bromide ► 1-pentene 

9-12 Predict the products of each of the following Grignard reactions before and 
after hydrolysis. Give reasoning or analogies for each. 



a. CH 3 MgI + HC0 2 C 2 H 5 ► 

b. CH 3 CH 2 MgBr + CS 2 ► 

c. CH 3 CH 2 MgBr + NH 3 »■ 

9-13 Show how each of the following substances can be synthesized from the 
indicated starting materials by a route that involves organometallic substances 
in at least one step : 

a. (CH 3 ) 3 C-D from (CH 3 ) 3 CCI 

b. CH 3 C=C-C0 2 H from CH^CH 

CH 3 
I 

c. CH 3 -C-CH 2 I from (CH 3 ) 4 C 

I 
CH 3 

CH 3 

I 

d. CH 3 -C-CH(CH 3 ) 2 



'•(OJt 



OH 

COH from < ^Br 
CH, 



*3 

I 

/ CH 3 -C-CH 2 CH 2 CH 2 CH 2 OH from (CH 3 ) 3 CCH 2 C1 



9-14 Explain what inferences about the stereochemistry of the E2 reaction can be 
made from the knowledge that the basic dehydrohalogenation of A is 
exceedingly slow compared with that of B. 







chap 9 alkyl halides and organometallic compounds 240 


CI 

]_ 


H 

1 


Cl 


Cl 


Ml 


p: 


c lAr 




H 

i 


Cl 


Cl 
B 


H 



9' 15 Classify each of the following reactions by consideration of yield, side reac- 
tions, and reaction rate as good, fair, or bad synthetic procedures for prepa- 
ration of the indicated products under the given conditions. Show your reason- 
ing and designate any important side reactions. 



CH, 



CH, 



CHL-C-C1 + CH 3 -C-ONa 



50° 



CH, 



CH, 



CH, 



CH 3 
I 
-C— o- 

I 

CH 3 



CH 3 
I 
-C-CH, + NaCl 
I 
CH, 



/■ 





CH 3 


CH 3 


CH 3 - 


1 25° 


1 

T _ /-< _ r\ 1 


1 *r C1I3 v- Uri ' *^H 4 t* ^113 \_, \j 1 




CH 3 


CH 3 




CH 3 CH 3 




CH 3 - 


1 H <~> 1 

-CH-CH-CH 3 ]0 2 0O > CH 3 -CH-CH=CH 2 + HC1 




1 
Cl 




H 3 C 


^N H Nai Hi V^4 




H K 


\ A^St' H\ / H 






CH 3 O 


CH 3 O 


CH 3 - 


1 II 50° 

-C-CH 2 C1 +CH 3 C-ONa ► CH 3 - 


1 II 
- C- CH 2 - O- C- CH 3 +NaCl 




CH 3 


CH 3 




CH 3 


CH 3 


CH 3 - 


1 35° 


1 
-C-0-CH 2 CH 3 +CH 3 CH 2 C1 


C U t LnjLn2ULH 2 trl3 ' Ln 3 " 




CH 3 


CH 3 




O 


O 


CH 2 = 
CH 2 = 


II 25° 


H-0-C-CH 3 + AgCl 
-CH 2 F + iSbCl 3 


-tntlTLrljL UAg * Cri 2 — C 

=CH-CH 2 C| +iSbF 3 5 °° > CH 2 =CH- 



9-16 Consider each of the following compounds to be in unlabeled bottles in pairs 
as indicated. Give for each pair a chemical test (preferably a test tube reaction) 
that will distinguish between the two substances and show what the observa- 
tions will be. Write equations for the reactions expected. 





Bottle A 


Bottle B 


a. 


(CH 3 ) 3 CCH 2 C1 


CH 3 CH 2 CH 2 CH 2 C1 


b. 


BrCH=CHCH 2 Cl 


ClCH=CHCH 2 Br 


c. 


(CH 3 ) 3 CC1 


(CH 3 ) 2 CHCH 2 C1 


d. 


CH 3 CH=CHC1 


CH 2 =CHCH 2 C1 


e. 


(CH 3 ) 2 C=CHC1 


CH 3 CH 2 CH=CHC1 


f. 


CH 3 CH 2 CH=CHC1 


CH 2 =CHCH 2 CH 2 C1 



-o 




UAh 


' — 


^- 






X 


■-.. •■■■ -■= 




o 
- o 










o 
- o 




— __ 




— ^ 
i 


o 
- o 

CO 


o 






- 



s 

o 



5<a 






-§ 








^- 


*Tfc 


. ., ... . -^^M 










- 


o 
-o 




1 


- 






' > ' -1 


- 




CQ 


1 




o 
- o 

CO 


CJ 


•" ' ' 1 


• 



c 



s„ 



1 3 
8* 



241 



UOISSIUISUBI} % 



chap 9 alkyl halides and organometallic compounds 242 

9-17 Show how the pairs of compounds listed in Exercise 9-16 could be distin- 
guished by spectroscopic means. 

9-18 Deduce the structures of the two compounds whose nmr and infrared spectra 
are shown in Figure 94. Assign as many of the infrared bands as you can and 
analyze the nmr spectra in terms of chemical shifts and spin-spin splittings. 

9-19 Write balanced equations for reactions that you expect would occur between 
the following substances (1 mole) and 1 mole of pentylsodium. Indicate 
your reasoning where you make a choice between several possible alternatives. 

a. water /. allyl chloride 

b. diethyl ether g. acetic acid (added slowly to the 

c. f-butyl chloride pentylsodium) 

d. pentyl iodide h. acetic acid (pentylsodium added 

e. propene slowly to it) 



chapter 10 
alcohols and ethers 



chap 10 alcohols and ethers 245 

Alcohols, ROH, and ethers, ROR, can be regarded as substitution prod- 
ucts of water. With alcohols, we shall be interested on the one hand in 
reactions that proceed at the O— H bond without involving the C— O bond or 
the organic group directly, and on the other hand with processes that result in 
cleavage of the C— O bond or changes in the organic group. The reactions 
involving the O— H bond are expected to be similar to the corresponding 
reactions of water. 

The simple ethers do not have O— H bonds, and the few reactions that they 
undergo involve the substituent groups. 

Alcohols are classed as primary, secondary, or tertiary according to the 
number of hydrogen atoms attached to the hydroxylic carbon atom. This 

RCH 2 OH R 2 CHOH R 3 COH 

primary alcohol secondary alcohol tertiary alcohol 

classification is necessary because of the somewhat different reactions the 
three kinds of compounds undergo. 

The nomenclature of alcohols has been discussed previously (page 187). 

The first member of the alcohol series, methanol or methyl alcohol, 
CH3OH, is a toxic liquid. In the past it was prepared by the destructive dis- 
tillation of wood and acquired the name wood alcohol. Many cases of 
blindness and death have resulted from persons drinking it under the impres- 
sion that it was ethyl alcohol, CH 3 CH 2 OH (ethanol). 

Ethanol is intoxicating in small amounts but toxic in large amounts. The 
higher alcohols are unpleasant tasting, moderately toxic compounds which are 
produced in small amounts along with ethanol by fermentation of grain. A 
mixture of higher alcohols is called fusel oil. (Fusel means bad liquor in 
German.) One hundred proof whiskey (or whisky) 1 contains 50% ethanol by 
volume and 42.5% ethanol by weight. Pure ethanol (200 proof) is a strong 
dehydrating agent and is corrosive to the gullet. 

An understanding of the chemistry of alcohols is important to under- 
standing the functioning of biological systems which involve a wide variety 
of substances with hydroxyl groups. The OH group attached to a carbon 
chain often dramatically changes physical properties and provides a locus for 
chemical attack. 



10-1 physical properties of alcohols 

Comparison of the physical properties of alcohols with those of hydrocar- 
bons of comparable molecular weight shows several striking differences, 
especially for the lower members. Alcohols are substantially less volatile and 
have higher melting points and greater water solubility than the corres- 
ponding hydrocarbons, although the differences become progressively 
smaller as molecular weight increases. x The first member of the alcohol 

1 Scotch and Canadian whisky, but Irish and American whiskey. 



chap 10 alcohols and ethers 246 

series, CH 3 OH (methanol or methyl alcohol), has a boiling point of 65°, 
whereas ethane, CH 3 CH 3 , with almost the same molecular weight boils at 
-89°. 

The profound effect of the hydroxyl group on the physical properties of 
alcohols is caused by hydrogen bonding. The way in which the molecules of 
hydroxylic compounds interact via hydrogen bonds was described earlier 
(Section 1-2B). 

The water solubility of the lower-molecular-weight alcohols is high and is 
also a result of hydrogen bonding. In methanol, the hydroxyl group accounts 
for almost half of the weight of the molecule, and it is thus not surprising that 
the substance is miscible with water in all proportions. As the size of the 
hydrocarbon group of an alcohol increases, the hydroxyl group accounts for 
progressively less of the molecular weight, and hence water solubility decreases 
(Figure 10-1). Indeed, the physical properties of higher-molecular-weight 
alcohols are very similar to those of the corresponding hydrocarbons. 

An interesting effect of chain branching on solubility can be seen in the 
four butyl alcohols. «-Butyl alcohol is soluble to the extent of 8 g in 100 g of 
water whereas ?-butyl alcohol is completely miscible with water (Table 10-1.) 
Branching also affects volatility. The highly branched alcohol, f-butyl 
alcohol, has a boiling point 35° lower than that of «-butyl alcohol. The 
effect of branching on melting point is in the opposite direction because 
crystal packing is improved by branching. The result is that ^-butyl alcohol 
has the highest melting point and the lowest boiling point of the four isomeric 
C 4 alcohols. (?-Butyl alcohol is liquid over a range of only 58° whereas 
«-butyl alcohol is liquid over a range of 208°.) See Table 10-1. 

Figure 10-1 Dependence of melting points, boiling points, and water solu- 
bilities of continuous-chain primary alcohols (C„H 2 „ + 2 0) on«. (The alcohols 
with C 3 or fewer carbons are infinitely soluble in water.) 



250 



200 



150 



100 




-100 



3 3 



sec 10.2 spectroscopic properties of alcohols — hydrogen bonding 247 
Table 10-1 Physical properties of the butyl alcohols 



name 


formula 


mp, 

°C 


bp, 
°C 


solubility, 
g/lOOg water 


n-butyl alcohol (1-butanol) 


CH 3 CH 2 CH 2 CH 2 OH 


-90 


118 


8.0 


isobutyl alcohol 
(2-methy 1- 1 -propanol) 


(CH 3 ) 2 CHCH 2 OH 


-108 


108 


10.0 


.s-butyl alcohol (2-butanol) 


CH 3 CH 2 CHOHCH 3 


-114 


100 


12.5 


/-butyl alcohol (2-methyl- 
2-propanol) 


(CH 3 ) 3 COH 


25 


83 


00 



10-2 spectroscopic properties of alcohols— hydrogen 
bonding 

The hydrogen-oxygen bond of a hydroxyl group gives a characteristic 
absorption band in the infrared and, as we might expect, this absorption is 
considerably influenced by hydrogen bonding. For example, in the vapor state 
(in which there is essentially no hydrogen bonding because of the large inter- 
molecular distances), ethanol gives an infrared spectrum with a fairly sharp 
absorption band at 3700 cm -1 owing to a free or unassociated hydroxyl 
group (Figure 10-2a). In contrast, this band is barely visible at 3640 cm -1 in 
the spectrum of a 10% solution of ethanol in carbon tetrachloride (Figure 
10-2b). However, there is a relatively broad band around 3350 cm -1 which is 
characteristic of hydrogen-bonded hydroxyl groups. The shift in frequency of 
about 300 cm -1 is not surprising, since hydrogen bonding weakens the 
O— H bond; its absorption frequency will then be lower. The association 
band is broad because the hydroxyl groups are associated in aggregates of 
various sizes and shapes, giving rise to a variety of different kinds of hydro- 
gen bonds and therefore a spectrum of closely spaced O— H absorption 
frequencies. 

In very dilute solutions of alcohols in nonpolar solvents, hydrogen bonding 
is minimized ; but as the concentration is increased, more and more of the 
molecules become associated and the intensity of the infrared absorption 
band due to associated hydroxyl groups increases at the expense of the free 
hydroxyl band. Furthermore, the frequency of the association band is a 
measure of the strength of the hydrogen bond. The lower the frequency 
relative to the position of the free hydroxyl group, the stronger is the hydrogen 
bond. As we shall see in Chapter 13, the hydroxyl group in carboxylic acids 
(RC0 2 H) forms stronger hydrogen bonds than alcohols and, accordingly, 
absorbs at lower frequencies (lower by about 400 cm -1 ). 

From the foregoing discussion of the influence of hydrogen bonding on the 
infrared spectra of alcohols, it should come as no surprise that the nuclear 
magnetic resonance spectra of the hydroxyl protons of alcohols are similarly 
affected. Thus the chemical shift of a hydroxyl proton is influenced by the 
degree of molecular association through hydrogen bonding and on the 



chap 10 alcohols and ethers 248 



m 



wavelength, fJL 



4 5 5 



9 10 12 14 



T— 1 T T 



~V 



,a, 



-I i i i _J i i ■ l_ 



3600 3200 2800 2400 2000 2000 




1 800 1 600 1 400 1 200 1 000 800 

frequency, cm - 



(b) 



J i i i i 



wavelength, jJ. 




10 12 14 



i 1 r 



3600 3200 2800 2400 2000 2000 



-. 1 i I I L_ 



1 800 1 600 1 400 

frequency, cm -1 



J l l 1_ 



1 200 1 000 800 



Figure 10-2 Infrared spectrum of ethanol in the vapor phase (a) and as a 10% 
solution in carbon tetrachloride (b). 



sec 10.3 preparation of alcohols 249 

strengths of the hydrogen bonds. Except for alcohols that form intramolecular 
hydrogen bonds, the OH chemical shift varies extensively with temperature, 
concentration, and the nature of the solvent. Also, resonance appears at 
lower magnetic fields (i.e., the chemical shift is larger relative to TMS) as the 
strengths of hydrogen bonds increase. Thus, the chemical shifts of the OH 
protons of simple alcohols as pure liquids generally fall between 4 and 5 ppm 
downfield with respect to tetramethylsilane, but when the degree of hydrogen 
bonding is reduced by dilution with carbon tetrachloride, the OH reso- 
nances move upheld. With ethyl alcohol, the shift is found to be 3 ppm between 
the pure liquid and very dilute solution in carbon tetrachloride. 

One may well question why it is that absorptions are observed in the infra- 
red spectrum of alcohols which correspond both to free and hydrogen- 
bonded hydroxyl groups, while only one OH resonance is observed in their 
nmr spectra. The answer is that the lifetime of any one molecule in the free 
or unassociated state is long enough to be detected by infrared absorption but 
too short to be detected by nmr. Consequently, one sees only the average OH 
resonance for all species present. (For a discussion of nmr and rate processes, 
see Section 7-6D.) 



10-3 preparation of alcohols 



We have already encountered most of the important methods of preparing 
alcohols, which are summarized below. 

1. Hydration of alkenes (Section 4-4). The direction of addition is governed 

RCH=CH 2 + H 2 HS > RCHOHCH3 

by Markownikoff 's rule and primary alcohols, therefore, cannot be made this 
way (except for CH 3 CH 2 OH). 

2. Hydroboration of alkenes (Section 4-4F). The direction of addition is 

RCH=CH 2 + B 2 H 6 ► (RCH 2 CH 2 ) 3 B H2 ° 2 . RCH 2 CH 2 OH 

" anti-Markownikoff " and primary alcohols, therefore, can be made this way. 

3. Addition of hypohalous acids to alkenes (Section 4-4). The HO group 

RCH=CH 2 + HOC1 ► RCHOHCH 2 Cl 

becomes bonded to the carbon atom bearing the least number of hydrogen 
atoms. 

4. S N 2 and S N 1 hydrolyses of alkyl halides (Sections 8-7 to 8T0). Primary 

RCH 2 CH 2 C1 + OH e ► RCH 2 CH 2 OH + Cl e S N 2 

r 3 CC1 H2 ° > R3COH S N 1 

and secondary but not tertiary alkyl halides require hydroxide ion. A side 
reaction is elimination, which can be especially important with tertiary 
halides and strong bases. 



chap 10 alcohols and ethers 250 

5. Grignard reagents to carbonyl groups (Section 9-9C). Grignard reagents 



o 

RMgBr + H— C-H ► RCH 2 MgBr H2 ° > RCH 2 OH primary 

alcohol 

O O MgBr OH 

II . | ho I secondary 

+ R'— C— H ► R'CH— R — ► R'-CH— R alcohol 



e e 

O O MgBr OH 

II I H.O I 

+ R' — C— R" ► R — C — R" ► R'-C-R" 

I I 

R R 



tertiary 
alcohol 



O O MgBr OH 

+ R' — C— OR" ► R' — C-R — ► R' — C-R tertiary 

| | alcohol 

R R 



with ketones can be used to produce tertiary alcohols in which all three R 
groups are different; with esters, tertiary alcohols result in which at least 
two of the R groups are identical. 

6. Reduction of carbonyl compounds (to be described in Section 11-4F): 



R— C ► R— CH 2 OH primary alcohol 



R— C ► R— CH 2 OH primary alcohol 

OR' 



° 0H , u , 

|| | secondary alcohol 

R— C-R ► R— CH— R 



A few of the reactions mentioned have been adapted for large-scale produc- 
tion. Ethanol, for example, is made in quantity by the hydration of ethene, 
using an excess of steam under pressure at temperatures around 300° in the 
presence of phosphoric acid : 

300°, H 3 PO„ 
CH 2 = CH 2 + H 2 , CH 3 CH 2 OH 

A dilute solution of ethanol is obtained which can be concentrated by distilla- 
tion to a constant boiling point mixture that contains 95.6 % ethanol by weight. 
Removal of the remaining few percent of water to give " absolute alcohol " is 
usually achieved either by chemical means or by distillation with benzene, 
which results in preferential separation of the water. Ethanol is also made in 
large quantities by fermentation, but this route is not competitive for indus- 
trial uses with the hydration of ethene. 



sec 10.4 reactions involving the O — H bond 251 

Isopropyl alcohol and /-butyl alcohol are also manufactured by hydration 
of the corresponding alkenes. The industrial synthesis of methyl alcohol 
involves hydrogenation of carbon monoxide. Although this reaction has a 
favorable AH value of —28.4 kcal, it requires high pressures and high tem- 
peratures and a suitable catalyst; excellent conversions are achieved using a 
zinc oxide-chromic oxide catalyst: 

^ ^ „ 4°°°. 200 atm 

CO + 2 H 2 -—± » CH 3 OH AH = - 28.4 kcal 

ZnO— Cr0 3 



chemical reactions of alcohols 

10-4 reactions involving the 0~H bond 

A. ACIDIC AND BASIC PROPERTIES 

Several important reactions of alcohols involve only the oxygen-hydrogen 
bond and leave the carbon-oxygen bond intact. An important example is salt 
formation with acids and bases. 

Alcohols, like water, are amphoteric and are neither strong bases nor 
strong acids. The acid ionization constant (K HA ) of ethanol is about 10" 18 — 
slightly less than that of water. Ethanol can be converted to a salt by the salt 
of a weaker acid such as ammonia (X HA ~ 1CT 35 ), but it is usually more con- 
venient to employ sodium or sodium hydride. The reactions are vigorous 
but can be more easily controlled than the analogous reactions with water. 



C 2 H,O a Na lB + 



1 2 

sodium amide sodium ethoxide 

(sodamide) 

C 2 H 5 OH + Na e H e ► C 2 H 5 O e Na® + H 2 

The order of acidity of various alcohols is generally primary > secondary > 
tertiary; ?-butyl alcohol is therefore considerably less acidic than ethanol. 
The anions of alcohols are known as alkoxide ions. 

CH 3 Oe C 2 H 5 O e (CH 3 ) 2 CHOe (CH 3 ) 3 CO e 

methoxide ethoxide isopropoxide f-butoxide 

Alcohols are bases comparable in strength to water and are converted to 
their conjugate acids by strong acids. An example is the reaction of methanol 
with hydrogen bromide to give methyloxonium bromide. 



CH,:0:H + HBr . CH 3 :0:H + Br e 



H„ 

methyloxonium bromide 



The reaction of hydrogen bromide with water proceeds in an analogous 
manner : 

Ha, 
H:6:H + HBr , H:0:H + Br e 

hydroxonium bromide 



chap 10 alcohols and ethers 252 
B. ETHER FORMATION 

Alkoxide formation is important as a means of generating a powerful 
nucleophile that will readily enter into S N 2 reactions. Whereas ethanol reacts 
only slowly and incompletely with methyl iodide, sodium ethoxide in ethanol 
solution reacts rapidly with methyl iodide and gives a high yield of methyl 
ethyl ether. 

fast 
CH 3 I + C 2 H 5 Oe Na ffl ► CH 3 OC 2 H 5 + Nal 

In fact, the reaction of alkoxides with alkyl halides or alkyl sulfates is an 
important general method for the preparation of ethers, and is known as the 
Williamson synthesis. Complications can occur because the increase of 
nucleophilicity associated with the conversion of an alcohol to an alkoxide 
ion is always accompanied by an even greater increase in eliminating power 
by the E2-type mechanism. The reaction of an alkyl halide with alkoxide may 
then be one of elimination rather than substitution, depending on the tem- 
perature, the structure of the halide, and the alkoxide (Section 8T2). For 
example, if we wish to prepare isopropyl methyl ether, better yields would be 
obtained if we were to use methyl iodide and isopropoxide ion rather than 
isopropyl iodide and methoxide ion because of the prevalence of E2 elimina- 
tion with the latter combination : 

CH3I + (CH 3 ) 2 CHO e ^— (CH 3 ) 2 CHOCH 3 + ie 

(CH 3 ) 2 CHI + CH 3 O e E ' > CH 3 CH=CH 2 +CH 3 OH + ie 

Potassium /-butoxide is often an excellent reagent to achieve E2 elimina- 
tion, since it is strongly basic but so bulky as to not undergo S N 2 reactions 
readily. 



C. ESTER FORMATION 

Esters are one of a number of compounds containing the carbonyl group 

O 

II 
(— C— ) that are important to a discussion of alcohols and whose detailed 
chemistry will be discussed in subsequent chapters. 

OO OOO 

II II II II II 

R-C— OR R-C— OH R— C — CI R-C— H R-C — R 

esters carboxylic acids acyl halides aldehydes ketones 

Esters are produced by the reactions of alcohols with either acyl halides or 
carboxylic acids. Acyl halides, for example, have a rather positive carbonyl 
carbon because of the polarization of the carbon-oxygen and carbon-halogen 



sec 10.4 reactions involving the O — H bond 253 

bonds. Addition of an electron-pair-donating agent such as the oxygen of an 
alcohol occurs rather easily. 



ie O> O e 

i9 v^ -\.. i . 

CH 3 — C — CI + CH 3 :0:H , CH 3 -C-C1 

<se <se " | 

CH3OH 

[1] 



The complex [1] contains both an acidic group (CH 3 — O— H) and a basic 

O e I 

I 
group (— C— ), so that one oxygen loses a proton and the other gains a 

I 
proton to give [2], which then rapidly loses hydrogen chloride by either an 

El or E2 elimination to form the ester. The overall process resembles an 



O e 6) O 

| iv . || 



CH3-C-CI . CH 3 -C^C1 ► CH 3 -C-0-CH 3 + HC1 

CH 3 OH CH 3 

e 

[I] [2] 



S N 2 reaction, but the mechanism is different in being an addition-elimination 
with three transition states rather than a one-stage displacement reaction with 
one transition state. 
A similar but less complete reaction occurs between acetic acid and 



O O e 

II .. I 

CH 3 -C-OH + CH 3 :0=H ;=i CH 3 -C-OH ^ 



CH,— O-H 



OH O 

I II 

CH 3 -C-OH . CH 3 -C-OCH 3 + H 2 

CH 3 -0 



methanol. This reaction is slow in either direction in the absence of a strong 
mineral acid. Strong acids catalyze ester formation from the alcohol provided 
they are not present in large amount. The reason for the " too much of a good 
thing " behavior of the catalyst is readily apparent from a consideration of 
the reaction mechanism. A strong acid such as sulfuric acid may donate a 
proton to the unshared oxygen electron pairs of either acetic acid or methanol : 



chap 10 alcohols and ethers 254 



CH, 



// 







+ H 2 S0 4 



OH 



.OH 



CH 3 


// 

-C + 
\ 
OH 

[3] 
H 



HSO, 


CH,- 


1 
-0— H + 


e 
HSO, 



CH,— O— H + H,S0 4 ;= 

a 

Clearly, formation of methyloxonium bisulfate can only operate to reduce the 
reactivity of methanol toward the carbonyl carbon of acetic acid. However, 
this anticatalytic effect is more than balanced (at low concentrations of 
H 2 S0 4 ) by protonation of the carbonyl oxygen of the carboxylic acid [3], 
since this greatly enhances the electron-pair accepting power of the carbonyl 
carbon : 



eOH 

r 

CH 3 -C— OH 



OH 
I 
CH,-C-OH 



+ CH,-0-H 



[3] 



CH, 



OH 
I 
-C— OH 



CH 3 -0-H 

[4] 



The resulting intermediate [4] is in equilibrium with its isomer [5], which can 
lose a water molecule to give the protonated ester [6] : 



OH 

CH 3 -C— OH 

I 
CH3-O-H 

e 

[4] 



c 



OH, 



CH,-C— OH 



[5] 



S 0H 
II 
CH 3 — C— OCH 3 + H 2 

[6] 



O 

II ffi 

CH 3 -C-OCH 3 + H 3 

methyl acetate 



Transfer of a proton from [6] to water gives the reaction product. 

At high acid concentrations, essentially all the methanol would be con- 
verted to inert methyloxonium ion and the rate of esteriflcation would then be 
very slow, even though more of the oxonium ion of acetic acid would be 
present. 

As mentioned earlier, esteriflcation is reversible and, with ethanol and 
acetic acid, has an equilibrium constant of about 4 at room temperature, 
which corresponds to 66% conversion to ester with equimolal quantities. 
Higher ester conversions can be obtained by using an excess of either the 



o 

II 

CH,C-OH 



C,H,OH 



O 
II 
CH 3 C-0-C 2 H 5 + H 2 



K = 



[CH 3 CQ 2 C 2 H 5 ] [H 2 Q] 
[CH 3 C0 2 H][C 2 H 5 OH] ' 



sec 10.5 reactions involving the C — O bond of alcohols 255 

alcohol or the acid. The reaction may be driven to completion by removing 
the ester and (or) water as they are formed. 

Steric hindrance is very important in determining esteriflcation rates, and 
esters with highly branched groups, in either the acid or alcohol parts, are 
formed at slower rates and with smaller equilibrium constants than their less 
highly branched analogs. In general, the ease of esteriflcation for alcohols is 
primary > secondary > tertiary with a given carboxylic acid. 

10-5 reactions involving the C~ O bond of alcohols 

A. HALIDE FORMATION 

Alkyl halide formation from an alcohol and a hydrogen halide offers an 
important example of a reaction in which the C— O bond of the alcohol is 



R-rOH + HBr . RBr + H,0 



broken. The reaction is reversible and the favored direction depends on the 
water concentration (see Exercise 8-9). Primary bromides are often best 
prepared by passing dry hydrogen bromide into the alcohol heated to just 
slightly below its boiling point. 

Reaction proceeds at a useful rate only in the presence of strong acid, 
which can be furnished by excess hydrogen bromide or, usually and more 
economically, by sulfuric acid. The alcohol accepts a proton from the acid to 
give an alkyloxonium ion, which is more reactive in subsequent displacement 
with bromide ion than the alcohol, since it can more easily lose a neutral 
water molecule than the alcohol can lose a hydroxide ion (Section 8-1 IB). 

H 
Bi^T^R-^p-H > RBr + H 2 S N 2 



H 

| (-H 2 0) Br e 
R-O-H , R e ► RBr S N 1 



Hydrogen chloride is less reactive than hydrogen bromide or hydrogen iodide 
toward primary alcohols, and application of heat and the addition of a 
catalyst (zinc chloride) are usually required for preparation purposes. 

A solution of zinc chloride in concentrated hydrochloric acid (Lucas 
reagent) is widely used, in fact, to differentiate between the lower primary, 
secondary, and tertiary alcohols. Tertiary alcohols react very rapidly to give 
an insoluble layer of alkyl chloride at room temperature. Secondary alcohols 
react in several minutes, whereas primary alcohols form chlorides only on 
heating. 

Thionyl chloride, SOCl 2 , is a useful reagent for the preparation of alkyl 
chlorides, especially when the use of zinc chloride and hydrochloric acid is 



chap 10 alcohols and ethers 256 

undesirable. Addition of 1 mole of an alcohol to 1 mole of thionyl chloride 
gives an unstable alkyl chlorosulfite, which generally decomposes on mild 
heating to yield the alkyl chloride and sulfur dioxide. 

o 

— HCl " 

ROH + SOCl 2 > R-O-S— CI ► RC1 + S0 2 

alkyl chlorosulfite 

Phosphorus tribromide, PBr 3 , is an excellent reagent for converting 
alcohols to bromides. A disadvantage compared to thionyl chloride is the 
formation of involatile P(OH) 3 rather than sulfur dioxide. 

3 ROH + PBr 3 ► 3 RBr + P(OH) 3 



B. ESTERS OF SULFURIC ACID— DEHYDRATION OF ALCOHOLS 

Alkyl hydrogen sulfate formation from alcohols and concentrated sulfuric 
acid may occur by a reaction rather closely related to alkyl halide formation. 



ROH + H 2 S0 4 , ROH 2 + HS0 4 ► ROS0 3 H + H 2 

alkyl hydrogen sulfate 

On heating, alkyl hydrogen sulfates readily undergo elimination of sulfuric 
acid to give alkenes and, in the reaction of an alcohol with hot concentrated 
sulfuric acid, which gives overall dehydration of the alcohol, the hydrogen 
sulfate may well be a key intermediate. This is the reverse of acid-catalyzed 
hydration of alkenes discussed previously (Section 4-4) and goes to comple- 
tion if the alkene is allowed to distill out of the reaction mixture as it is 
formed. 

(- H 2 0) 150° 
CH 3 CH 2 OH + H 2 S0 4 ' CH 3 CH 2 0SO 3 H ^ - CH 2 =CH 2 + H 2 S0 4 

The mechanism of elimination of sulfuric acid from ethyl hydrogen sulfate 
is probably of the E2 type, with water or bisulfate ion acting as the base. 

At lower temperatures, alkyl hydrogen sulfate may react by a displacement 
mechanism with excess alcohol in the reaction mixture with formation of a 
dialkyl ether. Diethyl ether is made commercially by this process. Although 
each step in the reaction is reversible, ether formation can be favored by 
distilling away the ether as fast as it forms. 

H 
-^— - CH 3 CH 2 -0— CH 2 CH 3 + HSO? 



CH 3 CH 2 OCH 2 CH 3 + H 2 SQ 4 



sec 10.S reactions involving the C — O bond of alcohols 257 

Most alcohols will also dehydrate at fairly high temperatures to give 
alkenes and (or) ethers in the presence of solid catalysts such as silica gel or 
aluminum oxide. The behavior of ethanol, which is reasonably typical of 
primary alcohols, is summarized in the following equations : 



Al 2 Q 3 

375° 



- CH,= CH, + H,0 



CH 3 CH 2 OH 



\ AUO3 | 
300° 



CH 3 CH 2 OCH 2 CH 3 + H 2 



Tertiary alcohols react with sulfuric acid at much lower temperatures than 
do most primary alcohols. The S N 1 and El reactions in Scheme I may be 
written for r-butyl alcohol and sulfuric acid. Di-f-butyl ether is unstable in 



I " H 2 S0 4 I 

CH3-C-OH . * CH 3 -C-OH 2 



CH, 



CH, 



-H 2 




HSO4 

H " <-«, CH, 

Is 
(CH^C-O— C(CH 3 ) 3 CH 3 -C = CH 2 CH 3 -C-OS0 3 H 

CH 3 



-H® 



(CH 3 ) 3 C-0-C(CH 3 ) 3 



polymer 
SCHEME I 



sulfuric acid solution and it has never been detected in reaction mixtures of 
this type. Its low stability may be due to steric crowding between the alkyl 
groups. 



CH, 



CH 



/* 



\/y? H 



dH ^ c ^H 3 

1 

CH, 



steric repulsions 



2-Methylpropene can be removed from the reaction mixture by distillation 
and is easily made the principal product by appropriate adjustment of the 



chap 10 alcohols and ethers 258 

reaction conditions. If the 2-methylpropene is not removed as it is formed, 
polymer is the most important end product. Sulfuric acid is often an unduly 
strenuous reagent for dehydration of tertiary alcohols. Potassium hydrogen 
sulfate, copper sulfate, iodine, phosphoric acid, or phosphorus pentoxide may 
give better results by causing less polymerization and less oxidative degrada- 
tion. The oxidizing action of sulfuric acid results in formation of sulfur 
dioxide. 

Rearrangement of the alkyl group of an alcohol is very common in dehydra- 
tion, particularly in the presence of sulfuric acid, which is highly conducive 
to carbonium ion formation. Shown are typical examples of both methyl and 
hydrogen migration. The key step in each such rearrangement involves an 
isomerization of a carbonium ion along lines discussed in Section 8T3. 



CH 3 
I 
CH 3 -C— CH-CH 3 



CH 3 

I 
C- 

I 
H 



CH 3 ,CH 3 

H 2 SO* \„ „/ 



C=C + H 2 

CH 3 CH 3 

CH, 



H 2 SO» 



Except in a few circumstances where thermodynamic control dominates and 
leads to different results from kinetic control, the final products always cor- 
respond to rearrangement of a less stable carbonium ion to a more stable 
carbonium ion. 



reactants 



R 

I e 
— C—C- 

I I 

less stable 
carbonium ion 



R 

e I 
-C— C— 

I I 

more stable 
carbonium ion 



products 



For the particular case of the dehydration of methyl-?-butylcarbinol, the 
sequence is 



CH 3 

I H 2 S0 4 

CH 3 — C— CH — CH 3 ;= 



(-OH e ) 



H,C OH 



CH 3 

CH 3 — C-CH-CH 3 

I e 

CH 3 

secondary 
carbonium ion 



I I 
-+ CH 3 — C — C— CH 3 

© I 
H 



tertiary 
carbonium ion 



-H« 



CH 3 CH 3 

\ / 

c=c 

/ \ 

CH 3 CH 3 



sec 10.6 oxidation of alcohols 259 



10-6 oxidation of alcohols 



Primary alcohols on oxidation first give aldehydes and thence carboxylic 
acids, whereas secondary alcohols give ketones. 



o ,o 

[O] S [O] # 

RCH 2 OH > R-C > R-C 

H OH 

R R 

\ [O] \ 

CHOH ► / C= ° 

R R 

Tertiary alcohols are oxidized with considerable difficulty and then only 
with degradation into smaller fragments by cleavage of carbon-carbon bonds. 
Similarly, oxidation of secondary alcohols beyond the ketone stage does not 
proceed readily and leads to degradation. 

Laboratory oxidation of alcohols is most often carried out with chromic 
acid (H 2 Cr0 4 ), which is usually prepared as required from chromic oxide 
(Cr0 3 ) or from sodium dichromate (Na 2 Cr 2 7 ) in combination with sulfuric 
acid. Acetic acid is the most generally useful solvent for such reactions. 



3 CH-OH + 2 H 2 Cr0 4 + 6 H® ► 3 C=0 + 2 Cr 3 ® + 8 H 2 

/ / 

CH 3 CH 3 

The mechanism of the chromic acid oxidation of isopropyl alcohol to 
acetone has been investigated very thoroughly and is highly interesting in that 
it reveals how changes of oxidation level can occur involving a typical inorgan- 
ic and a typical organic compound. The initial step is reversible formation of 
isopropyl hydrogen chromate [7], which is quite unstable and is not usually 
isolated (although it can be isolated by working rapidly at low temperatures). 

CH 3 CH 3 

X CH-OH + H 2 Cr0 4 . H 2 + \:H—0— Cr0 3 H 

CH 3 CH 3 

CH 3 

X CH-0— CrOjHf 
CH 3 

The subsequent step is the slowest in the reaction and appears to involve 
either (a) a cyclic decomposition of the protonated ester or (b) attack of a 
base (water) at the a, hydrogen of the protonated chromate ester concurrent 
with elimination of the grouping H 2 Cr0 3 , for which there is an obvious 
analogy with an E2 reaction (Section 8T2). 



chap 10 alcohols and ethers 260 



CH < / H I „ . CH \ / H -9 



(a) C CrO z Hf ► C li 



CH 
CH 3 



C— O + H 2 Cr0 3 + H a 



H 

CH 3 H<-:(/ ^ CH 3 

(b) / x ) ^ H ► JZ-0 + H 3 + H 2 Cr0 3 

CH 3 O— Cr0 3 H? CH 3 

Despite intensive study, the exact mechanism of decomposition of the 
chromate ester remains in doubt. The Cr IV shown as the reduction product 
(H 2 Cr0 3 ) in both (a) and (b) is a highly reactive oxidant and is rapidly con- 
verted to Cr m in subsequent steps. 

The Cr VI oxidation of alcohols is acid catalyzed, as shown above. Indeed, 
neutral and basic solutions of chromate are without any effect on alcohols. 
Permanganate oxidation of alcohols, on the other hand, is subject to both 
acid and base catalysis. The acid catalysis is the result of the conversion of 
MnOf to HMn0 4 , a more powerful oxidant, and requires a high concentra- 
tion of acid. The base catalysis, however, is the result of converting the 
neutral alcohol to the more easily oxidized alkoxide ion : 

R H R H 

X C 7 + OH e < \ + H 2 

R y X OH R 7 V 

The rate-controlling step of this reaction is transfer of hydrogen, possibly as 
hydride ion, to the Mn v ". 

R (^rP^M R 

X C ^ + MnOe > C=0 + HMn0 4 2e 

R X N > * r 7 

The ion UMn0 2 4 e (Mn v ) quickly disproportionates to Mn IV and Mn v ". 

Presumably Cr VI is ineffective in oxidizing alcohols by such a path because 
in basic solution the chromate ion, CrO^ 8 , carries a double negative charge 
and the repulsions between the reactants would be too great. 

Biological oxidation of alcohols is considered later in the book (Section 
18-4). 



10 -1 polyhydroxj alcohols 



The simplest example of an alcohol with more than one hydroxyl group is 
methylene glycol, HOCH 2 OH, the term " glycol " meaning a diol, a substance 



sec 10.7 polyhydroxy alcohols 261 

with two alcoholic hydroxyl groups. Methylene glycol is reasonably stable in 
water solution but attempts to isolate it lead only to its dehydration product, 
formaldehyde. 



HO-CH,-OH 



This behavior is typical of gem-diols (gem = geminal, i.e., with both hydroxyl 
groups on the same carbon atom), and the very few gem-diols that are 
isolable are those which carry strongly electron-attracting substituents, such 
as chloral hydrate and hexafluoroacetone hydrate. 

OH OH 

I I 

CI3C-C-OH CF3-C-CF3 
I I 

H OH 

chloral hydrate hexafluoroacetone hydrate 

Polyhydroxy alcohols in which the hydroxyl groups are situated on different 
carbons are relatively stable and, as we might expect for substances with a 
multiplicity of hydrogen bonding groups, they have high boiling points, 
considerable water solubility, and low solubility in nonpolar solvents. 

CH 2 CH 2 CH 2 CH 2 CH 2 CH 2 CH 2 -CH-CH 2 

II II III 

OH OH OH OH OH OH OH 

ethylene glycol tetramethylene glycol glycerol 

1 , 2-ethanediol 1 , 4-butanediol 1,2, 3-propanetrioI 

bp 197° bp 230° bp 290° 

1,2-Diols are usually prepared from alkenes by oxidation with reagents 
such as osmium tetroxide, potassium permanganate, or hydrogen peroxide 
(Section 4-4G). However, ethylene glycol is made on a large scale commer- 
cially from ethylene oxide, which in turn is made by air oxidation of ethene 
at high temperatures over a silver catalyst. 



CH 2 =CH 2 + 10 2 -^jU CH 2 -CH 2 H2 °' HS .. 

O OH OH 

ethylene oxide 



Ethylene glycol has important commercial uses. It is an excellent per- 
manent antifreeze for automotive cooling systems, because it is miscible with 
water in all proportions anda 50% solution freezes at -34° ( — 29°F). It is 
also used as a solvent and as ah-- intermediate in the production of polymers 
(polyesters) and other products, as shown in Section 28-4. 

The trihydroxy alcohol, glycerol, is a nontoxic, water-soluble, viscous, 
hygroscopic liquid that is widely used as a humectant (moistening agent). It 
is an important component of many food, cosmetic, and pharmaceutical 
preparations. At one time, glycerol was obtained on a commercial scale only 



chap 10 alcohols and ethers 262 

as a byproduct of soap manufacture through hydrolysis of fats, which are 
glycerol esters of long-chain alkanoic acids (see Chapter 13), but now the 
main source is synthesis from propene as described in Section 9-6. The 
trinitrate ester of glycerol (nitroglycerin) is an important but shock-sensitive 
explosive. Dynamite is a much safer and more controllable explosive, made 
by absorbing nitroglycerin in porous material like sawdust or diatomaceous 
earth. Smokeless powder is nitroglycerin mixed with partially nitrated 
cellulose. 



CH 2 ON0 2 

I nitroglycerin 

CHON0 2 (glyceryl trinitrate) 

CH,ONO, 



Glycerol plays an important role in animal metabolism as a constituent of 
fats and lipids. 



10-8 unsaturated alcohols 

The simplest unsaturated alcohol, vinyl alcohol, is unstable with respect to 
acetaldehyde and has never been isolated. Other simple vinyl alcohols undergo 

o 

// 

CH 2 =CHOH ^_ ' CH 3 -C 

H 

vinyl alcohol acetaldehyde 

similar transformations to carbonyl compounds. However, ether and ester 
derivatives of vinyl alcohols are known and the esters are used to make many 
commercially important polymers. 

Allyl alcohol, CH 2 =CH— CH 2 OH, unlike vinyl alcohol, is a stable com- 
pound. It displays the usual double-bond and alcohol reactions but, as expec- 
ted from the behavior of allylic halides (Section 9-6), it is much more reactive 
than saturated primary alcohols toward reagents such as Lucas reagent that 
cleave the C— O bond. 



ethers 

Substitution of both hydrogens in water by alkyl or similar groups gives 
compounds known as ethers, general formula R— O — R. Ethers thus lack the 



sec 10.9 preparation of ethers 263 

hydroxyl group which determines to such a great extent the physical and 
chemical properties of water and alcohols. Ethers are much more volatile than 
alcohols of the same molecular weight. Thus, diethyl ether (C 2 H 5 OC 2 H 5 ) 
boils at 35° (body temperature), which is 48° to 83° below the boiling points 
of the isomeric butyl alcohols. Its solubility in water, however, is 7 g per 100 g 
of water, about the same as «-butyl alcohol (but much less than that of 
(-butyl alcohol which is miscible with water). Diethyl ether is often used to 
extract organic substances out of water solution and it should be remembered 
that a considerable amount of ether can be retained in the water layer, which 
may release enough ether vapor to create a fire hazard. 

A number of important ethers are listed here. (Nomenclature of ethers was 
discussed in Section 8-3.) 

CH 3 CH 2 OCH 2 CH 3 CH 3 OCH=CH 2 H 2 C CH 2 HOCH 2 CH 2 OCH 2 CH 2 OH 

O 



diethyl ether 
bp35° 


methyl vinyl ether ethylene oxide 
bp 8° bpll" 


diethylene glycol 
bp 245° 


^}-OCH3 




H 2 C-CH 2 

H 2 C V /CH 2 
O 


H 2 C^ X CH 2 
H 2 C V .CH 2 


ethyl phenyl ether 
(anisole) 
bp 155° 




tetrahydrofuran 
bp65° 


1, 4-dioxane 
bp 101.5° 



10-9 preparation of ethers 



There are only two generally useful ways of preparing ethers and these have 
been previously discussed in connection with the reactions of alcohols. 
The first is the Williamson synthesis (Section 10-4B), to which the general 
rules governing S N 2 displacements and the competing elimination reactions 
apply (Sections 8-7, 8-12). 



R-X + R'CTNa 95 



-► R— O— R + Na* 



The second is dehydration of alcohols (Section 10- 5B), which can be accom- 



2 ROH 



-H 2 



-» R— O— R 



plished with concentrated H 2 S0 4 , H 3 P0 4 , or at high temperatures with 
A1 2 3 catalysis. Only symmetrical ethers can be made efficiently by this 
route; intramolecular dehydration to give alkenes is a competing reaction. 



chap 10 alcohols and ethers 264 

10-10 reactions of ethers 

In general, ethers are low on the scale of chemical reactivity, since the 
carbon-oxygen bond is not cleaved readily. For this reason, ethers are fre- 
quently employed as inert solvents in organic synthesis. Particularly important 
in this connection are diethyl ether, diisopropyl ether, tetrahydrofuran, and 
1,4-dioxane. The latter two compounds are both miscible with water. 

The mono and dialkyl ethers of ethylene glycol and diethylene glycol are 
useful high-boiling solvents. Unfortunately, they have acquired irrational 
names like " polyglymes," "cellosolves," and "carbitols"; for reference, 
cellosolves are monoalkyl ethers of ethylene glycol ; carbitols are monoalkyl 
ethers of diethylene glycol; polyglymes are dimethyl ethers of di- or triethyl- 
ene glycol, diglyme, and triglyme, respectively. 

CH 3 OCH 2 CH 2 OH C 4 H 9 OCH 2 CH 2 OCH 2 CH 2 OH 

methyl cellosolve butyl carbitol 

bp 124° bp231° 

CH 3 OCH 2 CH 2 OCH 2 CH 2 OCH 3 

diglyme 
bp 161° 

Unlike alcohols, ethers are not acidic and do not usually react with bases. 
However, exceptionally powerfully basic reagents, particularly certain 
alkali-metal alkyls, will react destructively with many ethers (see also Sec- 
tion 9-9B). 



CHf Na a> + H:CH 2 



CH 4 + CH 2 =CH 2 + CH 3 CH 2 ONa 



Ethers, like alcohols, are weakly basic and are converted to oxonium ions 
by strong acids (e.g., H 2 S0 4 , HC10 4 , and HBr) and to coordination com- 
plexes with Lewis acids (e.g., BF 3 and RMgX). 



C 2 H 5 6C 2 H 5 + HBr . C 2 H 5 OC 2 H 5 + Br e 

diethyloxonium bromide 

F F F 

\l/ 
B 

C 2 H 5 OC 2 H 5 + BF 3 ► C 2 H 5 OC 2 H 5 

boron trifluoride etherate 

Dialkyloxonium ions are susceptible to nucleophilic displacement and 
elimination reactions just as are the conjugate acids of alcohols (Section 
10-5A). Acidic conditions, therefore, are used to cleave ethers, and as appro- 
priate to the degree of substitution either S N 2 or S N 1 (El) reactions may 
occur. 



C,H,-0-C 2 H 5 + H s 



H 

I 



sec 10.11 cyclic ethers 265 



Br 9 



C 2 H 5 -0— C 2 H 5 „ . > C 2 H s Br + C 2 H 5 OH 



H 



(CH 3 ) 3 C-0-C(CH 3 ) 3 + H® . (CH 3 ) 3 C-0-C(CH 3 ) 3 



HS0 4 



El 



2(CH 3 ) 2 C=CH 2 «- 



(CH 3 ) 3 C-OH + (CH 3 ) 2 C=CH 2 + H 2 S0 4 



Ethers are quite susceptible to attack by radicals, and for this reason are not 
good solvents for radical-type reactions. In fact, ethers are potentially 
hazardous chemicals since, in the presence of atmospheric oxygen, a radical- 
chain process can occur, resulting in the formation of peroxides which are 
unstable, nonvolatile explosion-prone compounds. This process is called 
autoxidation and occurs not only with ethers but with many aldehydes and 
hydrocarbons. Commonly used ethers such as diethyl ether, diisopropyl 
ether, tetrahydrofuran, and dioxane often become contaminated with perox- 
ides formed by autoxidation on prolonged storage and exposure to air and 
light. Purification of ethers is frequently necessary before use, and caution 
should always be exercised in their distillation as the distillation residues may 
contain dangerously high concentrations of explosive peroxides. 



10-11 cyclic ethers 



Ethylene oxide, the simplest cyclic ether, is an outstanding exception to the 
generalization that most ethers resist cleavage. Like that of cyclopropane, the 
three-membered ring of ethylene oxide is highly strained and opens readily 
under mild conditions. Unlike ordinary ethers it reacts with Grignard 
reagents (Section 9-9C), a useful way of extending a chain by two carbon 
atoms. For example, conversion of 1-hexanol to 1-octanol can be easily 
accomplished this way via the alkyl bromide. 



CH 3 CH 2 CH 2 CH 2 CH 2 CH 2 OH + HBr 



-> CH 3 CH 2 CH,CH 2 CH 2 CH 2 Br + H,0 



CH 3 CH 2 CH 2 CH 2 CH 2 CH 2 Br + Mg dry ethet > CH 3 CH 2 CH 2 CH 2 CH 2 CH 2 MgBr 
CH 3 CH 2 CH 2 CH 2 CH 2 CH 2 MgBr + CH 2 -CH 2 

O 



CH 3 CH 2 CH 2 CH 2 CH 2 CH 2 CH 2 CH 2 MgBr 
H 2 
CH 3 CH 2 CH 2 CH 2 CH 2 CH 2 CH 2 CH 2 OH + HOMgBr 



' chap 10 alcohols and ethers 266 

The commercial importance of ethylene oxide lies in its readiness to form 
other important compounds; for example, ethylene glycol, diethylene 
glycol, the cellosolves and carbitols, dioxane, ethylene chlorohydrin, and 
polymers (Carbowax) (see Scheme II). 




H2 ° HOCH 2 CH 2 OH -5— -> (HOCH 2 CH 2 ) 2 



\V 

o 



l> 
H,C 



H® 




ethylene 
glycol 



methyl 
cellosolve 



H 2 C— CH 2 

H — > O O dioxane 

\ / 
H 2 C-CH 2 



ethylene 
* ClCH 2 CH 2 OH chlorohydrin 

> H-0-c-CH 2 -CH 2 — 0-);H 
SCHEME II 



diethylene 
glycol 

CH 3 (CH 2 ) 2 (CH 2 ) 2 OH 

methyl 
carbitol 



Carbowax 



The lesser-known four-membered cyclic ether, trimethylene oxide (CH 2 ) 3 0, 
is also cleaved readily, but less so than ethylene oxide. Tetramethylene oxide 
(tetrahydrofuran) is relatively stable and is a water-miscible compound with 
desirable properties as an organic solvent. It is often used in place of diethyl 
ether in Grignard reactions and reductions with lithium aluminum hydride. 



summary 

Alcohols, whether primary (RCH 2 OH), secondary (R 2 CHOH), or tertiary 
(R 3 COH), have higher melting and boiling points and much higher water 
solubilities than the corresponding hydrocarbons. The highest melting points 
and water solubilities (but not boiling points) belong to tertiary and other 
branched-chain alcohols. 

The H— O stretching frequencies in the infrared spectra of alcohols occur 
near 3700 cm -1 (free OH) and near 3350 cm" 1 (hydrogen-bonded OH). 
The nmr peaks of hydrogen-bonded OH protons occur 4-5 ppm downfield 
from TMS but shift upheld with dilution. 

Alcohols can be prepared in the following ways for a set of typical 
primary, secondary, and tertiary alcohols. 



exercises 267 



RCH,C0 2 R' RCHXHXI 



RCHXHO > RCH,CH 2 OH «- 



RCH 2 MgBr 



RCHCICHj 



O 



RCH=CH 2 ► RCHOHCH., < RC-CH 3 



RMgBr > R 2 COHCH 3 



R,C=CH 2 



R,CC1CH, 



The reactions of alcohols include the following: 



ROH 



RO e 
ROH 2 ® 
ROR' 
R'C0 2 R 

RX 

ROSO3H- 

(or R®)< 



(with powerful bases) 

(with strong acids) 

(Williamson synthesis; S N 2 mechanism) 

(with R'COCl or with R'C0 2 H; addition-elimination 
mechanism) 

(with HX; S N 1 or S N 2 mechanism) 

— > symmetrical ether 

, ^alkene - » polymer 

rearranged products 



On oxidation, primary alcohols give aldehydes and secondary alcohols give 
ketones. Tertiary alcohols are oxidized only with degradation. The Cr VI 
oxidation of alcohols proceeds via a chromate ester, ROCr0 3 H, while the 
Mn™ oxidation involves the anion of the alcohol, RO e . 

Important polyhydric or unsaturated alcohols include chloral hydrate, 
ethylene glycol, glycerol, and allyl alcohol. 

Ethers are less soluble and are more volatile than alcohols of the same 
molecular weight. They are prepared from alcohols by dehydration or by the 
Williamson synthesis (see above). Ethers are cleaved by strong acids, 
ROR + HX ->■ RX + ROH, and are susceptible to attack by radicals and 
exceptionally powerful bases (R e ). 

Some important cyclic ethers include 1,4-dioxane, tetrahydrofuran, and 
ethylene oxide ; the latter is not a typical ether because it reacts with Grignard 
reagents. 



exercises 

10-1 c«-l,2-Cyclopentanediol is appreciably more volatile (bp 124° at 29 mm) than 
frww-l,2-cyclopentanediol (bp 136° at 22 mm). Explain. 

10-2 Suggest a likely structure for the compound of molecular formula C 4 H e O 
whose nmr and infrared spectra are shown in Figure 10-3a. Show your 
reasoning. Do the same for the compound of formula C 3 Hs0 2 whose spectra 
are shown in Figure 10-3b. 










UOISSIUISUBI} % 








T 


r 


^l^^^^^^i 


-^---^-h-----^ 1213 * 




o 
o 




















f 


^^sSi^4l^l|liil^^^(i 










•3 




*- — 







o 
o 




°* ii 




fflmSsKt 




■ 


o 








s:^ 




■* . 












J 5 














a. 




r 


^^^^^i|^^^^^^|| 




1 


JS 










m ° 














C 




\ 






till -o >» 






J 






; c 






1 








> 




f 






1 = 


S 











o a 1 
Jo -j 




„|i 


i 

\ 




d 


Jg 

o 
o 






7**~ 










-*■ b 


^ 


c; 




lift o 

Jo 

'.J 
























H$s||« 




























li 


C 






filo 
$m o 

r 



UOISSIUISUBIJ % 




Figure 10-3 Nuclear magnetic resonance spectra and infrared spectra of 
C 4 H 6 (a) and C 3 H 8 2 (b); see Exercise 10-2. 



268 



exercises 269 

10-3 What type of infrared absorption bands due to hydroxyl groups would you 
expect for ?ra«.y-l,2-cyclobutanediol and 1 ,2-butanediol (a) in very dilute 
solution, (b) in moderately concentrated solution, and (c) as pure liquids ? 



10-4 Show the reagents and conditions for the reactions used to prepare alcohols 
that are summarized on p. 267. 



10-5 Show how you can convert 1-phenylethanol to 2-phenylethanol and 2-phenyl- 
ethanol to 1-phenylethanol. 



10-6 Show how you can prepare 2-methyl-2-butanol starting with (a) 2-butanol, 
(b) 2-propanol. 

10-7 In the esterification of an acid with an alcohol, how could you distinguish 
between C— O and O— H cleavage of the alcohol using heavy oxygen ( ls O) 
as a tracer? 

O O 

i II ! H ® II 

CH3O-T-H + RCi-OH ► RCOCH3 + H 2 

or 

O O 

i II ! h« II 

CH 3 -rOH + RCO-i-H ► RCOCHj + H 2 



What type of alcohols would be most likely to react by C— O cleavage, and 
what side reactions would you anticipate for such alcohols? 

10-8 Predict the major products of the following reactions : 

a. (CH 3 ) 3 CCH 2 1 + C 2 H 5 O e > 

b. (CH 3 ) 3 CBr + CH 3 O e ► 

c. I Vci + (CH 3 ) 3 CO e ► 






d. L V-CH 2 Br + (CH 3 ) 2 CHCH,O e 



10-9 Predict the products likely to be formed on cleavage of the following ethers 
with hydrobromic acid: 

XH 2 

a. CH 2 =CH-CH 2 -0-CH 3 H 2 C^ \ 2 

d. I O 

b. CH 3 CH 2 -0-CH = CH 2 2 ^CH 2 

c. (CH 3 ) 3 CCH 2 -0-CH 3 e . ^ r A_ -CH 3 



chap 10 alcohols and ethers 270 

10-10 What would be the products of the following reactions? Indicate your 
reasoning. 



a. (CH 3 ) 3 COH + CH 3 OH COnC - H2S °' 

H H 

I I 

b. H-C-C-H . 



I I 75%H 2 S0 4 



I 1 100° 

HO OH 

CH 3 

I 
c. CH,-C— CH,-OH 



95%H 2 SO« 

> 

25° 
100% H 2 S0 4 



CH, 25° 

cone. H 2 S0 4 



I I 

d. CH 3 — C-O-C-CH3 
I I 

CH 3 CH 3 



0° 



10-11 The reaction of methyl acetate with water to give methanol and acetic acid 
is catalyzed by strong mineral acids such as sulfuric acid. Furthermore, when 
hydrolysis is carried out in ls O water, the following exchange takes place 
faster than formation of methanol. 



O ,S P 

f H® // 

CH 3 -C + H 2 ls O 7==t CH 3 -C x + H 2 

X OCH 3 O-CH3 

No methanol- 1 s O (CH3 1 8 OH) is formed in hydrolysis under these conditions. 

a. Write a stepwise mechanism which is in harmony with the acid catalysis 
and with the results obtained in ls O water. Mark the steps of the reac- 
tion that are indicated to be fast or slow. 

b. The reaction depends on methyl acetate having a proton-accepting 
ability comparable to that of water. Why? Consider different ways of 
adding a proton to methyl acetate and decide which is most favorable 
on the basis of structural theory. Give your reasoning. 

c. Explain how the reaction could be slowed down in the presence of high 
concentrations of sulfuric acid. 

10-12 Indicate how you would synthesize each of the following substances from 
the given organic starting materials and other necessary organic or inorganic 
reagents. Specify reagents and conditions. 

a. CH 3 CH 2 CH 2 C(CH 3 ) 2 C1 from CH 3 CH 2 CH 2 OH 

b. CH 3 CH 2 CHCH 2 CH 3 from CH 3 CH 2 OH 

I 

O-C-CH3 
II 
O 

c. (CH 3 ) 2 CH-CH 2 Br from (CH 3 ) 3 COH 



exercises 271 



d. CH 3 CH 2 CHCH 3 from CH 3 CH 2 CH 2 CH 2 OH 

I 
OSO3H 

e. CH 3 CH 2 C(CH 3 ) 2 CH0 from (CH 3 ) 2 C(OH)CH 2 CH 3 
/ CH 3 OCH 2 CH 2 OCH 3 from ethene 




from 



H 3 C 




OH 



(cis) 
CH, 



(trans) 
CH 3 



h. CH 3 — C— CH 2 — CH 3 from CH 3 — C-OH 



CH, 



/. (CH 2 =CHCH 2 ) 2 



J. rf 



,CH 3 



OCH 2 CH 3 
CH, 




CH 3 

=Cr 

CH 2 
from I I 

kCH, 



OCH, 



H 



from 



AT 



10-13 Give for each of the following pairs of compounds a chemical test, preferably 
a test tube reaction, which will distinguish between the two substances. 
Write an equation for each reaction. 



CH, 



I 
H 

and CH 3 CH=CH-CH 2 OH 

CH 3 
I 
and CH 3 -CH-CH 2 CH 2 OH 



CH 3 
I 

a. CH 3 -C-CH 3 

I 
OH 

b. CH 2 =CH-CH 2 CH 2 OH 

CH 3 
I 

c. CH 3 -C-CH 2 OH 

I 
CH 3 

d. CH 3 CH 2 -0-S0 2 -0-CH 2 CH 3 and CH 3 CH 2 CH 2 CH 2 -0-S0 3 H 

o o 

II II 

e. CH 3 -C-CI and C1CH 2 C-0H 

18 o o 

II II 

/. CH 3 -C-0CH 3 and CH 3 -C- 18 0CH 3 

CH 3 CH 3 



g. CH 3 -C-0-Cr0 3 H 



CH, 



and CH 3 -C— CH 2 — O— Cr0 3 H 
H 



chap 10 alcohols and ethers 272 



h. CH 2 -CH-CH 3 and CH 2 -CH 2 -CH 2 

II I I 

OH OH OH OH 



H,C O and CH 3 CH-CH 2 

\ / \ / 

CH 2 V 



I 3 ' I J I J T T 

CH 3 -C-CH 2 -0-CH 2 -C-CH 3 and CH 3 -C-0-CH 2 -CH 2 -C-CH 3 
CH 3 CH, CH 3 CH 3 



10-14 Suppose you were given unlabeled bottles, each of which is known to contain 
one of the following compounds: 1-pentanol, 2-pentanol, 2-methyl-2- 
butanol, 3-penten-l-ol, 4-pentyn-l-ol, di-«-butyl ether, and 1-pentyl 
acetate. Explain how you could use simple chemical tests (test tube reactions 
with a visible result) to identify the contents of each bottle. 

10-15 Predict the principal features with approximate chemical shifts in the nmr 
spectra of the following compounds (neat liquid unless otherwise noted) : 

a. isobutyl alcohol 

b. neopentyl alcohol 

c. methyl /3-methylvinyl ether 

d. /-butyl alcohol 

e. f-butyl alcohol in carbon tetrachloride 

10-16 Sketch out the energy profile for the reaction of CH 3 COCl with CH 3 OH as 
discussed in Section 10-4C, following the arguments in Section 8-9. It will 
help, in this connection, if you calculate Aif for the addition of ROH to a 
carbon-oxygen double bond to estimate the energy of [2] in Section 10-4C. 






llf 

ear uSvs ■fSss 

illil 









- 








ffW. 



BOB 














i./.; -f|- ..--': ;%..•' &! 



:i'AjrtW^ V ft* CSCfari^WV *g%-aT8i«Wa M'^^f.'.-t £S5 vig^iV'Vv 






ff ^iiC**'^^ 



i ••.frf--',lV 

! •* 



¥$$&*:& j 



^^v-^ 



M^^"_^i- '♦•■^ 



F? 1 "??!;' ;"'rit ;^"5 






liip 





f i*"*- '• 


















*W' 






i0 i ~S& 


*;i 


f. 




>*C"'-i'r .* 


\Vj, 


"-^ts?; vv^-V»r].*' *iv 


rivo 


Si* 


;•>;?■; 


* ',<"" < 




fc.r** 1 '! ** 


rt'i-' 






■ ■_ ■. ' 


AiV 










•V? 


* -£',' --" 


;£& 


t^f 


**T* 


i-/-Z*&< 









chap 11 aldehydes and ketones I 275 

o 

II 

The carbonyl group, — C— , is an extremely important functional group, and 
indeed the chemistry of carbonyl compounds is virtually the backbone of 
synthetic organic chemistry. We shall divide our study of these substances 
into three parts. In this chapter we shall consider first methods for the 
synthesis of simple aldehydes and ketones and then the reactions of aldehydes 
and ketones which involve only their carbonyl groups. In Chapter 12, con- 
sideration will be given to the way in which the carbonyl function activates 
groups on adjacent carbons. In Chapter 13, we shall discuss the role of the 
carbonyl group in reactions of carboxylic acids and their derivatives. Through- 
out these discussions attention will be given to the differences in behavior of 
various kinds of carbonyl groups — differences that may be correlated with 
electrical and steric effects. 



11' 1 nomenclature of aldehydes and ketones 

o o 

Aldehydes have the general formula R— C and ketones R— C— R. 

H 
Since their nomenclature follows along the lines discussed previously for 
other types of compounds, the most widely used naming systems are sum- 
marized in Table 11-1. 



11 -2 carbonyl groups of aldehydes and ketones 



A. COMPARISON WITH CARBON-CARBON DOUBLE BONDS 

The carbon-oxygen double bond is both a strong and a reactive bond. Its 
bond energy (179 kcal) is rather more than that of two carbon-oxygen single 
bonds (2 x 86 kcal) in contrast to the carbon-carbon double bond (146 
kcal), which is weaker than two carbon-carbon single bonds (2 x 83 kcal). A 
typical difference in reactivity is seen in hydration : 

CH 2 = + H 2 , CH 2 -0 

I 



OH H 



CH,=CH, + H,0 



I ' I 
OH H 



Formaldehyde adds water rapidly and reversibly at room temperature 
without an added catalyst; but the addition of water to ethene does not occur 
in the absence of a very strongly acidic catalyst, even though the equilibrium 
constant is considerably larger. 

The reactivity of the carbonyl bond is primarily due to the difference in 
electronegativity between carbon and oxygen, which leads to a considerable 
contribution of the dipolar resonance form with oxygen negative and carbon 
positive. 



chap 1 1 aldehydes and ketones I 276 



Table 11-1 Nomenclature systems for carbonyl compounds 





P 
Aldehydes, R— C 






H 




formula 


IUPAC name = longest 
straight chain" -\- suffix -al 


name as derivative of a 
carboxylic acid (i.e., 
a carboxaldehyde) 


H 2 C=0 


methanal 


formaldehyde 


ClCH 2 CH 2 CHO 


3-ohloropropanal 


/3-chloropropionaIdehyde 


CH 3 

| 






CH 2 =CCHO 


2-methylpropenal 


methacrolein 
(methacrylaldehyde) 1 ' 


[^CHO 


cyclopropylmethanal 


cyclopropanecarboxal- 
dehyde c 




R 






Ketones, p = ° 
R' 








formula 


IUPAC name = longest 
straight chain" -\- suffix -one 


name as ketone 11 


(CH 3 ) 2 C=0 


propanone 


dimethyl ketone 
(acetone) 8 


O 
II 
CH 3 CCH 2 CH=CH 2 


4-penten-2-one 


methyl allyl ketone 


O 






Q4-CH S 


phenylethanone 


methyl phenyl ketone 
(acetophenone)" 


o 


cyclohexanone 


— 



" The longest straight chain which includes the functional group is understood. 

b Propenoic acid is commonly called acrylic acid and 2-methylpropenoic acid methacrylic acid. 

c The acid (CH 2 ) 2 CHC0 2 H is commonly called cyclopropanecarboxylic acid. 

d Ketone is a separate word even though aldehyde is not. Thus (CH 3 CH 2 ) 2 CO is diethyl 
ketone but (CH 3 CH 2 ) 2 CHCHO is diethylacetaldehyde (see Section 8-6). 

e A few ketones are named as derivatives of carboxylic acids. Names by this system stem 
from the synthesis (real or imagined) of the ketone by the reaction RC0 2 H-f R'C0 2 H-> 
RCOR'+C0 2 + H 2 0. 



sec 11.2 carbonyl groups of aldehydes and ketones 277 

\ \<b P. \t& se 

C=0 < ► -C-O: ~ C=~0 

The polarity of the carbonyl bond is expected to facilitate addition of water 

0® se ae a® 
and other polar reagents such as H— X and R— MgX relative to addition of 
the same reagents to alkene double bonds. However, we must always keep in 
mind the possibility that, whereas additions to carbonyl groups may be 
rapid, their equilibrium constants may be small, because of the strength of 
the carbonyl bond. 

The polarity of the carbonyl group is manifested in many of the other 
properties of aldehydes and ketones. Boiling points for the lower members of 
the series are 50° to 80° higher than for hydrocarbons of the same molecular 
weight. The water solubility of the lower-molecular-weight aldehydes and 
ketones is large. 

O 

II 
Formaldehyde, H— C— H, the first member of the aldehyde series, is a gas, 

bp —21°. It is readily soluble in water, forming the hydrate (Section 10-7). 

O 

// 
The second member, acetaldehyde, CH 3 — C , has a boiling point near 

H 

O 

II 
room temperature (21°). Acetone, CH 3 — C— CH 3 , the first member of the 

ketone series, boils at 56°. It is completely miscible both with water and with 

benzene. As might be expected, the higher aldehydes and ketones are insoluble 

in water. 



B. SPECTROSCOPIC PROPERTIES 

The infrared stretching frequencies for the carbonyl groups of aldehydes 
and ketones generally fall between 1705 and 1740 cm -1 and the absorption 
intensities are much greater than for carbon-carbon double bonds. 

All carbonyl compounds have absorption in the accessible part of the 
ultraviolet spectrum. This absorption, which is weak for ordinary carbonyl 

O 

II 
groups such as that in methyl ethyl ketone, CH 3 — C— CH 2 CH 3 (Figure 11-1), 

is designated as an n -> n* transition, meaning that one of the nonbonding 

electrons on oxygen is elevated to an antibonding n* orbital. 

Conjugation with a carbon-carbon double bond has a large effect on the 

spectrum of the carbonyl group. The spectrum of methyl vinyl ketone, 

O 

II 
CH 3 — C— CH=CH 2 , shown in Figure 11-1 reveals a significant shift 

toward the visible. (The tail of the band just reaches the visible part of the 

spectrum and imparts a slight yellow cast to the compound.) This means that 

less energy is required to excite a nonbonding electron in the conjugated 

ketone than in the saturated ketone. We can understand this behavior on the 



chap 1 1 aldehydes and ketones I 278 




2500 3000 

wavelength, A 



Figure 11*1 Ultraviolet spectra of methyl ethyl ketone, a, and methyl vinyl 
ketone, b, in cyclohexane solution. 



basis of differences in the degree of resonance stabilization of the excited states 
of the two kinds of compounds. 

More significant in practical terms is the appearance in the accessible part 
of the spectrum of methyl vinyl ketone of an intense absorption band that is 
due to a n -* n* transition. The almost vertical rise of the absorption curve in 
Figure 11-1 is caused by this intense absorption, whose peak is at 2190 A. A 
comparable band for the carbonyl group of methyl ethyl ketone is not shown 
in Figure 11-1 because it occurs at much shorter wavelength, about 1850 A, 
which is out of the range of most spectrometers. 

The character of the carbonyl bond is such as to give very low-field nmr 
absorptions for protons of the aldehyde group (RCHO). As Table 7-2 shows, 
these absorptions come some 4 ppm below vinylic hydrogens. Much of this 
difference is probably due to the polarity of the carbonyl group. It is carried 
over in much smaller degree to hydrogens in the a position and we find that 

O 

II 
protons of the type CH 3 — C— R come at lower fields (0.3 ppm) than those of 

R 

I 

CH,-C=CR, . 



11-3 preparation of aldehydes and ketones 

Of the six methods listed below for the preparation of aldehydes and 
ketones, the first four have already been met. The other two will, therefore, be 
described in more detail. There are a number of additional routes to aromatic 
aldehydes and ketones, in which the carbonyl group is directly attached to a 
benzene ring, and these will be encountered in Chapter 24. 

1. Oxidation of alcohols (Section 10-6). The primary alcohols give aide- 



sec 11.3 preparation of aldehydes and ketones 279 

primary alcohols RCH 2 OH > RCHO 

secondary alcohols R 2 CHOH ► R 2 C = 

hydes readily with many oxidants but subsequent oxidation of the aldehyde 
to a carboxylic acid, RC0 2 H (Section 11-4G), can occur; secondary alcohols 
give ketones in good yield; tertiary alcohols are inert unless drastic conditions 
are used, in which case carbon-carbon bond rupture occurs to give a mixture 
of cleavage products. 

2. Oxidative cleavage of alkenes (Section 4-4G). The number of alkyl 

R 2 C=CHR' ° 3 > ^" > R 2 C = + R— CHO 

groups at the double bond determines whether aldehydes or ketones are 
produced ; many oxidants will cleave a carbon-carbon double bond but ozone is 
one of the few that will not further oxidize the aldehyde produced; oxidative 
cleavage of 1,2-glycols (Section 10-7) by HI0 4 or Pb(OAc) 4 produces the 
same products (aldehydes or ketones). 

3. Hydration of alkynes (Section 5-4). This method can be used only to 

o 

RC^CR + H 2 ,"1» RCCH 2 R 

prepare ketones (and acetaldehyde, CH 3 CHO). 

4. Organocadmium compounds react with acyl chlorides (Section 9-9C). 

O O 

II II 

R— C— CI + R'CdCl ► R— C— R' + CdCl 2 

Grignard reagents react the same way but, being more reactive than organo- 
cadmium compounds, they usually attack the ketone to give tertiary alcohols. 

5. Reduction of carboxylic acids to aldehydes. Conversion of a carboxylic 
acid to an aldehyde by direct reduction is not easy to achieve because acids 
are generally difficult to reduce, whereas aldehydes are easily reduced. Thus 
the problem is to keep the reaction from going too far. 

The most useful procedures involve conversion of the acid to a derivative 
that either is more easily reduced than an aldehyde, or else is reduced to a 
substance from which the aldehyde can be generated. The so-called Rosen- 
mund reduction involves the first of these schemes; in this procedure, the acid 
is converted to an acyl chloride, which is reduced with hydrogen over a 
palladium catalyst to the aldehyde in yields up to 90 %. The rate of reduction 
of the aldehyde to the corresponding alcohol is kept at a low level by poison- 
ing the catalyst with sulfur. 

R-C0 2 H ( S q° C ^ c|) > RCOC1 p " 2 s) ' RCHO + HC1 

Reduction of an acid to a substance that can be converted to an aldehyde is 
usefully achieved by way of lithium aluminum hydride reduction of the 
nitrile corresponding to the acid. The following scheme outlines the sequence 



chap 1 1 aldehydes and ketones I 280 

of reactions involved starting with the acid : 
RCO,H SOC ' 2 > RCOC1 NHa > RCONH, 



POCI = > RC.N -t^ R _ C =N-Li 



(-H z O) 

H H 

_2«^ R _(L N H "1"^ R_<L 



R 




R 


1 1 

-c-c- 

1 | 


H® , 


1 

— c-c 

ii 1 


HO OH 




II 1 

o 



The reduction step is usually successful only {{inverse addition is used; that is, 
a solution of lithium aluminum hydride is added to a solution of the nitrile in 
ether, preferably at low temperatures. 

[Vc-N t!A!Hi_^ ^J^ [^cho 

l""^^ (C 2 H 5 ) 2 0, - 50° ^ 

70% 
cyclopropanecarbonitrile cyclopropanecarboxaldehyde 

If the nitrile is added to the hydride, the reduction product is a primary 
amine, RCH 2 NH 2 . 

6. Rearrangements of 1,2-glycols. Many carbonyl compounds can be 
usefully synthesized by acid-catalyzed rearrangements of 1,2-glycols, the 
so-called "pinacol-pinacolone" rearrangement: 



+ H,0 



R = alkyl, aryl, or hydrogen 

The general characteristics of the reaction are similar to those of carbonium 
ion rearrangements (see Section 8-13). The acid protonates one of the —OH 
groups and makes it a better leaving group. The carbonium ion which results 
can then undergo rearrangement by shift of the neighboring group R with its 
pair of bonding electrons to give a new, more stable, species with a carbon- 
oxygen double bond. 

An important example is provided by the rearrangement of pinacol to 
pinacolone, as follows : 

H 3 C CH 3 H 3 C CH 3 H 3 C CH 3 

II H® II -H,0 "->l 

CH 3 -C-C-CH 3 . CH 3 -C-C-CH 3 > CH 3 -C-C-CH 3 

ii ii Vi • 

HO OH HO OH, k :OH 

® * 
2, 3-dimethyl-2, 3-butanediol 
(pinacol) 



CH 3 CI 

I -H- I 



O CH 3 HO CH 3 

3, 3-dimethyl-2-butanone 
(pinacolone) 



sec 11.4 reactions of aldehydes and ketones 281 

Alkenes may be converted to carbonyl compounds with the same number of 
carbon atoms by hydroxylation, followed by rearrangement. Isobutyralde- 
hyde is made on a large scale this way from 2-methylpropene. 

CH 3 CH 3 CH 3 

CH 3 -C=CH 2 — ^ CH 3 -C-CH 2 -^^-* CH 3 -C-CH 2 

CI CI HO OH 



CH 3 

CH C-CHO 

I 
H 



11-4 reactions of aldehydes and ketones 

The important reactions of carbonyl groups characteristically involve 
addition at one step or another. We have already discussed additions achieved 
by Grignard reagents as part of Chapter 9, It will be recalled that steric 
hindrance plays an important role in determining the ratio between addition 
and other competing reactions. Similar effects are noted in a wide variety cf 
other reactions. We shall expect the reactivity of carbonyl groups in addition 
processes to be influenced by the size of the substituents thereon because 
when addition occurs, the substituent groups are pushed back closer to one 
another. On this basis, we anticipate decreasing reactivity with increasing 
bujkiness of substituents, as in the accompanying series. 



H. R CH 



H 3 C ,CH 3 

V 



) C =0 > >=0 > )c=0 » gx' \ = 
H X CH 3 CH 3 *%/ 

H 3 C X X CH 3 

The term " reactivity " is often used somewhat loosely in connection with 
discussions of this type. Although it is probably best confined to considera- 
tions of reaction rate it is also rather widely employed in connection with 
equilibrium constants — -that is, how far the reaction goes, as well as how fast 
it goes. If steric hindrance is high we expect equilibrium reactions to go 
neither very fast nor very far. In the case of carbonyl additions, this is gen- 
erally true, but there are exceptions which will be noted later. 

Cyclic ketones almost always react more rapidly in addition processes than 
their open-chain analogs. 

O O 

H 2 c' N CH 2 > H 2 C' X CH 2 
H 2 C-CH 2 H 3 C CH 3 



chap 11 aldehydes and ketones I 282 

This is because the alkyl groups of the open-chain compounds have consider- 
ably more freedom of motion and produce greater steric hindrance in transi- 
tion states for addition. 

Electrical effects are also important in influencing the ease of addition to 
carbonyl groups. Electron-attracting groups facilitate the addition of nucleo- 
philic reagents to carbon by increasing its positive character. Thus we find 
that compounds such as the following add nucleophilic reagents readily : 

O O 

II II 

CC1 3 — C— H H0 2 C — C— C0 2 H 

trichloroacetaldehyde (chloral) oxomalonic acid 

Although Grignard reagents, organolithium compounds, and the like gen- 
erally add to aldehydes and ketones (Section 9-9C) rapidly and irrevers- 
ibly, this is not true of many other reagents. Their addition reactions may 
require acidic or basic catalysts and have relatively unfavorable equilibrium 
constants. Some of these reactions are discussed in considerable detail in the 
following sections. 



A. CYANOHYDRIN FORMATION 

Hydrogen cyanide adds to many aldehydes and ketones to give a-cyano- 
alcohols, usually called cyanohydrins. 

o OH 

II 
CH3-C-CH3 + HCN 

The products are useful in synthesis — for example, in the preparation of 
cyanoalkenes and hydroxy acids : 



( - H 2 0) 



CH 3 

I 
* CH 2 =C — C=N 

OH / a-methacrylonitrile 

I 



CH3-C-CH3 



OH 



acetone cyanohydrin _ >jh / ^^ 3 — — ^^ 3 

C0 2 H 

2-hydroxy-2-methylpropanoic acid 
(dimethylglycolic acid) 

An important feature of cyanohydrin formation is that it requires a basic 
catalyst. In the absence of base, the reaction does not proceed, or is at best 
very slow. In principle the basic catalyst might activate either the carbonyl 
group or the hydrogen cyanide. With hydroxide ion as the base, one reaction 
to be expected is a reversible addition of hydroxide to the carbonyl group. 



sec 11.4 reactions of aldehydes and ketones 283 



5® 8e e H3C X> 

(CH 3 ) 2 C=^p +OH ;=± V 

H 3 C OH 
HI 
However, such addition is not likely to facilitate formation of cyanohydrin, 
because it represents a competitive saturation of the carbonyl double bond. 
Indeed, if the equilibrium constant for this addition were large, an excess of 
hydroxide ion could inhibit cyanohydrin formation by tying up acetone as the 
adduct [1]. 

Hydrogen cyanide itself has no unshared electron pair on carbon and is 
unable to form a carbon-carbon bond to a carbonyl carbon (indeed, the fact is 
that hydrogen cyanide, when it does react as a nucleophilic agent toward 
carbon, forms C— N rather than C— C bonds). However, an activating 
function of hydroxide ion is clearly possible through conversion of hydrogen 
cyanide to cyanide ion, which can function as a nucleophile toward carbon. 
A complete reaction sequence for cyanohydrin formation is shown. 

H— C=N + OH . :C=N + H 2 

CH 3 H 3 C O e 

C=0 + :C = N , C 

/ IT /-•' \ 



CH, 



C = N 



H 3 C O e H 3 C OH 

C + H 2 ^z^ r' + OH 

H 3 C C^N H 3 C C^N 

The last step regenerates the base catalyst. All steps of the overall reaction 
are reversible but, with aldehydes and most nonhindered ketones, formation 
of the cyanohydrin is reasonably favorable. In practical syntheses of cyano- 
hydrins, it is convenient to add a strong acid to a mixture of sodium cyanide 
and the carbonyl compound, so that hydrogen cyanide is generated in situ. 
The amount of acid added should be insufficient to consume all the cyanide 
ion, so that sufficiently alkaline conditions are maintained for rapid addition. 
If sodium cyanide alone is used the reaction is rapid but does not go to com- 
pletion because of the reversibility of the final step. 

Table 1 1 .2 shows the extent of reaction for some simple carbonyl com- 
pounds. The effect of introducing an isopropyl group in the 2 position of 
cyclohexanone is seen to be considerable and is probably steric in origin. 
However, in other cases where steric effects might be expected to be impor- 
tant, no evidence for such effects is reported. Thus virtually the same equilib- 
rium constant is found for acetone and methyl ?-butyl ketone. This is difficult 
to explain and it appears that the factors governing cyanohydrin formation 
require further study. 



B. HEMIACETAL AND ACETAL FORMATION 

Hemiacetals and acetals are products of addition of alcohols to aldehydes 
— thus, for acetaldehyde and methanol, 



chap 1 1 aldehydes and ketones I 284 

Table 11-2 Equilibrium in cyanohydrin formation in 96% ethanol solution 
at 20° 



carbonyl compound 


equilibrium constant, 
liters/mole 


% cyanohydrin at 
equilibrium" 


CH 3 COCH 3 


32.8 


84 


CH 3 COCH 2 CH 3 


37.7 


85 


CH 3 COCH(CH 3 ) 2 


81.2 


90 


CH 3 COC(CH 3 ) 3 


32.3 


84 


H 2 C C ^ 

I c=o 


67 


89 


H * C ^CH 2 






H^C CH-» 

/ \ 2 

H,C C = 
\ / 
H 2 C-CH 2 






11,000 


~100 






H 2/ C-CH 2 
H 3 C— HC C=0 

H 2 C-CH CH 3 

c 
h' x ch 3 






15.3 


78 


(menthone) 







" Starting with 1 M concentrations of carbonyl compound and hydrogen cyanide. 



CH, H,C OH „ H,C OCH 3 

\ 3 CH3OH 3 \ / CH jOH, - H 2 Q 3 \ / 3 

c=o ^=t C - A 

H 7 H X OCH3 » OCH3 

a hemiacetal acetaldehyde dimethyl acetal 

(1-methoxyethanol) (1, 1-dimethoxyethane) 

First, we shall consider hemiacetal formation, which is catalyzed by both 
acids and bases. The base catalysis is similar to that involved in cyanohydrin 
formation. The slow step here is the addition of alkoxide ion to the carbonyl 
group. 

=± CH 3 O e + H 2 

CH 3 H 3 C X)CH 3 

\ x — -^e slow 3 \ / 

C = 0-CH 3 . C. 

H ' ^> H y V 

H3C X/ OCH3 fast H 3C x/ OCH 3 



t + H 2 . C + OH e 

H O e W OH 



sec 11.4 reactions of aldehydes and ketones 285 

Acid catalysis of hemiacetal formation might involve activation of either 
the alcohol or the carbonyl compound. However, the only simple reaction 
one would expect between various species of alcohols and proton donors is 
oxonium ion formation, which hardly seems the proper way to activate an 
alcohol for nucleophilic attack at the carbonyl group of an aldehyde. 



CH,OH + H a 



H 

I 

CH,-0- H 



methyloxonium ion 

On the other hand, formation of the oxonium ion (or conjugate acid) 1 of 
the carbonyl compound is expected to provide activation for hemiacetal 
formation by increasing the positive character of the carbonyl carbon. 



CH 3 

\ 

c=o 

/ 

H 



CH^-^ 

C=?=OH +O-CH3 
H H 



CH 3 

\ © 
C = OH 

H 



CH 3 



C— OH 



/ 
H 



conjugate acid of acetaldehyde 
,CH 3 



H,C. 



O 



/ 



W y \>H 



H,C N /O-CH, 
H /C \)H 



+ H 9 



hemiacetal 
In water solution most aldehydes form hydrates. This reaction is analogous 



\ 
C=0 + H,0 



R OH 

X 

H OH 



to hemiacetal formation and is catalyzed by both acids and bases. The equilib- 
rium constant for hydrate formation depends on steric and electrical factors. 
In contrast to hemiacetal formation, acetal formation is catalyzed only by 
acids. Addition of a proton to a hemiacetal can occur two ways to give either 
[2] or [3]. 



CH, 



h 3 c ;o 

X H 

H OH 



H 3 C x ^OCH, 
H OH 



+ H tt 



[2] 



H 3 C x OCH 3 

C 

/ \© 
H O-H 
I 
H 



[3] 



1 The conjugate acid of X is XH®, while the conjugate base of HY is Y e . 



chap 11 aldehydes and ketones I 286 

The first of these [2] can lose CH 3 OH and yield the conjugate acid of acetalde- 
hyde. This is the reverse of acid-catalyzed hemiacetal formation. 



H 3 C ffi q ch, 

V H . C=OH + CH3OH 

/ > / 

H M)H h 

[2] 

The second of these [3] can lose H 2 and give a new entity, the methyloxo- 
nium cation of the aldehyde [4]. 

H 3 C /O-CH3 CHj 

C , C=0-CH 3 + H 2 

/ ,\s> / 

H VO-H h 

I 
H [4] 

[3] 

The reaction of [4] with water gives back [3], but reaction with alcohol leads 
to the conjugate acid of the acetal [5]. Loss of a proton from [5] gives the 
acetal. 

CH 3 
H,C. ffl .O. ,,„ H 3 C, ,OCH, 



\ ^© ~^\ "3-^ / x _ H S 3- x / 

C=0-CH 3 + 0-CH 3 ■ C H , C 

H h H ° _CH 3 H OCH 3 

[4] [5] 

The fact that acetals are formed only in an acid-catalyzed reaction has the 
corollary that acetal groups are stable to base. This can be synthetically very 
useful, as illustrated by the following synthesis of glyceraldehyde from 
readily available acrolein (CH 2 =CHCHO). Hydrogen chloride in ethanol 
adds in the anti-Markownikoff manner to acrolein to give /?-chloropropion- 
aldehyde, which then reacts with the ethanol to give the acetal. (Markow- 
nikoff 's rule does not apply when a carbon-carbon double bond is conjugated 
to groups such as carbonyl.) 



CH 2 =CHCHO + HCI - 

ClCH 2 CH 2 CHO + 2C 2 H 5 OH ► ClCH 2 CH 2 CH(OC 2 H 5 ) 2 

The key step in the synthesis involves E2 dehydrochlorination of the chloro- 
acetal without destroying the acetal group, which is stable to base. 

ClCH 2 -CH 2 -CH(OC 2 H 5 ) 2 + KOH E2 » CH 2 =CH-CH(OC 2 H 5 ) 2 + KC1 + H 2 

Hydroxylation of the double bond with neutral permanganate then gives the 
diethyl acetal of glyceraldehyde. This kind of step would not be possible with 
acrolein itself because the aldehyde group reacts with permanganate as easily 



sec 11.4 reactions of aldehydes and ketones 287 

Table 1 1*3 Conversions of aldehydes to acetals with various alcohols 
(1 mole of aldehyde to S moles of alcohol) 



°/ conversion to acetal 
aldehyde 



isopropyl J-butyl 

ethanol cyclohexanol alcohol alcohol 



CH3CH0 


78 


56 


43 


(CH 3 ) 2 CHCHO 


71 




23 


(CH 3 ) 3 CCHO 


56 


16 


11 


QH5CHO 


39 


23 


13 



23 



as does the double bond. 

CH 2 =CH-CH(OC 2 H 5 ) 2 ^^ CH 2 -CH-CH(OC 2 H 5 ) 2 

OH OH 

Finally, mild acidic hydrolysis of the acetal function yields glyceraldehyde. 

CH 2 -CH-CH(OC 2 H 5 ) 2 + H 2 "" > CH 2 -CH-CHO + 2 C 2 H,OH 
II II 

OH OH OH OH 

glyceraldehyde 

The position of equilibrium in acetal and hemiacetal formation is rather 
sensitive to steric hindrance. Large groups in either the aldehyde or the alcohol 
tend to make the reaction less favorable. Table 11-3 shows some typical 
conversions in acetal formation when 1 mole of aldehyde is allowed to come 
to equilibrium with 5 moles of alcohol. 

Hemiacetal formation occurs readily in an intramolecular manner when a 
five- or six-membered ring can be formed from a hydroxyaldehyde. This type 

H 2 C-CH 2/ OH 

HOCH 2 CH 2 CH,CHO < > 2 | C 

H 2 CV \ H 

of reaction is especially important for carbohydrates as will be discussed in 
Chapter 15. 



C. POLYMERIZATION OF ALDEHYDES 

A reaction closely related to acetal formation is the polymerization of alde- 
hydes. Both linear and cyclic polymers are obtained. For example, formalde- 
hyde in water solution polymerizes to a solid long-chain polymer called para- 
formaldehyde or "polyoxymethylene." This material, when strongly heated, 

n-CH 2 =0 + H 2 ► H— 0-CH 2 ~(-O— CH 2 ^-0-CH 2 — O— H 

paraformaldehyde 



chap 1 1 aldehydes and ketones I 288 

reverts to formaldehyde; it is therefore a convenient source of gaseous 
formaldehyde. When heated with dilute acid, paraformaldehyde yields the 
solid trimer trioxymethylene (mp 61°). The cyclic tetramer is also known. 

CH 2 
of ^O 

I I trioxymethylene 

H 2 C. /CH 2 

Long-chain formaldehyde polymers have become very important as plas- 
tics in recent years. The low cost of paraformaldehyde (10-15 cents/lb) is 
highly favorable in this connection, but the instability of the material to 
elevated temperatures and dilute acids precludes its use in plastics. However, 
the "end-capping" of polyoxymethylene chains through formation of esters 
or acetals produces a remarkable increase in stability, and such modified 
polymers have excellent properties as plastics. Delrin (DuPont) and Celcon 
(Celanese) are stabilized formaldehyde polymers with exceptional strength 
and ease of molding. 



D. CONDENSATIONS OF CARBONYL COMPOUNDS WITH 
DERIVATIVES OF AMMONIA 

Ammonia adds readily to acetaldehyde to give a crystalline compound. 

OH 

I 
H3C-C-NH, 

I 
H 

The corresponding adducts with most other aldehydes are not very stable, 
undergoing dehydration and polymerization rapidly. Amines and other 
derivatives of ammonia react in a similar way except that the dehydration 
products are usually stable compounds that do not polymerize. Reactions of 
this sort, in which two molecules combine and then water is split out, are 
often called condensation reactions and usually require acid or base catalysis. 
In the case of condensations between carbonyl compounds and amines, catal- 
ysis is normally brought about by acids. 

\ u® \ 

C=0 + H 2 N-R ► C 

/ / 

Table 11-4 summarizes a number of important reactions of this type and the 
nomenclature of the reactants and products. 



E. HYDROGEN HALIDE ADDITION AND REPLACEMENT BY 
HALOGEN 

Addition of hydrogen halides to carbonyl groups is so easily reversible as to 
prevent isolation of the products. 



sec 1 1 .4 reactions of aldehydes and ketones 289 

Table 1 1*4 Condensation reactions of carbonyl compounds with 
derivatives of ammonia 



reactant 




typical product 


class of product 


H 2 N-R (R = alkyl, aryl 


CH 3 CH=N-CH 3 


inline" 


or hydrogen) 




acetaldehyde methylimine 


(Schiff's base) 


amine 




CH 3 
C = N-NH, 




NH 2 -NH 2 




hydrazone 


hydrazine 




/ 
CH 3 

acetone hydrazone 








CH 3 CH 3 

\ / 3 
C=N-N=C 
/ \ 
CH 3 CH 3 


azine 






acetone azine 








N0 2 




H 2 N-NHR (R = 


= alkyl, aryl, 


or <X>=N — NH-f \-N0 2 


substituted 


hydrogen) 




hydrazone* 


substituted hydrazine 


cyclobutanone 2,4-dinitro- 








phenylhydrazone 




O 




H 




II 
H 2 NNHCNH 2 




{ VCH==N— N-C-NH, 


semicarbazone 6 


semicarbazide 




benzaldehyde semicarbazone 




HO-NH 2 




CH 2 =N-OH 


oxime 6 


hydroxylamine 




formaldoxime 





" Most unsubstituted imines, that is, R = H, are unstable and polymerize. 
* Usually these derivatives are solids and are excellent for the isolation and characterization of 
aldehydes and ketones. 



CH 3 

\ 

C=0 + HC1 - 

/ 
H 



CH 3 CI 

X 

H OH 



However, many aldehydes react with alcohols in the presence of an excess of 
hydrogen chloride to give a-chloro ethers : 



CH 3 
\ 
C=0 + HC1 + CH 3 OH 

/ 3 

H 



CH 3 O — CH 3 
\ / 
/ C N + H 2 

H CI 



Replacement of the carbonyl function by two chlorines occurs with phos- 



chap 1 1 aldehydes and ketones I 290 

phorus pentachloride in ether, and by two fluorines with sulfur tetrafluoride : 
CI 




F. REDUCTION OF CARBONYL COMPOUNDS 

Formation of Alcohols. The easiest large-scale reduction method for con- 
version of aldehydes and ketones to alcohols is by catalytic hydrogenation. 

H >r c >=o ȣ&& h ^ c >hoh 

h > c Vh 2 <N0 ' 50 h > c Vh 2 

95-100% 

The advantage over most chemical reduction schemes is that usually the 
product can be obtained simply by filtration from the catalyst followed by 
distillation. The usual catalysts are nickel, palladium, copper chromite, or 
platinum promoted with ferrous iron. Hydrogenation of aldehyde and ketone 
carbonyl groups is much slower than of carbon-carbon double bonds and 
rather more strenuous conditions are required. This is not surprising, because 
hydrogenation of carbonyl groups is calculated to be less exothermic than 
that of double bonds. It follows that it is generally not possible to reduce a 

\ / II 

C=C + H 2 ' > -C— C- A#=-30kcal 

/ \ II 

H H 

\ I 

C=0 + H, ► — C— OH Atf=-12kcal 

/ I 

H 

carbonyl group with hydrogen in the presence of a double bond without also 
saturating the double bond. 

In recent years inorganic hydrides such as lithium aluminum hydride, 
LiAlH 4 , and sodium borohydride, NaBH 4 , have become extremely important 
as reducing agents of carbonyl compounds. These reagents have considerable 
utility, especially with sensitive and expensive carbonyl compounds. The 
reduction of cyclobutanone to cyclobutanol is a good example : 



^o^^O" 



OH 

90% 

With the metal hydrides the key step is transfer of a hydride ion to the carbonyl 
carbon of the substance being reduced : 



sec 11.4 reactions of aldehydes and ketones 291 

R H r 0-AlH 3 Li 

\ r< |e © \ / 

C=0 + H— Al-H Li ► .C 

Lithium aluminum hydride is handled very much like a Grignard reagent, 
since it is soluble in ether and is sensitive to both oxygen and moisture. 
(Lithium hydride is insoluble in organic solvents and is not an effective 
reducing agent for organic compounds.) All four hydrogens on aluminum can 
be utilized. 

H OCH 3 

Li® H-A1^H + 4CH 2 = ► Li e CH 3 0-A1-0CH 3 -**sitiU 

H OCH 3 

4CH 3 OH + Al 3 ® + Li® 

The reaction products must be decomposed with water and acid as with the 
Grignard complexes. Any excess lithium aluminum hydride is decomposed by 
water and acid with evolution of hydrogen. 

LiAlH 4 + 4H® ► Li® + AF© + 4 H 2 

Lithium aluminum hydride usually reduces carbonyl groups without affec- 
ting carbon-carbon double bonds. It is, in addition, a strong reducing agent 
for carbonyl groups of carboxylic acids, esters, and other acid derivatives, 
as will be described in Chapter 13. 

Sodium borohydride is a milder reducing agent than lithium aluminum 
hydride and will reduce aldehydes and ketones but not acids or esters. It 
reacts sufficiently slowly with water in neutral or alkaline solution so that 
reductions which are reasonably rapid can be carried out in water solution 
with only slight hydrolysis of the reagent. 

NaBH 4 + 4CH 2 =0 + 4H 2 ► 4CH 3 OH + NaB(OH) 4 

Reduction of Carbonyl Compounds to Hydrocarbons. There are a number of 
methods of transforming 

\ \ 

C = to CH 2 

In some cases, the following sequence of conventional reactions may be 
useful : 

, .0 *f OH H V H 

\y \y \ r s \y 

<f ^L <f -^ ? -£- i 
c c c ' c 

1 H ' H ' H 



chap 1 1 aldehydes and ketones I 292 

This route is long, requires a hydrogen a to the carbonyl function, and may 
give rearrangement in the dehydration step (Section 10-5B). 

More direct methods may be used depending on the character of the R 
groups of the carbonyl compound. If the R groups are stable to a variety of 
reagents there is no problem, but with sensitive R groups not all methods are 
equally applicable. When the R groups are stable to acid but unstable to base, 
the Clemmensen reduction with amalgamated zinc and hydrochloric acid is 
often very good. The mechanism of the Clemmensen reduction is not well 

o 

"^ " ^ ^VcH 2 CH 2 C0 2 C 2 H 5 



^ y— <~— <~n 2 <~<-» 2 ^ 2 n 5 HC | 

59% 

understood. It is clear that in most cases the alcohol is not an intermediate, 
because the Clemmensen conditions do not suffice to reduce most alcohols to 
hydrocarbons. 

When the R groups of the carbonyl compound are stable to base but not 
to acid, the Wolff-Kishner reduction is often 1 very convenient. This involves 
treating the carbonyl compound with hydrazine and potassium hydroxide 
in dimethyl sulfoxide solution. 

X CH 2 CH 2 

H 2 <\ C-O + NH 2 -NH 2 -&¥* H 2 C^ Vh 2 + N 2 + H 2 
CH 2 \rf 2 

90% 



G. OXIDATION OF CARBONYL COMPOUNDS 

Aldehydes are easily oxidized by moist silver oxide or by potassium per- 
manganate solution to the corresponding acids. 

o o 

« [O] // 

R-C l J > R-C 
\ \ 

H OH 

Ketones are much more difficult to oxidize at the carbonyl group than 
aldehydes. Ketones with a hydrogens can be oxidized in acidic or basic solu- 
tions because of enol formation, as will be described in Chapter 12. Thus, 

O OH 

II OH e (or H®) I 

CH 3 -C— CH 3 , CH 3 -C=CH 2 

ketone enol 

Enols, being unsaturated, are easily attacked by reagents which oxidize 
unsaturated molecules. 

Methyl ketones can be oxidized to carboxylic acids with the loss of a 
carbon atom by the haloform reaction (Section 12- 1C). 



sec 11.4 reactions of aldehydes and ketones 293 
H. THE CANNIZZARO REACTION 

A characteristic reaction of aldehydes without a hydrogens is the self 
oxidation-reduction that they undergo in the presence of strong base. Using 
formaldehyde as an example, 

O 

heat " ff 

2CH 2 =0 + NaOH ™ Q > CH 3 OH + H—C 

ONa 

If the aldehyde has a hydrogens, other reactions usually occur more rapidly. 
The mechanism of the Cannizzaro reaction combines many features of 
other processes studied in this chapter. The first step is reversible addition of 
hydroxide ion to the carbonyl groups. 

H H O e 

\ e \ / 

C = + OH . C^ 

H 7 H OH 

The hydroxyalkoxide ion so formed can act as a hydride ion donor like lithium 
aluminum hydride and reduce a molecule of formaldehyde to methanol. 



hVs . 



H 2 C=0 + X C^ -^ H3C-O + H-C — ^U CH3OH + HC0 2 
H OH OH 



I. DISTINGUISHING BETWEEN ALDEHYDES AND KETONES 

Most of the reactions described in earlier sections of this chapter take place 
with both aldehydes and ketones. There are a number of tests, however, to 
distinguish between these classes of compounds, one of which is the use of 
ammoniacal silver ion. This is a mild reagent which oxidizes aldehydes but 

RCHO _ Ag(NH ^ g ', RCO.H + AgCs) 

leaves most ketones untouched. If the reagent and substrate are carefully 
mixed in a test tube the metallic silver will deposit on the walls to form a 
mirror. A similar test for aldehydes involves the oxidizing action of com- 
plexed cupric ion (Fehling's reagent); a red precipitate of cuprous oxide 
indicates the oxidation of an aldehyde or other easily oxidized substance. 

Another reaction that is characteristic of aldehydes and a few ketones is 
addition of sodium bisulfite to the carbonyl group. The resulting ionic 

O OH 

R-C + Na®HSOf . R-C-S0 3 e Na ffl 

\ I 

H H 



chap 1 1 aldehydes and ketones I 294 

compound is a salt of a sulfonic acid (Section 19-2D), although it is usually 
called simply the bisulfite addition compound of the particular aldehyde. It 
contains a carbon-sulfur bond. The reaction is reversible and addition of acid 
regenerates the aldehyde by converting the sodium bisulfite to sulfur dioxide. 
Most bisulfite addition compounds are nicely crystalline and aldehydes are 
sometimes purified through them. The few ketones that form bisulfite addition 
compounds have relatively unhindered carbonyl groups — for example, 
acetone and cyclopentanone. 

Although the carbonyl stretching frequencies of aldehydes and ketones 
lie close together (1705-1740 cm -1 ) the aldehydic C— H stretching frequency 
occurs at somewhat lower frequency (near 2700 cm -1 ) than the C— H 
absorption in ketones and most other compounds. This band is not always 
easy to recognize, however, and a more characteristic absorption is the nmr 
peak of the aldehydic proton which occurs at very low field (9.7 ppm from 
TMS). 

summary 

// II 

Aldehydes, R— C , and ketones, R— C— R, both contain the carbonyl 

H 
group. The IUPAC system names aldehydes as alkanals — methanal, for 
example — and ketones as alkanones — propanone, for example. In addition, 
the simple aldehydes can be named as analogs of carboxylic acids— formalde- 
hyde, for example — and ketones can be named by giving the names of the two 
groups attached to carbonyl — dimethyl ketone, for example. 

The polarity of the carbonyl group accounts for aldehydes and ketones being 
part way between alkanes and alcohols in terms of most physical properties. 

Aldehydes and ketones have characteristic infrared, ultraviolet, and nmr 
spectra. Strong absorption just above 1700 cm -1 in the infrared is due to 
C=0 stretching. Rather weak absorption near 2750 A in the ultraviolet is 
due to an n -> tc* electronic transition in saturated aldehydes or ketones ; 
conjugation of the carbonyl group with a double bond causes an increase in 
both A max and e max and the appearance of powerful n -> n* absorption just 
above 2000 A. The aldehydic proton absorbs at very low field in the nmr. 

The methods of preparing aldehydes and ketones may be summarized as 
follows: 
RCH,OH ^ z- RCH=CR 2 



RCHO q R 2 C=0 «- 



RC0 2 H 



RC-C1 

O OH OH ° 

II II II 

RC=CR > RCCH 2 R R 2 C-CR 2 ► RCCR 3 

Most of the reactions of aldehydes and ketones at the carbonyl group 
involve addition processes. These usually occur more readily with aldehydes 
or with cyclic ketones than with open-chain ketones, especially those with 



summary 



295 



bulky groups attached. In almost all the reactions shown below a nucleophile 
attacks the carbon atom of the carbonyl group. In some cases the active 
nucleophile is produced by the action of base; in others, the carbonyl group 
is activated by protonation. 



p 



R-C 



O 

II 

-C-+H® 



RCHOHR' 



-► RCHOHCN 

OH 
I 
-» RCOR 

I 
H 



OR 
I 
RC-OR 

I 
H 

-> polymers 



-» RCH=NZ 



-> RCHXOR 



-► RCHX, 



->■ RCH 2 OH 



-> RCH 3 



-* RCH 2 OH + RC0 2 H 



®OH 

II 
— C- 

R 2 C=0 



OH 

I 



-> -C- 

I 

N 



-> R 2 R'COH (by the Grignard 
reaction) 



► R 2 COHCN 


(cyanohydrms) 


OH 




* 1 
► RC— OR 

I 




R 








(hemiacetals, ace- 
tals, hemiketals, 






ketals) 


OR 

I 




RC-OR 

1 




R 




* 


C=NZ 


(Z = -H, -NH 2 , 






O 




Jl 
-NHCNH 2 ,-OH) 




(with HX and ROH) 


► R 2 CX 2 


(with PX 5 ) 



-♦ R 2 CHOH (with LiAlH 4 or 
cat. H 2 ) 



-* RCH 2 R 



(Clemmensen or 
WoJff-Kishner 
reductions) 

(Cannizzaro 
reaction) 



Acid catalyzed through prior protonation of the carbonyl group. 



Aldehydes can be distinguished from most ketones by their ability to 
reduce Ag 1 or Cu 11 , by their ability to form bisulfite addition compounds, and 
by the characteristic low-field nmr absorption of the aldehydic proton. 



chap 1 1 aldehydes and ketones I 296 



exercises 



11-1 Draw structural formulas for each of the following substances. 



a. hexanal 

b. divinyl ketone 

c. phenylethanal 



d. 3-phenyl-2-butenal 

e. cyclohexyl phenyl ketone 



11-2 Name the following substances according to the systems developed in 
Table 11-1. 



a. CH 3 CH=CHCHO 



<x 



O 



d. CF3COCF3 

e. CH 3 CH— CHCHO 

I I 
OH CH 3 

/. (CH 3 ) 2 C(CH 2 ) 2 CHCHO 



c. C fi H,COCH,COC„H, 



11-3 Write structural formulas for each of the following : 

a. bromomethyl 1 ,2-dimethylcyclopropyl ketone 

b. diallyl ketone 

c. 4-bromo-2-methyl-3-butynal 

d. diacetylacetylene (and supply the IUPAC name; 

O 
II 
acetyl =CH 3 -C-) 

e. 3-acetyl-2-cyclohexenone 

11-4 Show how you could prepare 3-hexanone using, in turn, each of the following 
compounds as starting material. 



a. butanal 

b. 3-hexyne 



c. 3-ethyl-2-hexene 

O 

// 

d. propanoyl chloride (CH 3 CH 2 C ) 

CI 



11-5 A compound C 6 Hi 2 was treated with ozone and the reaction products 
boiled with water and zinc dust. 

A single organic product was obtained which had a strong infrared absorp- 
tion band near 1725 cm -1 and a low-field absorption in the nmr (almost 10 
ppm from TMS). What is the structure and name of the compound C 6 Hi 2 ? 

11-6 Identify the reagents A-F in each of the steps below: 



2-pentanol 



2-pentanone 



D C 

butanal < < butanoic acid 



exercises 297 

11-7 Predict the products to be expected from acid-catalyzed rearrangements of 
1 ,2-propanediol and 2-methyl-2,3-butanediol. 

11-8 Treatment of tetramethylethylene oxide (CH 3 ) 2 C — C(CH 3 ) 2 with acid 
produces pinacolone (pp. 280-281). Explain. O 

11-9 How might one dehydrate pinacol to 2,3-dimethyl-l,3-butadiene without 
forming excessive amounts of pinacolone in the process ? 

11-10 Arrange the following pairs of compounds in order of expected reactivity 
toward addition of a common nucleophilic agent such as hydroxide ion to 
the carbonyl bond. Indicate your reasoning. 



O O 

II II 

a. CH3-C-CH3 and CH 3 -C-CC1 3 

O O 

II II 

b. (CH 3 ) 3 C-C-H and CH 3 -C— CH 3 

o o 

II II 

c. CH 3 — C-OCH 3 and CH 3 — C— CH 3 

o o 

II II 

d. CH 3 -C-C1 and CH 3 -C-CH 3 

, ^CH 2 CH 2 

e. I C=0 and H 2 C C=0 
u r / \ / 



11-11 What should be the rate law for the formation of acetone cyanohydrin by the 
mechanism given on p. 283 if the first step is slow and the others fast ? If the 
second step is slow and the others fast? 

a. Calculate A.H for the formation of acetone cyanohydrin from hydro- 
gen cyanide and acetone in the vapor phase at 25°. Do the same for 
the formation of dimethylethynylcarbinol from ethyne and acetone. 

b. What are the prospects for carrying out addition of ethyne to acetone 
by the same procedure used for hydrogen cyanide? Explain. 

c. What would you expect to happen if one attempted to prepare 
acetone cyanohydrin with acetone and a solution of sodium cyanide ? 

11-12 The equilibrium constant for hydration is especially large for formaldehyde, 
trichloroacetaldehyde, and cyclopropanone. Explain. 

11-13 Write equations for the synthesis of the following substances based on the 
indicated starting materials. Give the reaction conditions as accurately as 
possible. 

a. (CH 3 ) 2 CHCHO from (CH 3 ) 2 CHCH 2 CH 2 OH 

O 

II 

CH 2 -CH-C-CH 3 CH— CHC0 2 H 

b. I I from I I 
CH 2 — CH 2 



chap 11 aldehydes and ketones I 298 



CH 2 -CHO C H 2 

c. CH, from | C=0 

\ H 2 CX / 

CH 2 — CHO CH 2 

CH 2 -CH 2 CH 2 -C=CH 2 

d I I from 

CH 2 -CH 2 

e. | CHC0 2 H from | C=0 

HjC ~~c£ 2 H ^CH 2 

11-14 Write reasonable mechanisms for each of the following reactions. Support 
your formulations with detailed analogies insofar as possible. 

O O 

II II H 2 

a. H-C-C-H + NaOH —+ HOCH 2 C0 2 Na 

H H 

b. CH3-C-C-CH3 „ g ,! » CH 3 -C-CH 2 CH3 

Br OH O 

11-15 The following reactions represent "possible" synthetic reactions. Consider 
each carefully and decide whether or not the reaction will proceed as written. 
Show your reasoning. If you think side reactions would be important, write 
equations for each. 

excess 

a. CH 3 CH(OC 2 H 5 ) 2 + 2 NaOCH 3 CH 5 3 ° H > CH 3 CH(OCH 3 ) 2 +2 NaOC 2 H 5 

b. (CH 3 ) 3 CCOCH 2 CH 3 + KMn0 4 ^° > (CH 3 ) 3 COH + CH 3 CH 2 C0 2 K 

c. CH 3 CC1(CH 3 )CH 2 C1 + 2 NaOCH 3 CH3 ° H > CH 3 C(CH 3 )(OCH 3 )CH 2 OCH 3 + 2 NaCl 

d. 0=CH-C0 2 H + NaBH 4 CH3 ° H > 0=CH-CH 2 OH 

e. CH 2 =0 + CH 3 C0 2 CH 3 + NaOH > HC0 2 Na + CH 3 CH 2 OH + CH 3 OH 



11-16 Name each of the following substances by an accepted system : 

a. CH 3 OCH 2 C(OCH 3 ) 3 

b. CH 3 -CH-S0 3 Na 

I 
OH 

c. CH 3 C(OCH 3 ) 3 

d. [(CH 3 ) 3 CO] 3 Al 

e. CH 3 (CH 3 C0 2 )C(CN)(CH 3 ) 

/ (CH 3 ) 2 C=N-N(CH 3 ) 2 
g. (CH 2 ) 2 CHC(CH 3 )NOH 

11-17 Show how structures may be deduced for the two substances with respective 
infrared and nmr spectra as shown in Figure 11-2. 



■* : 










o 
o 
























8 




















o> 


<d 






' -' ~^~- 


o 
o 








00 


"* g TTT ' ' - - '• 






■-^Zpmm . ^ 


o 
o 










: 


l-l 


a. 


^hfiH^HI^piHl^BHffiBHH& 


'fi 


.g 


^^^^^^^^^^^^^^^^^^^^ 


8°. 

•o >> 


C o 


- 


a 




i 




> 
td 


^^ s : 


3 
o O" 
2 o 
2*1 




^^^^^^^^^^^^^^^^^^^^^^^^B 


8 




"sf^^^SM^SiM^^^^ff^^^Mifg^^^s^^^^Si^S^^^S^ 


o 










■* 


o 
o 




W^^fVw^^^B^^^^^s^WwWvfW^i^^M^f^f^- 


o 
o 


-t 


S»fcft(lH^WftSi^8Hfc^^(^^^w 


£N 




«r 


8 




iS^^^^^^^^^^^^^^^^^^^^^tf^^H 


o 
o 






§ 

•o 




. 








-r 








= L 


J..... 


~__ 




o I 


'«== 


=_ 




o pi 


.~£ 






so U 


{-p**"' 






("v 








3. 










fi 


o 




wilt 

1 




4> 











UOKSIUISUBJ} % 




chap 1 1 aldehydes and ketones I 300 

11-18 The mass spectra (Section 7-2B) of propanal and butanone both show strong 
peaks at mass 57. What is the ionic fragment that accounts for these peaks 
in the two cases? Propanone (acetone), on the other hand, has almost no 
peak at 57 but has a strong peak at mass 43. What does this suggest about 
the ease of bond breaking in carbonyl compounds ? 

11-19 Both periodic acid and lead tetraacetate are useful reagents for cleaving 
1,2-diols to aldehydes or ketones. *ra«.y-9,10-Decalindiol, however, does not 
react with periodic acid at all, although it is cleaved slowly by lead tetra- 
acetate. 




What general conclusions can you draw about the reaction pathways used by 
these two reagents? 



chap 12 aldehydes and ketones II 303 

Carbonyl groups often have a profound effect on the reactivity of their 
substituents. This is particularly evident when the substituents have hydrogen 
and halogen atoms on the carbon a to the carbonyl group, or when there is a 
double bond in the oc,/i position. In this chapter we shall consider first a 
number of very important synthetic reactions involving a hydrogens and later 
reactions of unsaturated and polycarbonyl compounds. 



12-1 halogenation of aldehydes and ketones 

Halogenation of saturated aldehydes and ketones usually occurs exclusively 
by replacement of hydrogens a to the carbonyl groups. The characteristics of 

o o 

II II 

CH3CCH3 + Cl 2 ► C1CH 2 CCH 3 + HC1 

chloroacetone 
Br 

<^ \=0 + Br 2 «• / \=0 + HBr 

2-bromocyclohexano ne 

such reactions are very different from those of the halogenation of alkanes 
(Chapter 3). Acetone has been particularly well studied and reacts smoothly 
with chlorine, bromine, and iodine. 

An important feature of the reaction is that acetone reacts at the same rate 
with each halogen. Indeed, the rate of formation of halogenated acetone is 
independent of the concentration of halogen, even at quite low halogen 
concentrations. Furthermore, halogenation of acetone is catalyzed by both 
acids and bases. The rate expressions for formation of halogenated acetone in 
water solution are 

e 
v = &[CH 3 COCH 3 ][OH] at moderate concentrations of OH e 

v = /t'[CH 3 COCH 3 ][H ffi ] at moderate concentrations of H® 

(The ratio of fc to k' is 12,000, which means that hydroxide ion is a much more 
effective catalyst than hydrogen ion.) 

To account for the role of the catalysts and the lack of effect of halogen 
concentration on the rate, the acetone must necessarily be slowly converted 
by the catalysts to an intermediate that reacts rapidly with halogen to give the 
products. This intermediate is a-methylvinyl alcohol, which is the unstable 
enol form of acetone. 

O OH e OH 

II H®(orOH e ) I Br 2 II e 

CH 3 -C-CH 3 — ► CH 3 -C=CH 2 -jt^- CH 3 -C-CH 2 -Br + Br 

enol 
(a-methylvinyl alcohol) 

O 

— ^ CH 3 -C-CH 2 Br + H ffi + Br e 



chap 12 aldehydes and ketones II 304 

As long as the first step is slow compared with the second and third steps, the 
rate will be independent of both the concentration of halogen and its nature, 
whether chlorine, bromine, or iodine. 

We shall now discuss each step in the reaction in more detail. First, there is 
the question of how the catalysts function to convert the ketone to its enol 
form. 



A. BASE-CATALYZED ENOLIZATION 

With a basic catalyst, the first step is removal of a proton from the a position 
to give the enolate anion [1]. Normally, C— H bonds are highly resistant to 



o 

II 

CH.-C- 



•CH, + OH 



slow 



CH, 



II e 

-C-CH,: 



[1] 



I 
CH,— C = CH, 



+ H,0 



attack by basic reagents, but removal of a hydrogen a to a carbonyl group 
results in the formation of a considerably stabilized anion with a substantial 
proportion of the negative charge on oxygen. As a result, hydrogens a to 
carbonyl groups have acidic character and can be removed as protons. In con- 
trast to the dissociation of many weak acids (e.g., CH 3 C0 2 H, H 3 B0 3 , HF), 
the acidic proton on carbon is removed slowly and equilibrium between the 
ketone and its enolate anion [1] is not established rapidly. The reverse reac- 
tion is also slow and, as a result, the enolate anion has ample time to add a 
proton to oxygen to form the enol (this process is at least 10 10 times faster 
than conversion to the ketone). 



O 

I: 

CH,-C— CH, 



+ H a 



fast 



CH 



OH 
I 
-C=CH 2 



Both the enol and the enolate anion can combine rapidly with halogen to give 
the a-halo ketone. 



O 



OH 

I 



+ Br, 



fasi 



O 



ffi OH 

fast II 
► CH 3 -C-CH 2 Br + Br' 



O 

a ( - HBr) II 

e > CH,-C-CH,Br 



The slowest step in the whole sequence is the formation of the enolate anion, 
and the overall rate is thus independent of the halogen concentration. 



B. ACID-CATALYZED ENOLIZATION 

Catalysis of the enolization of acetone by acids involves, first, oxonium ion 
formation and, second, removal of an a proton with water or other proton 
acceptors. 



sec 12.1 halogenation of aldehydes and ketones 305 



O ®OH 

II _ fast II 

CH 3 — C— CH 3 + H ffi ; CH 3 — C— CH 3 

®OH OH 

B f "^ slow I $ 

CH 3 -C-CH/+ H 2 > CHj-C-CHj +H 3 

This differs from base-catalyzed enolization in that the enol is formed 
directly and not subsequently to the formation of the enolate anion. Also, 
proton addition to the carbonyl oxygen greatly facilitates proton removal 
from the a carbon by the electron-attracting power of the positively charged 
oxygen. Nevertheless, the rate of enolization (or halogenation) is determined 
by this last step. 

The characteristics of acid- and base-catalyzed enolization, as revealed 
by the halogenation of acetone, are displayed in a wide variety of other, 
usually more complicated, reactions. For this reason the halogenation reaction 
has been considered in rather more detail than is consistent with its intrinsic 
importance as a synthetic reaction. 



C. HALOFORM REACTION 

The preceding discussion on the halogenation of ketones is incomplete in 
one important respect concerning base-induced halogenation. That is, once 
an a-halo ketone is formed, the other hydrogens on the same carbon are 
rendered more acidic by the electron-attracting effect of the halogen and are 
replaced much more rapidly than the first hydrogen. 



OOO 

II slow H fast II 

CH3-C-CH3 ' CH 3 -C-CH 2 Br > CH 3 -C-CHBr 2 

fast 

O 

II 

CH 3 -C-CBr 3 

The result is that if the monobromo ketone is desired, the reaction is best 
carried out with an acidic catalyst rather than a basic catalyst. A further 
complication in the base-catalyzed halogenation of a methyl ketone' is that the 
trihalo ketone formed is readily attacked by base with cleavage of a carbon- 
carbon bond. 

o o e o 

11 * j^ ... in^ s .ow^ CHj _ c // + 



OH OH 

:CBr 3 e . CH 3 COf + HCBr 3 

bromoform 

This sequence is often called the haloform reaction because it results in the 
production of chloroform, bromoform, or iodoform, depending on the 



chap 12 aldehydes and ketones II 306 

halogen used. The haloform reaction is a useful method for detecting methyl 
ketones, particularly when iodine is used because iodoform is a highly 
insoluble, bright yellow solid. The reaction is also useful for the synthesis of 
carboxylic acids when the methyl ketone is more available than the corres- 
ponding acid. Because halogens readily oxidize alcohols to carbonyl com- 

o 

i^ II l.Br 2 ,OH e , H 2 r-. 

P-C-CH 3 rS5 ► [>-C0 2 H + CHBr 3 

85% 

pounds it follows that a positive haloform test will also be given by alcohols 
containing the — CHOHCH 3 group. 



12-2 reactions of end ate anions 



A. THE ALDOL ADDITION 

Many of the most important synthetic reactions of carbonyl compounds 
involve enolate anions, either as addends to suitably activated double bonds 
or as participants in nucleophilic substitutions. When the addition is to a 
carbonyl double bond, it is often called an aldol addition. S N reactions of 
enolate anions are considered in Chapter 13 with regard to synthetic applica- 
tions. 

The course of the aldol addition is typified by the reaction of acetaldehyde 
with base, and, if carried out under reasonably mild conditions, gives 
/?-hydroxybutyraldehyde (acetaldol). If the mixture is heated, the product is 
dehydrated to crotonaldehyde (2-butenal). 







OH 


2 CH3CHO 


dilute NaOH 

> 


1 
CH 3 CHCH 2 CHO 


OH 






CH3-CH- 


-CH,-CHO 


► CH 3 CH= 



Formation of the enolate anion by removal of an a hydrogen by base is the 
first step in the aldol addition. 



o 

e^ II 

HO +CH,-C-H 



O o° 

e II I 

:CH,— C — H « ► CH 2 =C— H 



The anion then adds to the carbonyl bond in a manner analogous to the 
addition of cyanide ion in cyanohydrin formation (Section 1 1 ■ 4A). You would 
expect from consideration of the two resonance forms of the enolate anion 
that addition might take place in either of two ways: The anion may attack 



sec 12.2 reactions of enolate anions 307 



to form either a carbon-carbon or a carbon-oxygen bond, leading to the 
aldol [2] or to a-hydroxyethyl vinyl ether [3]. Although the formation of [3] is 



O a O 

I II 

CH 3 -C-CH 2 -CH 
I 

H 



H® 



OH 

i CH,— C— CH,CHO 
I 
H 

[2] 



CH, 



O 

II 
-C— H + 



CH, 



O e 
I 
=C-H <- 



O 

e II 

-> :CH 2 — C- 



O fc 



CH, — C— O— CH=CH, 5P 
I 
H 



OH 

I 



=t CH,— C— O— CH=CH, 



I 
H 



[3] 



mechanistically reasonable, it is much less so on thermodynamic grounds. 
Indeed, AH for the formation of [3] from acetaldehyde (calculated from 
vapor-state bond energies) is +20 kcal. 

The equilibrium constant is favorable for the aldol addition of acetaldehyde, 
as in fact it is for most aldehydes. For ketones, however, the reaction is much 
less favorable. With acetone, only a few percent of the addition product 
" diacetone alcohol " [4] is present at equilibrium. 



o 

II 

2CH,CCH, 



OH 



OH O 

I II 

CH3 C~~ CH2 C CH 3 



I 

CH, 



[4] 



This is understandable on the basis of steric hindrance and the fact that the 
ketone-carbonyl bond is about 3 kcal stronger than the aldehyde-carbonyl 
bond. Despite the unfavorable equilibrium constant, it is possible to prepare 
diacetone alcohol in good yield with the aid of an apparatus such as shown in 
Figure 12-1. 

The acetone is boiled and the hot condensate from the reflux condenser 
flows back through the porous thimble over the solid barium hydroxide 
contained therein and comes to equilibrium with diacetone alcohol. The 
barium hydroxide is retained by the porous thimble and the liquid phase is 
returned to the boiler where the acetone (boiling 110° below diacetone alcohol) 
is selectively vaporized and returned to the reaction zone to furnish more 
diacetone alcohol. 

The fundamental ingredients in the key step in aldol addition are an elec- 
tron-pair donor and an electron-pair acceptor. In the formation of acetaldol 



chap 12 aldehydes and ketones II 308 



porous thimble 

containing 

Ba(OH) 2 




Figure 12-1 Apparatus for preparation of diacetone alcohol. 



and diacetone alcohol, both roles are played by one kind of molecule, but 
there is no reason why this should be a necessary condition for reaction. 
Many kinds of mixed additions are possible. Consider the combination of 
formaldehyde and acetone: Formaldehyde cannot form an enolate anion 
because it has no a hydrogens, but it should be a particularly good electron- 
pair acceptor because of freedom from steric hindrance and the fact that it has 
an unusually weak carbonyl bond (166 kcal vs. 179 kcal for acetone). Acetone 
forms an enolate anion easily but is relatively poor as an acceptor. Conse- 
quently, the addition of acetone to formaldehyde should and does occur 
readily. 



o 

CH,-C-CH, 



OH 



o 

II 



+ CH 3 — C-CH 2 CH 2 OH 



The problem is not to get addition, but rather to keep it from going too far. 
Indeed, all six a hydrogens can be easily replaced by — CH 2 OH groups. 



o 

II 

CH,-C-CH, + 



6 

OH 



(HOH 2 C) 3 C- 



O 

II 

-c- 



C(CH 2 OH) 3 



To obtain high yields of the monohydroxymethylene derivative, it is usually 
necessary to use an apparatus such as shown in Figure 12-2. The scheme here 
is to have the addition take place in the presence of a large excess of acetone 
to assure favorable formation of the monoadduct. The reaction is then 



sec 12.2 reactions of enolate anions 309 

quenched and the acetone separated and used again. This is achieved by 
boiling rapidly a solution of acetone containing an excess of a nonvolatile 
weak organic acid, such as succinic acid (CH 2 )2(C0 2 H) 2 . The vapor is 
condensed and then mixed with a very slow trickle of an alkaline formalde- 
hyde solution. The addition then occurs while the mixture flows down over a 
column of glass beads, and gives the monoadduct because the acetone is 
present in great excess. The reaction stops and reversal is prevented when the 
alkali is neutralized by the acid in the boiler. The excess acetone is revaporized 
and sent up to react with more formaldehyde. 

Aldol addition products can be converted to a variety of substances by 
reactions that have been discussed previously. Of particular importance is the 
dehydration of aldols to a,/?-unsaturated carbonyl compounds, which occurs 
spontaneously in the presence of acid. The formation of crotonaldehyde by 
dehydration of acetaldol has already been mentioned in this section. Another 
example is the dehydration of diacetone alcohol to mesityl oxide. 



CH 3 O 

I II 

CH 3 -C-CH 2 -C-CH 3 

I 
OH 



H e 



( - H 2 0) 



CH 3 
\ 
C=CH- 

CH, 



O 

II 

-c- 



CH, 



4-methyl-3-penten-2-one 
(mesityl oxide) 



Figure 12-2 Apparatus for preparation of monohydroxymethylene aldol- 
addition products from formaldehyde and carbonyl compounds with more 
than one a hydrogen. 




cold formaldehyde 
solution containing 
dilute sodium 
hydroxide 



volatile carbonyl 
compound containing 
a weak nonvolatile 
organic acid 



chap 12 aldehydes and ketones II 310 
B. NUCLEOPHILIC SUBSTITUTION INVOLVING ENOLATE ANIONS 

An enolate anion can be formed in good yield from a ketone and a power- 
fully basic reagent, such as sodium or potassium amide, provided that the 
ketone has an a hydrogen. The enolate anion so formed can theoretically 
undergo S N reactions with an alkyl halide in two different ways. Thus, for 
f-butyl methyl ketone and methyl iodide, we could have the reactions shown 
in Scheme I, which differ only in the position of attack at the enolate anion. 



H 3 C O 
CH ,_c-C-CH, KNH > 



( - NH 3 ) 



H,C 



H 3 C O 

I II e 

CH 3 — C— C — CH 2 : 



-+ CH,— C— C=CH 2 



H,C 



CH 3 -C-C-CH 




C-alkylation 



O-alkylation 



CH, 



CH, 



SCHEME I 



The possibility of the enolate anion's acting as though its charge were 
effectively concentrated on carbon or on oxygen was discussed in the pre- 
vious section in connection with aldol addition (Section 12-2A). However, the 
situation there is actually quite different from the one here, because the 
reaction on oxygen was indicated to be thermodynamically unfavorable over- 
all (AH = + 20 kcal). However, O- and C-alkylation of the anion are both 
thermodynamically favorable. Furthermore, alkylation, unlike the aldol 
addition, is not reversible under ordinary conditions, and therefore the 
O-alkylation product is not expected to go over to the C-alkylation product, 
even though the latter is considerably more stable. 

Whether C- or O-alkylation occurs often depends on the reactivity of the 
halide; the lower the S N 2 reactivity of the halide, the more C-alkylation is 
favored. 

Useful alkylation procedures for the preparation of a-substituted ketones 
will be discussed below and in Chapter 13. 



unsaturated carbonyl compounds 

The combination of a carbonyl function and a double bond in the same 
molecule leads to exceptional properties only when the groups are close to one 
another. The cumulated and conjugated arrangements are of particular 
interest. We shall consider first the conjugated or a,/?-unsaturated carbonyl 



sec 12.3 a,(3-unsaturated aldehydes and ketones 311 

compounds, because their chemistry is closely related to that of the substances 
already discussed in this chapter and in Chapter 11. 



12-3 a, ^-unsaturated aldehydes and ketones 

The most generally useful preparation of a,/?-unsaturated carbonyl com- 
pounds is by dehydration of aldol addition products, as described in the 
previous section. Conjugation of the carbonyl group and double bond has a 
marked influence on spectroscopic properties, particularly on ultraviolet 
spectra, as the result of stabilization of the excited electronic states which, for 
n-*n* transitions, can be described in terms of important contributions by 

\e I I e 
polar-resonance structures such as C — C = C — O (see also Sections 7-5 and 

11-2B). / 

The effect of conjugation is also reflected in nmr spectra. The protons on 
the P carbon of a,/?-unsaturated carbonyl compounds usually come at 0.7 to 
1 .7 ppm lower than ordinary olefmic protons. The effect is smaller for the a. 
proton. 

cc,/5-Unsaturated carbonyl compounds may undergo the usual addition and 
condensation reactions at the carbonyl group, such as cyanohydrin and 
hydrazone formation and addition of organometallic compounds. These 
reactions, however, may be complicated, if not overshadowed, by " 1,4 addi- 
tion " (conjugate addition, Section 6-2) which gives as the overall result addi- 
tion to the carbon-carbon double bond. The balance between the two modes 
of reaction is so delicate that relatively small changes in steric hindrance are 
sufficient to cause one or the other process to predominate. 

With hydrogen cyanide, cyanohydrin formation is usually more rapid than 
1,4 addition, and if the equilibrium is favorable, as with most aldehydes, only 
1,2 addition is observed. 

OH 

e | 

CH 3 -CH=CH-CHO + HCN OH > CH 3 — CH=CH-C-H 

I 

With ketones, cyanohydrin formation is less favorable, and 1,4 addition 

results to give an enol intermediate which is unstable with respect to the 

jS-cyano ketone. 

O 
II 
CH,= CH-C-CH, + HCN 



OH O 

I k II 

CH 2 =CH-C-CH 3 ;= ' N=C-CH 2 -CH 2 -C-CH 3 

C=N 

Addition of hydrogen halides to a,/?-unsaturated aldehydes, ketones, and 

acids occurs 1,4 and places the halogen on the /? carbon atom. The mechanism 

of this reaction is illustrated for the reaction of hydrogen chloride with pro- 



chap 12 aldehydes and ketones II 312 

penal (acrolein). 1,2 Addition to the carbonyl group would lead to an unstable 



o 
II 
CH,= CH-C-H + H s 



ffi OH 
II 
CH, = CH-C-H <- 



OH 
B CH, — CH = C — H 



CI° 



OH 
C1CH, - CH=C - H -^r- 



O 

II 
-C-H 



a-halo alcohol (Section 1 1 -4E). 

0,y-Unsaturated aldehydes and ketones are usually relatively difficult to 
synthesize and are found to rearrange readily to the a, jS-un saturated isomers, 
particularly in the presence of basic reagents. (See Exercise 12-14.) 



CH,=CH-CH,-CHO 



base 



+ CH,-CH=CH-CHO 



12-4 ketenes 

Substances with cumulated carbonyl and carbon-carbon double bonds, 

C = C = 0, are called ketenes and, as might be expected, have interesting 

and unusual properties. 

There are few general preparations for ketenes although several special 
methods are available for ketene itself. The most convenient laboratory prep- 
aration is to pass acetone vapor over a coil of resistance wire heated electri- 
cally to a dull red heat. Air is excluded to avoid simple combustion. 



750° 



+ CH,=C=0 + CH. 



O 

II 

CH 3 -C— CH 3 

ketene 
bp-56° 

Ketene is a very useful acetylating agent for ROH and RNH 2 compounds. 
It reacts rapidly, and since the reactions involve additions, there are no by- 
products to be separated. 

p 



CH,= C=0 



H 2 



CH 3 C0 2 H 



CH 3 CH 2 OH 



+ CH, 



\ 
OH 

O O 

II II 

-C-O-C-CHj 

o 



* CH 3 — C-0-CH 2 CH 3 



O H 

CH 3 NH 2 II I 
— ^-> CH,-C-N-CH, 



sec 12.5 1,2-dicarbonyl compounds 313 

The considerable convenience of ketene as an acetylating agent would 
make it an excellent candidate for commercial sale in cylinders except for the 
fact that the substance is unstable with respect to formation of a dimer known 
as diketene. The dimer is itself a highly reactive substance with such unusual 
characteristics that its structure was not firmly established until the early 
1950's, some 40 years after it was first prepared. 

CH 2 =C=0 C-O 

i i ► I I 

H 2 C=C=0 H 2 C— C 

O 

diketene 
(vinylaceto-£-lactone) 



polycarbonyl compounds 

12-5 1 ,2-dicarbonyl compounds 

The structures of some typical and important members of this class are shown. 

00 9 9 9 9 

n II ii II II II 

H-C-C-H CH3-C-C-CH3 C 6 H 5 -C-C-C 6 H 5 

glyoxal biacetyl benzil 

(ethanedial) (2, 3-butanedione) (diphenylethanedione) 

Most of the 1,2-dicarbonyl compounds are yellow in color. Glyoxal is 
unusual in being yellow in the liquid state, but green in the vapor state. It has 
very reactive aldehyde groups. 

Glyoxal undergoes an internal Cannizzaro reaction (Section 11-4H) with 
alkali to give glycolic acid. 

OO OH OH o 

II II e I h® I // 

H-C-C-H + OH » H-C-COf ► H-C-C 

I I \ 

H H OH 

An analogous reaction occurs with benzil, except that the migrating group 
is phenyl rather than hydrogen and this results in a rearrangement of the 
carbon skeleton. This is one of a very few carbon-skeleton rearrangements 
brought about by basic reagents, and is known as the benzilic acid rearrange- 
ment. 

OO OH OH 

II II e I H ® I // 

C 6 H 5 -C-C-C 6 H 5 + OH ► C 6 H 5 -C-CO? ► C 6 H 5 -C-C x 

benzil C 6 H 5 C 6 H 5 OH 

benzilic acid 



chap 12 aldehydes and ketones II 314 

12-6 1 ,3-dicarhonjl compounds 

Most of the important properties of 1,3-dialdehydes, aldehyde ketones, and 
diketones that are characteristic of the 1 ,3 relationship are well illustrated by 
2,4-pentanedione (acetylacetone). This substance is unusual in existing to the 
extent of 85 % as the enol form. 

o o 

II i II 

CH-C-s-CH 2 -C-CH 3 

2, 4-pentanedione 

O O OH O 

II 1 I II 



,c „c. 



~CH- - C H 3 

15% 85% 

The nmr spectrum of 2,4-pentanedione (Figure 12-3) is very informative 
about the species present in the pure liquid. Resonance lines for both the keto 
and enol forms can be readily distinguished. The keto form has its CH 2 reson- 
ance at 218 Hz and its CH 3 resonances at 120 Hz, whereas the enol form 
shows its CH 3 and vinyl-CH resonances at 110 Hz and 334 Hz, respectively. 
The enol-OH proton comes at the very low field value of 910 Hz with respect 
to tetramethylsilane. That each form may be observed separately indicates 
that the lifetime of each form is longer than 0.1 sec at room temperature (see 
Section 7-6D). However, when a basic catalyst is added, the lines broaden 



Figure 12-3 Nuclear magnetic resonance spectrum of 2,4-pentanedione, 
CH 3 COCH 2 COCH 3 , at 60 MHz. Calibrations are relative to tetramethylsilane. 





800 


600 


400 


200 


ii/. 










enol 










MC.H ,-.... 










kc-ni 








IliBiifliii'iilSi 


(ch»-), j; 








=c— 


iiiiilpliill 


enol 








-CH- 


-OH 

i 


' : 


i 


.. i 

1 J 


_ ._ _ 







16.0 12.0 8.0 4.0 Oppm 



sec 12.6 1,3-dicarbonyl compounds 315 

considerably; when the mixture is heated, the lines coalesce to an average 
spectrum as expected for rapid equilibration. 

The very large chemical shift of the enol-OH proton is the consequence of 
internal hydrogen bonding involving the carbonyl group. This type of hydro- 
gen bonding is also important in stabilizing the enol form, as evidenced by an 
increase in the percentage of enol in those solvents, such as hexane, that can- 
not effectively solvate the ketone groups of the keto form. 

-hydrogen bond 



o 

1 II 

H 3 C^ C CH 3 

I 
H 

The enol form is also considerably stabilized by resonance which, in turn, 
increases the strength of the hydrogen bond. Such stabilization is, of course, 
not possible for the keto form. 

o /H "o t „ »o^ H "-o e 

I II II I 

h 3 c" C \h C ^ch 3 h 3 c /C ^c^ v ch 3 

The term conformer is often used to designate one of two or more stereo- 
chemical arrangements that are interconverted by rotation about single bonds. 
This is usually so rapid that it prevents the isolation of discrete conformers, 
except at very low temperatures (see Section 3-4B). The term tautomer is 
employed in exactly the same sense for structural isomers in rapid equilibrium, 
such as the keto and enol forms of 2,4-pentanedione. When tautomerization 
involves only the shift of a proton, as in the acetaldehyde-vinyl alcohol 
equilibrium (Section 10-8), it is sometimes called a prototropic change. While 
in principle the only difference between the rapid interconversion of 2,4- 
pentanedione on the one hand and the slow isomerization of 1,4-pentadiene 
on the other is a matter of relative reaction rate, rightly or wrongly, the term 
tautomeric change is usually applied only to the more rapid process. 

2,4-Pentanedione is moderately acidic with K HA ~ 1CT 9 compared with 
K HA of 10~ 5 for acetic acid and 10~ 16 for ethanol, the charge derealization 
in the anion being responsible. 

o o o e o o o e 

II II « > I II « — ► II I 

H 3 C^ ^CH^ V CH 3 H 3 C^ V CH^ ^CH 3 H 3 C^ ^CH^ ^CH 3 

Alkali metal salts of the compound can be alkylated with alkyl halides of 
good S N 2 reactivity and generally give C-alkylation. A number of syntheti- 
cally important alkylations of other 1,3-dicarbonyl compounds are discussed 
in Chapter 13. 

Polyvalent metal cations often form very stable and slightly polar enolate 
salts with acetylacetone, better known as metal chelates. Cupric ion is a parti- 



chap 12 aldehydes and ketones II 316 

cularly good chelating agent. The corresponding beryllium compound is a 



H 

I 

I: 
O. 

':cu; 
o-' -o 

il 



il 

o 



,c*. 



-A 



I 

H 



cupric acetylacetonate (dark blue) 
(Cu" salt of 2, 4-pentanedione) 



CH, 



further example of a metal chelate; it melts at 108°, boils at 270°, and is 
soluble in many organic solvents. 



summary 

Aldehydes and ketones [5] undergo a number of reactions at the a position 
involving either the enolate anion [6] or the enol [7]; [6] results only from the 
action of a base, while [7] can be generated by the action of either an acid or a 
base. The rates of interconversion of [7] and [5], not their equilibrium con- 
centrations, are affected by acid or base. 



(-H®K, 

O 
II 
R-CH 2 — C-Z 

[5] 



O 
e II 
R— CH— C— Z 



I 
-» R— CH=C— Z 



[6] 



(+H«) 



R-CH, 



ffi OH 
-C-Z 



( - H®)_ 



\(+ H») 

OH 
I 
R-CH=C-Z 

[7] 



a-Halogenation can proceed via [6] or [7] with both aldehydes and ketones. 
Methyl ketones in basic solution undergo trisubstitution and this is followed 
by chain cleavage (the haloform reaction). 



o 

II 

RCJHL C CH3 



o 

II 



-► RCH 2 — C — CX 3 



-► RCH,-CO,H + CHX, 



Aldehydes with a hydrogens undergo aldol addition in base to give products 
which often eliminate water to give qc,/?-unsaturated aldehydes. Although the 
equilibrium is normally unfavorable for the corresponding condensation of 
two molecules of a ketone, the reaction can nonetheless often be made to 
occur (Figure 12-1). 



summary 317 



e I 

RCH— CHO ► RCH,CH — CH— CHO 

I 
R 



OH 



RCH 2 CH-CH-CHO ► RCH 2 CH=C-CHO 

I I 

R R 

Ketones are alkylated (predominantly at carbon) through reaction of the 
enolate anions with alkyl halides. 

o o 

e II II 

R— CH— C— R ► R— CH— C— R 

I 
R' 

a,/MJnsaturated aldehydes and ketones absorb strongly in the ultraviolet 
and the protons on their /J-carbon atoms appear at lower field in the nmr than 
other alkenic protons : They are usually subject to 1,4-addition reactions of the 
type RCH=CHCHO + HZ -> RCHZCH 2 CHO. 

Ketenes contain cumulated double bonds and react rapidly with hydroxylic 
(and amino) compounds. 

O / o 

II II 

R 2 C = C = + ZOH ► R 2 CHCOZ \Z=H — , RC— , R- 



Ketene itself, which is prepared in the laboratory by the pyrolysis of ace- 
tone, readily dimerizes to give diketene, a /^-lactone. 

1,2-Dicarbonyl compounds include glyoxal (OHCCHO), biacetyl 

o o o o 

II II II II 

(CH3C-CCH3), and benzil (C 6 H 5 C-CC 6 H 5 ). In basic solution, glyoxal 
undergoes an internal Cannizzaro reaction and benzil undergoes a carbon- 
skeleton rearrangement to give benzilic acid, (C 6 H 5 ) 2 COHC0 2 H. 

Two important characteristics of 1,3-dicarbonyl compounds are their 
acidity and their tendency to enolize. Stabilization of the anion and enol can 
be ascribed to resonance stabilization and hydrogen bonding, respectively. 

OO O^O 

li e i| I || 

R— Ci- .-C— R R— C*. ^C— R 
CH^ V CH 

The keto-enol forms are called tautomers (structural isomers in rapid 
equilibrium). Enols form complex salts called chelates with many metal 
cations. 



chap 12 aldehydes and ketones II 318 

exercises 

12-1 At what point would the system shown in Figure 12-1 cease to produce more 
diacetone alcohol? What would happen if some barium hydroxide were to 
get through a hole in the thimble and pass into the boiler? Why is barium 
hydroxide more suitable for the preparation than sodium hydroxide? 

12-2 What would be the products expected from aldol additions involving pro- 
panal, 2,2-dimethylpropanal, and a mixture of the two aldehydes? 

12-3 Predict the principal products to be expected in each of the following reac- 
tions; give your reasoning: 

a. CH 3 CHO + (CH 3 ) 2 CO NaOH > 

b. (CH 3 ) 2 C(OH)CH 2 COCH 3 ^^U 

c. CH 2 + (CH 3 ) 3 CCHO NaOH > 



12-4 Show how the following compounds can be synthesized from the indicated 
starting materials by way of aldol-addition products : 

a. CH 3 -CH-CH 2 -CH 2 from acetaldehyde 

I I 

OH OH 

b. CH 3 CH=CH-CH 2 OH from acetaldehyde 

c. (CH 3 ) 2 CHCH 2 CH 2 CH 3 from acetone 

O 

II 

d. CH 3 CH— C— CH 2 CH 3 from propionaldehyde 

I 
CH 3 



12-5 Although trivial names for organic compounds are slowly passing out of 
common use a few such names remain; for example, diacetone alcohol 
(Section 12-2A), benzilic acid (Section 12-5), and benzoin (Section 24-6 and 
Exercise 12-27). Provide the IUPAC names for these three compounds and 
suggest a reason for the reluctance of chemists to abandon the trivial names 
completely. 

12-6 Aldol additions also occur in the presence of acidic catalysts. For example, 
acetone with dry hydrogen chloride slowly yields (CH 3 ) 2 C=CHCOCH 3 
(mesityl oxide) and (CH 3 )2C=CHCOCH=C(CH 3 ) 2 (phorone). Write 
mechanisms for the formation of these products giving particular attention to 
the way in which the new carbon-carbon bonds are formed. Review Sections 
4-4H and 11-4B. 

12-7 The following reactions represent "possible" synthetic reactions. Consider 
each carefully and decide whether or not the reaction will proceed as written. 
Show your reasoning. If you think side reactions would be important, write 
equations for each. 



exercises 319 



a. CH3COCH3 + 6 Br 2 + 8 NaOH ► 2 CHBr 3 +Na 2 C0 3 + 6 NaBr+ 6 H 2 

b. CH3CHO + NaNH 2 + (CH 3 ) 3 CC1 ► (CH 3 ) 3 CCH 2 CHO + NH 3 + NaCl 

c. (CH 3 ) 2 CHCOCH 3 + CH 2 = Ca(OH>2 . (CH 3 ) 2 C(CH 2 OH)COCH 3 

OH e 

d. CH3CHO + CH 3 C0 2 C 2 H s CH 3 CHCH 2 C0 2 C 2 H 5 

OH 
12-8 Write equations for a practical laboratory synthesis of each of the following 
substances, based on the indicated starting materials (several steps may be 
required). Give reagents and conditions. 

O O 

P^CirO-CH 2 ^ from [>-C-CH 3 



a 



1 

b. CH 2 =CHCOCH 3 from CH 3 COCH 3 

c. (CH 3 ) 3 CC0 2 H from (CH 3 ) 2 C(OH)C(CH 3 ) 2 OH 

d. (CH 3 ) 3 CCOC(CH 3 ) 3 from CH 3 CH 2 COCH 2 CH 3 

e. (CH 3 ) 2 CHCH 2 CH(CH 3 ) 2 from CH 3 COCH 3 
/. (CH 3 ) 3 CCH 2 CH 2 CH 3 from (CH 3 ) 3 CCOCH 3 

12-9 Give for each of the following pairs of compounds a chemical test, preferably 
a test tube reaction, that will distinguish between the two compounds. (You 
may wish to review Section 1 1 -41 in connection with some of these.) 

a. CH 3 COCH 2 CH 2 COCH 3 and CH 3 COCH 2 COCH 3 

b. (CH 3 CH 2 CH 2 CH 2 ) 2 CO and [(CH 3 ) 3 C] 2 CO 

c. (C 6 H 5 ) 2 CHCH 2 CHO and (C 6 H 5 CH 2 ) 2 C=0 

d. C 6 H 5 COCOC 6 H 5 and C 6 H 5 COCH 2 COC 6 H 5 

e. CH 3 CH=C=0 and CH 2 =CH-CH=0 

12-10 How might spectroscopic methods be used to distinguish between the two 
isomeric compounds in the following pairs : 

a. CH 3 CH=CHCOCH 3 and CH 2 =CHCH 2 COCH 3 

b. C 6 H 5 COCH 2 COC 6 H 5 and /,-CH 3 C 6 H 4 COCOC 6 H 5 

c. CH 3 CH=C=0 and CH 2 =CH-CHO 



0> 



and CH 3 COCH 2 CH 3 



12-11 Sketch out an energy profile with the various transition states for the reaction 
CH3COCH3 + OH e + Br 2 -> CH 3 COCH 2 Br + H 2 + Br e described in 
Section 12-1A, using the general procedure of Section 8-9. Note that in this 
case the enol form, unlike the carbonium ion in Figure 8-2, is a rather stable 
intermediate. 

12-12 Interpret the proton nmr spectra given in Figure 12-4 in terms of structures 
of compounds with the molecular formulas C 6 H 10 O and C 9 H s O. The latter 
substance has a phenyl (C 6 H 5 ) group. 



chap 12 aldehydes and ketones II 320 




ppm 





1 

3f,0 


i 
500 


450 


i l 

4uo ;i r )0 ii/. 


—r- 




C.,H,0 


-'\ 


7 

/ 

I 


r ^j~ j ~~^~ 


fllil 








t^^^^/t: 


...J 


*J 






rV fWl 
V» Iflk 








y " * -" -fj—r~r 

i i i i 



9.0 



8.0 



7.0 



6.0 ppm 



Figure 12-4 Proton nmr spectra at 60 MHz with TMS as standard. See Exer- 
cise 12-12. 



12-13 Calculate AH for vapor-phase 1,2 and 1,4 additions of hydrogen cyanide to 
methyl vinyl ketone. Write a mechanism for 1,4 addition that is consistent 
with catalysis by bases and the fact that hydrogen cyanide does not add to 
an isolated carbon-carbon double bond. 

12-14 Write a reasonable mechanism for the base-induced rearrangement of 
3-butenal to 2-butenal. Why is 2-butenal the more stable isomer? 



12-15 Write reasonable mechanisms for the reaction of ketene with alcohols and 
amines. Would you expect these reactions to be facilitated by acids or bases 
or both? 



exercises 321 

12-16 The following structures have been proposed or could be proposed for 
diketene. Show how infrared, ultraviolet, and nmr spectroscopy might be 
used to distinguish between the possibilities. (If necessary, review Chapter 7.) 



CH 2 

V 

1 


-0 

1 


H 2 C- 


1 

- c % ( 


[i] 


I 


CH 2 

V 

1 


-o 

1 


1 
O- 


1 

-s 




M 




o 



o 



CH, 



\ 



C-CH 2 

I I 



H,C-C. 



\ 



[i>] 



O 



H 2 C-C 
H 2 C— C, 

[vi] 



// 



P 



V 



HO 



\ 



C=CH 



H,C— C 



V 



[iii] 



// 



H 2 C-C 

I I 
HC=C 



OH 



[vii] 



CH 3 — C— CH=C=0 (the favored structure for many years) 
[ix] 



HO 



C=CH 

I I 
HC=C 



OH 



[iv] 



OH 



HC=C 

I I 
HC = C 



\ 



OH 



[viii] 



12-17 2,6-Bicyclo[2.2.2]octanedione exhibits no enolic properties. Explain. 

O^ y^ JO 




12-18 What experiments might be done to prove or disprove the following mecha- 
nism for rearrangement of glyoxal to glycolic acid ? 



HC — CH 



O 

II e 
H— C— C=0 



I 

H-C=C=0 



OH 

1 

c=c= 



=o 



OH i 
-^°. H-C-C 7 ' 



H 



X OH 



12-19 Write a mechanism analogous to that for the Cannizzaro reaction for the 
benzil-benzilic acid transformation. Would you expect the same type of 
reaction to occur with biacetyl ? Why or why not ? 



12-20 Account for the considerable K H a of the enol of acetylacetone with respect 
to ethyi alcohol. Arguing from the proportions of each at equilibrium, 
which is the stronger acid, the keto or the enol form of acetylacetone? 
Explain. 



chap 12 aldehydes and ketones II 322 



400 



300 



200 



t 
US. 1 ppni 



100 



11/ 



C,„H,,.0, 



I 



8.0 



6.0 



4.0 



2.0 



ppm 



*K 



(:i00 



! I : 

400 



(i:hi:h=C-0) 



ion 



1 



AVW*«»V^WN«Mty^*wrtjM>*»v<»fSU»v^^ 



Hz 



Owiv 



10.0 



8.0 



4.0 



ppm 



Figure 12-5 Proton nmr spectra at 60 MHz with TMS as standard. See Exer- 
cise 12-21. 



12-21 Interpret the proton nmr spectra shown in Figure 12-5 in terms of structures 
of the compounds with molecular formulas C10H10O2 and (CH 3 CH = 
C=0) 2 . See also Exercise 12-16. 



12-22 Write a reasonable mechanism, supported by analogy, for the acid-catalyzed 
dehydration of 2,4-pentanedione to 2,5-dimethylfuran. 

H,C^O CH, 

\J 
2, 5-dimethylfuran 



exercises 323 



300 : 



200 



— j — 

Mil'! 



Hz 



Wr- 



h 



O.H-OCl 



:H 



ppm 



_! — 

200 : 



n 

n 11/ 



5.0 



J J 
rr'r ' 



:(' H-()Ht: 



I 
4.0 



r 



3.0 



2.0 



1.0 



t 



ppm 



Figure 12-6 Proton nmr spectra at 60 MHz with TMS as standard. See Exer- 
cise 12-23. 



12-23 The nmr spectra of two compounds of formulas C 4 H 7 OCl and CUH 7 OBr 
are shown in Figure 12-6. Assign to each compound a structure that is con- 
sistent with its spectrum. Show your reasoning. Give a concise description of 
the chemical properties to be expected for each compound. 



12-24 When the formation of /3-hydroxybutyraldehyde is carried on in DjO con- 
taining OD e , using moderate concentrations of undeuteriated acetaldehyde, 
the product formed in the early stages of the reaction contains no deuterium 



chap 12 aldehydes and ketones II 324 

bound to carbon. Assuming the mechanism shown in Section 12-2A to be 
correct, what can you conclude as to which step in the reaction is the slow 
step? What would then be the kinetic equation for the reaction? What 
would you expect to happen to the kinetics and the nature of the product 
formed in D 2 at very low concentrations of acetaldehyde ? 

12-25 1,2-Cyclopentanedione exists substantially as the monoenol, whereas 
biacetyl exists as the keto form. Suggest explanations for this behavior that 
take into account possible conformational differences between the two sub- 
stances. How easily would you expect dione [8] to enolize? Why? 



O 




o 

[8] 

12-26 A detailed study of the rate of bromination of acetone in water, using acetic 
acid-acetate buffers, has shown that 

v={6x 10" 9 + 5.6x 10- 4 [H 3 O®]+1.3x 10- 6 [CH 3 CO 2 H] + 7[OH e ] 
+ 3.3 x 10- 6 [CH 3 CO 2 e ]+ 3.5 x 10- 6 [CH 3 CO 2 H][CH 3 CO 2 e ]} 
[CH 3 COCH 3 ] 

in which the rate is expressed in moles per liter per second when the con- 
centrations are in moles per liter. 

a. Calculate the rate of the reaction for 1 M acetone in water at pH 7 in 
the absence of acetic acid or acetate ion. 

b. Calculate the rate of the reaction for 1 M acetone in a solution made 
by neutralizing 1 M acetic acid with sufficient sodium hydroxide to 
give pH 5.0 (X HA of acetic acid = 1.75 x 10 - 5 ). 

12-27 The carbon skeletons of diphenylethanedione (benzil) and benzoin are 
identical, as are the skeletons of benzilate ion and diphenylhydroxyethanal. 
Yet, diphenylethanedione rearranges under the influence of base to give 
benzilate ion and the aldehyde rearranges under the influence of acid to 
give benzoin. Explain why these two reactions proceed in the directions that 
they do and provide reasonable mechanisms for both processes. 

O O 

C 6 H 5 -C-C-C 6 H 5 -531» (C 6 H 5 ) 2 COHC0 2 e 

(diphenylethanedione) (benzilate ion) 

O 

II H® 

C 6 H 5 -C-CHOHC 6 H 5 *- 2 — (C 6 H 5 ) 2 COHCHO 

(benzoin) (diphenylhydroxyethanal) 

12-28 Describe the course of the following reaction: 
O O O 

2 (C 6 H 5 ) 3 CCCH 3 < c ' H '>» ceNa ' t (C 6 H 5 ) 3 CCCH 2 CCH 3 + (C 6 H 5 ) 3 CH 

What would be a suitable solvent in which to conduct this reaction ? 



exercises 325 

12-29 The commercially important tetrahydroxy alcohol C(CH 2 OH) 4 known as 
pentaerythritol is prepared by alkaline addition of formaldehyde and acet- 
aldehyde according to the following equation: 



CH3CHO + 4CH 2 -^ L -» C(CH 2 OH) 4 + HC0 2 H 



Work out a sequence of steps which reasonably accounts for the course of the 
reaction. 



js. iX-v&i'A vi-i*i«^ 



chapter 13 

carboxylic acids and 

: ,,, i derivatives 



chap 13 carboxylic acids and derivatives 329 

We shall be concerned in this chapter with the chemistry of the carboxylic 
acids, RC0 2 H, and some of their functional derivatives of the type RCOX. 

P 

Although the carboxyl function — C is a combination of a hydroxyl 

O — H 

and a carbonyl group, the combination is such a close one that neither group 
behaves independently of the other. However, we shall be able to make a 
number of helpful comparisons of the behavior of the hydroxyl groups of 
alcohols and acids, and of the carbonyl groups of aldehydes, ketones, and 
acids. 

The carboxyl group is acidic because of its ability to donate a proton to a 
suitable base. In water most carboxylic acids are only slightly dissociated 
(K HA ~ 10" 5 , degree of ionization of a 1 M solution ~0.3%). 



RCO,H + H,0 



RCO, e + H,O a 



Aqueous solutions of the corresponding carboxylate salts are basic because of 
the reaction of the carboxylate anion with water (hydrolysis). 



CH 3 C0 2 e Na® + H 2 
sodium acetate 



CH 3 C0 2 H + Na® OH e 
acetic acid 



The nomenclature of carboxylic acids, which was discussed previously 
(Section 8-4), is illustrated with some representative compounds in Figure 13-1. 



Figure 13-1 Representative carboxylic acids. The IUPAC names are given 
first and some common names in parentheses. 



CH 3 CH 2 C0 2 H 

propanoic acid 
(propionic acid) 



CH 3 CH 2 CH 2 CH 2 CH 2 C0 2 H 



hexanoic acid 
(caproic acid) 



BrCH 2 C0 2 H 

bromoethanoic acid 
(bromoacetic acid) 



propenoic acid 
(acrylic acid) 



I CHC0 2 H 

cyclopropane- 

carboxylic acid 

(same) 



CH 3 CHC0 2 H 
I 
OH 



CHsCC0 2 H 

propynoic acid 
(propiolic acid) 



CO,H 



benzoic acid 
(same) 



NCCH 2 C0 2 H 



2-hydroxypropanoic acid cyanoethanoic acid 

(lactic or a-hydroxypropionic acid ) (cyanoacetic acid) 



CH 3 CHC0 2 H 

NH 2 

2-aminopropanoic acid 

(alanine or a-aminopropionic acid) 



O 

II 



butan-3-on-l-oic acid 
(acetoacetic acid) 



CH 2 C0 2 H 
CH 2 C0 2 H 

butanedioic acid 
(succinic acid) 



chap 13 carboxylic acids and derivatives 330 

Some common names are given in parentheses. The IUPAC names will be 
used wherever practicable in this chapter although the common names for the 
C x and C 2 compounds, formic acid and acetic acid, will be retained. 

Carboxylic acids with R as an alkyl or alkenyl group are also called fatty 
acids, but this term is more correct applied to the naturally occurring continuous 
chain, saturated and unsaturated aliphatic acids which, in the form of esters, 
are constituents of the fats, waxes, and oils of plants and animals. The most 
abundant of the fatty acids are palmitic, stearic, oleic, and linoleic acids ; they 
occur as glycerides, which are esters of the trihydroxy alcohol glycerol. 

CH 3 (CH 2 ) 14 C0 2 H palmitic acid 

CH 3 (CH 2 ) 16 C0 2 H stearic acid 

CH 3 (CH 2 ) 7 CH=CH(CH 2 ) 7 CO 2 H oleic acid (cis) 

CH 3 (CH 2 ) 4 CH=CHCH 2 CH=CH(CH 2 ) 7 C0 2 H linoleic acid 

Alkaline hydrolysis of fats affords salts of the fatty acids, those of the alkali 
metals being useful as soaps. The cleansing mechanism of soaps is described 
in the next section. 

o 

II 

CH 2 OCR 

I O CH 2 OH 

II NaOH i . ~ ffl H® ,„„„,. 



CHOCR > CHOH + 3RC0 2 a Na tt 

I 

O CH 2 OH 

II 
CH 2 OCR 

fat 
(a glyceride) glycerol soap fatty acid 



13-1 physical properties of carboxylic acids 

The physical properties of carboxylic acids reflect a considerable degree of 
association through hydrogen bonding. We have encountered such bonding 
previously in the case of alcohols (Section 10-1); however, acids form stronger 
hydrogen bonds than alcohols because their O— H bonds are more strongly 

8e Se 
polarized as — O — H. In addition, carboxylic acids have the possibility of 

forming hydrogen bonds to the rather negative oxygen of the carbonyl dipole 
rather than just to the oxygen of another hydroxyl group. Indeed, carboxylic 
acids in the solid and liquid states exist mostly as cyclic dimers. These dimeric 



Se 




O- 


■H — O 


i®// 


\« 


-c 


c 


\ 


// 


o- 


-H-0 




se 



structures persist in solution in hydrocarbon solvents and to some extent even 
in the vapor state. 

The physical properties of some representative carboxylic acids are listed in 



sec 13.1 physical properties of carboxylic acids 331 
Table 134 Physical properties of representative carboxylic acids 







solubility, 


mp, 


bp, 


JSTha(H 2 0) 


acid 


structure 


g/100gH 2 O 


°C 


°C 


at 25° 


formic 


HC0 2 H 


00 


8.4 


100.7 


1.77 x 10" 4 


acetic 


CH 3 C0 2 H 


00 


16.6 


118.1 


1.75 x 10- 5 


propanoic 


CH 3 CH 2 C0 2 H 


CO 


-22 


141.1 


1.3 x 10~ 5 


butanoic 


CH 3 CH 2 CH 2 C0 2 H 


GO 


-8 


163.5 


1.5 x 10- 5 


2-methylpropanoic 


(CH 3 ) 2 CHC0 2 H 


20 


-47 


154.5 


1.4 x 10- 5 


pentanoic 


CH 3 (CH 2 ) 3 C0 2 H 


3.3 


-34.5 


187 


1.6 x 10- 5 


palmitic 


CH 3 (CH 2 ) 14 C0 2 H 




64 


390 




stearic 


CH 3 (CH 2 ) 16 C0 2 H 




69.4 


360 d 




chloroacetic 


C1CH 2 C0 2 H 




63 


189 


1.4 x 10- 3 


dichloroacetic 


2 CHC0 2 H 


8.63 


5 


194 


5 x 10- 2 


trichloroacetic 


C1 3 CC0 2 H 


120 


58 


195.5 


3 x 10- 1 


trifluoroacetic 


F 3 CC0 2 H 


00 


-15 


72.4 


strong" 


2-chlorobutanoic 


CH 3 CH 2 CHC1C0 2 H 






1 Ai 1 5mm 


1.4 x 10~ 3 


3-chlorobutanoic 


CH 3 CHC1CH 2 C0 2 H 




44 


1 1/j22ram 


8.9 x 10- 5 


4-chlorobutanoic 


C1CH 2 CH 2 CH 2 C0 2 H 




16 


1Qg22mm 


3.0 x 10- 5 


5-chloropentanoic 


C1CH 2 (CH 2 ) 3 C0 2 H 




18 


130 llmm 


2 x 10" 5 


methoxyacetic 


CH 3 OCH 2 C0 2 H 






203 


3.3 x 10- 4 


cyanoacetic 


N=CCH 2 C0 2 H 




66 


l 08 15mm 


4 x 10~ 3 


vinylacetic 


CH 2 =CHCH 2 C0 2 H 




-39 


163 


3.8 xlO- 5 


benzoic 


C 6 H 5 C0 2 H 


0.27 


122 


249 


6.5 X 10" 5 


phenylacetic 


C 6 H 5 CH 2 C0 2 H 


1.66 


76.7 


265 


5.6 x 10" 5 



" The term "strong" acid means essentially complete dissociation in dilute aqueous solution; 
that is, the concentration of the neutral molecule is too low to be measured by any analytical 
technique now available. 



Table 13-1. The notably high melting and boiling points of acids relative to 
alcohols and chlorides can be attributed to the strength and degree of hydro- 
gen bonding. The differences in volatility are shown more strikingly by Figure 
13-2, which is a plot of boiling point versus n for the homologous series 



Figure 13-2 Boiling points of acids, CH 3 (CH 2 )„_ 2 C0 2 H, alcohols, 
CH 3 (CH 2 )„_ 2 CH 2 OH, and chlorides, CH 3 (CH 2 )„_ 2 CH 2 C1. 



200 










^^ ''.''' 




^^ ^ ,' 


160 


^^ alcohols^ - s 




jS^ s' *' 




^y^ ** '' 




j^ ^ ,* 


C 120 


^ ^ y 




^ 








^^^^^^^^^^^^^^^^^^^f^^m 


80 


^'chloric! os 




















:##itsfflst«^ 


40 


Wlmtm^^S^^X9^^^^^^^, 












1 1 . . 1 1 1 1 



chap 13 carboxylic acids and derivatives 332 

CH 3 (CH 2 )„_ 2 X, in which X is -C0 2 H, -CH 2 OH, and -CH 2 C1. 

Hydrogen bonding is also responsible for the high water solubility of the 
simple aliphatic acids — formic, acetic, propanoic, and butanoic — which are 
completely miscible with water in all proportions. As the alkyl chain increases 
in length (and in degree of branching) the solubility decreases markedly. On 
the other hand, the salts of carboxylic acids retain their moderately high solu- 
bilities in water even when the alkyl group becomes large. This enables car- 
boxylic acids to be extracted from solutions in benzene and other low- 
polarity solvents by aqueous base. Sodium bicarbonate is sufficiently basic 
to convert any carboxylic acid to the anion, RC0 2 H + HCOf -> RCOf 
+ H 2 + C0 2 , and is usually used for this purpose. Separation of the liquid 
layers, followed by acidification of the aqueous layer, then precipitates the free 
carboxylic acid. The separation of a mixture of a water-insoluble alcohol and 
water-insoluble carboxylic acid is illustrated in Scheme I. 



2 H 


1 . dissolve mixture in benzene 




2 OH 


2. extract benzene solution in 
separatory funnel with aqueous 
NaHCOj solution 






benzene layer 


aqueot 


is layer 

e Na 9 ) 








(RCO, 




distill off 
benzene 






H (b 




RCH 


2 OH 


RC( 


) 2 H 




SCHEME I. Thesef 


aration 


af a mix 


ure of 



a carboxylic acid. 

Though the salts of long-chain carboxylic acids are moderately soluble in 
water, the resulting solutions are usually opalescent as a result of the grouping 
together of molecules to form colloidal particles. These are called micelles, and 
their formation reflects the antipathy of the long hydrocarbon chains for the 
aqueous environment into which they have been drawn by the attraction of 
the water for the anionic group at the carboxylate end of the molecule. The 
alkyl groups cluster together in the micelle with the charged groups on the 
outside in position to be solvated by water in the usual way (Figure 13-3). 

Ordinary soaps are sodium or potassium salts of C 16 and C 18 acids, such as 
palmitic and stearic acids (Table 1 3-1), and their cleansing action results from 
the abilities of their micelles to dissolve grease and other nonpolar substances 
that are insoluble in water alone. 

In minute concentrations, salts such as sodium stearate, C 17 H 35 COf Na® 
(sodium octadecanoate), do not form micelles in water but, instead, concen- 
trate at the surface of the liquid with the charged ends of the salt molecules 
immersed in the water and the hydrocarbon parts forming a surface layer. 
This results in a sharp drop in surface tension of water by even minute con- 



sec. 13.1 physical properties of carboxylic acids 333 




Figure 13-3 A micelle, in which long-chain carboxylate salt molecules group 
together in aqueous solution. The interior of the micelle is a region of very 
low polarity. The total negative charge on the micelle is balanced by the 
positive charge of sodium ions in the solution. 



centrations of soap. As the concentration of long-chain carboxylate salt is 
increased and after the surface has become saturated, the system can either 
form micelles or attempt to increase its surface. Agitation will allow the latter 
to occur (frothing) but otherwise, at a certain concentration, micelles will 
begin to form. This point is called the critical micelle concentration. 

Micelles have some interesting catalytic properties. Not only do they pro- 
vide a nonpolar environment in an aqueous system, but they have a large 
charge concentrated at their surface. Each of these characteristics can be 
important in accelerating the rates of certain reactions, and the catalytic 
behavior of micelles is being actively investigated. 

The lower-molecular-weight aliphatic acids have quite characteristic 
odors. Although formic, acetic, and propanoic acids have sharp odors, those 
of the C 4 to C 8 acids (butanoic to octanoic) are disagreeable and can be 
detected in minute amounts, especially by dogs. 1 It has been shown that a dog's 
tracking ability stems from its recognition of the particular blend of com- 
pounds, mostly aliphatic acids, released by the sweat glands in the feet of the 
person being followed. Each person's metabolism produces a characteristic 



X A dog can detect butanoic acid at a concentration of 10~ 17 mole per liter of air, about a 
million times less than the concentration required by man (R. H. Wright, The Science of 
Smell, George Allen and Unwin, London, 1964). 



chap 13 carboxylic acids and derivatives 334 

spectrum of compounds although those from identical twins differ little from 
each other. The higher-molecular-weight carboxylic acids have low volatilities 
and hence are essentially odorless. 



13-2 spectra of carboxylic acids 

The infrared spectra of carboxylic acids provide clear evidence of hydrogen 
bonding. This is illustrated in Figure 13-4, which shows the spectrum of 
acetic acid in carbon tetrachloride solution, together with those of ethanol and 
acetaldehyde for comparison. The spectrum of ethanol has two absorption 
bands, characteristic of the OH bond; one is a sharp band at 3640 cm" 1 , 
corresponding to free or unassociated hydroxyl groups, and the other is a 
broad band centered on 3350 cm -1 due to hydrogen-bonded groups. The 
spectrum of acetic acid shows no absorption due to free hydroxyl groups but, 
like that of ethanol, has a broad intense absorption ascribed to associated OH 
groups. However, the frequency of absorption, 3000 cm -1 , is shifted appre- 
ciably from that of ethanol and reflects a stronger type of hydrogen bonding 
than in ethanol. The absorption due to the carbonyl group of acetic acid 
(1740 cm -1 ) is broad but not shifted significantly from the carbonyl absorp- 
tion in acetaldehyde. 

The carboxyl function does absorb ultraviolet radiation, but the wavelengths 
at which this occurs are appreciably shorter than for carbonyl compounds 
such as aldehydes and ketones, and, in fact, are barely in the range of most 
commercial ultraviolet spectrometers. Some idea of how the hydroxyl sub- 
stituent modifies the absorption properties of the carbonyl group in carboxylic 
acids can be seen from Table 13-2, in which are listed the wavelengths of 
maximum light absorption (l max ) and the extinction coefficients at maximum 
absorption (e max ) of several carboxylic acids, aldehydes, and ketones. 

In the nmr spectra of carboxylic acids, the carboxyl proton is found to 
absorb at unusually low magnetic fields. The chemical shift of carboxylic acid 
protons comes about 5.5 ppm toward lower magnetic fields than that of the 
hydroxyl proton of alcohols. This behavior parallels that of the enol hydro- 
gens of 1,3-dicarbonyl compounds (Section 12-6) and is probably similarly 
related to hydrogen-bond formation. 



Table 13'2 Ultraviolet absorption properties 
of carboxylic acids, aldehydes, and ketones 



compound 




^-max 


solvent 


acetic acid 


204 


40 


water 


acetic acid 


197 


60 


hexane 


acetaldehyde 


293 


12 


hexane 


acetone 


270 


16 


ethanol 


butanoic acid 


207 


74 


water 


butyraldehyde 


290 


18 


hexane 



sec 13.2 spectra of carboxylic acids 335 




Figure 13-4 Infrared spectra of ethanol (a), acetic acid (b), and acetaldehyde 
(c); 10% in carbon tetrachloride. 



chap 13 carboxylic acids and derivatives 336 

13-3 preparation of carboxylic acids 

The first three methods listed below have already been met in earlier 
chapters. 

1. Oxidation of a primary alcohol or aldehyde (Sections 10-6 and 11-4G); 

RCH,OH ► RCHO ► RCO,H 



2. Cleavageof alkenes or 1,2-glycols (Sections 4-4Gand 10-7). Any powerful 



RCH=CHR 



RCHOHCHOHR ► 2RC0 2 H 

oxidant may be used ; carboxylic acids are produced only if the carbons being 
cleaved possess hydrogen atoms — otherwise, ketones result. Cleavage of 
aryl side chains can also be brought about by drastic oxidation, ArCH 2 R -» 
ArC0 2 H (Section 24-1). 

3. Carbonation of Grignard reagents (Section 9-9C). This is a useful method 

O 

t H 2 

RMgX + C0 2 > R-C H e > RC0 2 H 

OMgX 

of extending a chain by one carbon atom. 

4. Hydrolysis of nitriles (to be described in Section 16-2B): 

RC=N W^ RCO > H 

5. Malonic ester synthesis (to be described in Section 13-9C): This is a 

RX + e CH(C0 2 Et) 2 ► RCH(C0 2 Et) 2 ► RCH(C0 2 H) 2 



RCH 2 C0 2 H 

useful method of extending a chain by two carbon atoms. 

O 

II 
Hydrolysis of esters (RC0 2 R), amides (RC— NH 2 ), and acid chlorides 

O 

II 
(RC— CI) also gives carboxylic acids, but these compounds are usually pre- 
pared from carboxylic acids in the first place. 



13-4 dissociation of carboxylic acids 

A. THE RESONANCE EFFECT 

Compared with mineral acids such as hydrochloric, perchloric, nitric, and 
sulfuric acids, the fatty acids, CH 3 (CH 2 )„_ 2 C0 2 H, are weak. The extent of 



sec 13.4 dissociation of carboxylic acids 337 

dissociation in aqueous solution is relatively small, the acidity constants, 
K HA , being approximately 10~ 5 (see Table 13-1). 

RC0 2 H + H 2 ■ ' RC0 2 e + H 3 O e 
[RC0 2 e ][H 3 O e ] 



[RC0 2 H] 



10" 5 for R = CH 3 (CH 2 )„ 



Even though they are weak, the fatty acids are many orders of magnitude 
stronger than the corresponding alcohols, CH 3 (CH 2 )„_ 2 CH 2 OH. Thus, 
the K HA of acetic acid, CH 3 C0 2 H, is 10 10 times larger than that of ethanol, 
CH 3 CH 2 OH. 

The acidity of the carboxyl group can be accounted for by resonance stabili- 
zation of the carboxylate anion, RCOf , which has the unit of negative charge 
distributed to both oxygen atoms (Section 6-4). 



R-C 



[la] 

The neutral carboxylic acid is expected to possess some stabilization asso- 
ciated with the ionic resonance structure [2b], but this is a minor contributor 
to the hybrid compared to [2a]. For the carboxylate anion, on the other hand, 
the two contributing forms [la] and [lb] are equivalent: 




O o e 

// / 

-C < > R-C 

\ \m 

OH OH 

[2a] [2b] 



P 
-C 
\ 
OH 



Alcohols are much weaker acids than are carboxylic acids because the 
charge in the alkoxide ion is localized on a single oxygen atom. 



B. THE INDUCTIVE EFFECT 

Although unsubstituted alkanoic acids with two carbons or more vary 
little in acid strength, substitution in the alkyl group can cause large acid- 
strengthening effects to appear (Table 13T). Formic acid and almost all the 
a-substituted acetic acids of Table 13-1 are stronger than acetic acid; trifluoro- 
acetic acid is in fact comparable in strength to hydrochloric acid. The nature of 
the groups which are close neighbors of the carboxyl carbon obviously has a 
profound effect on the acid strength, a phenomenon which is commonly 
called the inductive effect (symbolized as +/). The inductive effect is dis- 
tinguished from resonance effects of the type discussed earlier by being 
associated with substitution on the saturated carbon atoms of the fatty acid 
chain. It is taken as negative ( — /) if the substituent is acid enhancing, and 
positive ( + /) if the substituent is acid weakening. 



o 


arrows show 


// 
c 


movement of average 


position of electrons 


0<~H 


toward chlorine 




(-/effect) 



chap 13 carboxylic acids and derivatives 338 

The high acid strength of a-halogen-substituted acids (e.g., chloroacetic), 
compared with acetic acid, results from the electron-attracting power (elec- 
tronegativity) of the substituent halogen relative to the carbon to which it is 
attached. The electron-attracting power of three such halogen atoms is of 
course expected to be greater than that of one halogen; hence trichloroacetic 
acid (K HA , 3.0 x 1CT 1 ) is a markedly stronger acid than chloroacetic acid 
(^ HA ,1.4xlO- 3 ). , 



C1<-CH, 



As would be expected the inductive effect falls off rapidly with increasing 
distance of the substituent from the carboxyl group. This is readily seen by the 
significant difference between the K HA values of the 2-, 3-, and 4-chlorobuta- 
noic acids (see Table 13T). 

Many other groups besides halogen exhibit an acid-enhancing, electron- 
withdrawing (-/) effect. Among these are nitro (— N0 2 ); methoxyl 

(CH 3 0— ); carbonyl ( C = 0, as in aldehydes, ketones, acids, esters, and 

amides); cyano (—C=N); and trialkylammonio (R 3 N— ). Alkyl groups — 
methyl, ethyl, isopropyl, etc. — are the only substituents listed in Table 13T 
that are acid weakening relative to hydrogen (as can be seen by comparing 
their ^ HA 's with those of formic and acetic acids). This means that alkyl 
groups release electrons to the carboxyl group and thus exhibit a + / effect. 
The magnitude of the electrical effects of alkyl groups does not appear to 
change greatly in going from methyl to ethyl to propyl, and so on (compare 
the K HA values of acetic, propanoic, butanoic, and pentanoic acids). 



CH,-»C 



the arrows represent shifts 

/ in the average positions of the 

bonding electrons from the 
0->H methyl group toward the car- 
boxyl group ( + / effect) 



In addition to their acidic properties, carboxylic acids also can act as 
weak bases, the carbonyl oxygen accepting a proton from a strong acid such 
as H 2 S0 4 or HC10 4 (Equation 13T). Such protonation is an important step 
in acid-catalyzed esterification, as discussed in Section 10-4C. 

RCQ 2 H + H 2 S0 4 , RC0 2 H 2 ' e + HS0 4 e (13-1) 

It requires a 12.8 AT solution of sulfuric acid (74 % H 2 S0 4 , 26 % H 2 0) to 
" half-protonate " acetic acid. This means that if a small amount of acetic 
acid is dissolved in 12.8 M sulfuric acid, half of the acetic acid molecules at 
any instant would be in the cationic form, RC0 2 Hf , whereas in 1 M sulfuric 
acid, only about one in a million would be. (The acidity of solutions of strong 
acids rises very sharply as the medium becomes less aqueous.) 

The cation formed by protonation of acetic acid (" acetic acidium ion " 



sec 13.5 reactions at the carbonyl carbon of carboxy lie acids 339 

or "conjugate acid of acetic acid") has its positive charge distributed to both 
oxygen atoms and, to a much lesser extent, to carbon. 



CH 3 -C 



f H 



\ 
OH 



OH 



OH 



OH 



CH,-C« 



OH 



OH 



= ch,— c;< 



= CH,C0 2 H? 



OH 



Although oxygen is more electronegative than carbon, the resonance forms 
with oxygen bearing the positive charge are expected to be more important 
than the one with carbon bearing the positive charge because the former have 
one more covalent bond than the latter. 

Another cation can be reasonably formed by protonation of acetic acid, 



CH, — C 



O 

// 



OH, 



This ion is in rapid equilibrium with the isomer that has a proton on each 
oxygen, but is present in smaller amount. Note that these isomeric ions are 
tautomers (Section 12-6) and not resonance forms, because they are inter- 
converted only by changing the atomic positions. 



13-5 reactions at the carbonyl carbon 
of carboxy lie acids 

Many important reactions of carboxylic acids involve attack on carbon of 
the carbonyl group by nucleophilic species. These reactions are frequently 
catalyzed by acids, since addition of a proton or formation of a hydrogen 
bond to the carbonyl oxygen makes the carbonyl carbon more strongly 
electropositive and hence more vulnerable to nucleophilic attack. The follow- 
ing equations illustrate an acid-catalyzed reaction involving a negatively 
charged nucleophile (:Nu e ): 



R-C 



/° 



-OH 



+ H* 



OH 



R— C> 



:Nu e 



OH 



OH 
R— C— OH 

Nu 



Subsequent cleavage of a C— O bond and loss of a proton yields a displace- 
ment product : 



OH 
R-C-OH 

Nu 



-OH e 



R-C 

\ 



r 



,0 



R-C 



Nu 



Nu 



An important example of this type of reaction is the formation of esters as dis- 
cussed in Section 10-4C. Similar addition-elimination mechanisms occur in 



chap 13 carboxylic acids and derivatives 340 

many reactions at the carbonyl groups of acid derivatives. A less obvious 
example of addition to carboxyl groups involves hydride ion (H : e ) and takes 
place in lithium aluminum hydride reduction of carboxylic acids (Section 
13-5B). 



A. ACID-CHLORIDE FORMATION 

Carboxylic acids react with phosphorus trichloride, phosphorus pentachlo- 
ride, or thionyl chloride with replacement of OH by CI to form acid (acyl) 
chlorides, RCOC1. 

O O 

// sort v 

(CH 3 ),CHCH 2 C — ^^ (CH 3 ) 2 CHCH 2 C + S0 2 + HC1 

OH CI 

3-methylbutanoic acid 3-methylbutanoyl chloride 

(isovaleric acid) (isovaleryl chloride) 

Formyl chloride, HCOC1, is unstable and decomposes rapidly to carbon 
monoxide and hydrogen chloride at ordinary temperatures. 



B. REDUCTION OF CARBOXYLIC ACIDS 

In general, carboxylic acids are difficult to reduce either by catalytic hydro- 
genation or by sodium and alcohol. Reduction to primary alcohols proceeds 
smoothly, however, with lithium aluminum hydride, Li A1H 4 . 

RC q 2 h y™± ^^°* RCH 2 OH 

I i A 1 H H ® H O 

CH 2 =CHCH 2 C0 2 H *-> L - ± -* CH 2 =CHCH 2 CH 2 OH 

3-butenoic acid 3-buten-l-ol 

(vinylacetic acid) (allylcarbinol) 

The first step in lithium aluminum hydride reduction of carboxylic acids is 
formation of a complex aluminum salt of the acid and liberation of 1 mole of 
hydrogen : 

O O 

/ t 

R— C + LiAlH 4 ► R-C + H 2 

\ \ e ffi 

OH OAlHj Li 

Reduction then proceeds by successive transfers of hydride ion, H: e , from 
aluminum to carbon. Two such transfers are required to reduce the acid salt to 
the oxidation level of the alcohol : 

RC0 2 s + 2 H e > RCH 2 O e + 2e 

The anions shown as products in this equation are actually in the form of 
complex aluminum salts from which the product is freed in a final hydrolysis 
operation. The overall reaction can be shown as follows : 



4RC0 2 H + 3LiAlH 4 



sec 13.6 decarboxylation of carboxylic acids 341 

[(RCH 2 0) 4 Al]Li + 4H 2 + 2LiA10 2 
H 2 0, HC1 
4RCH,0H + A1C1, + LiCl 



13-6 decarboxylation of carboxylic acids 

The ease of loss of carbon dioxide from the carboxyl group varies greatly 
with the nature of the acid. Some acids require to be heated as their sodium 
salts in the presence of soda lime (in general, however, this is not a good 
preparative procedure). 



CH, 



Na a 



NaOH, CaO 



-> CH 4 



CO, 



sodium acetate 



Other acids lose carbon dioxide simply by being heated at moderate tempera- 
tures. 



/ 140M60 
CH 2 ► CH 3 C0 2 H + C0 2 

C0 2 H 

malonic acid acetic acid 



Thermal decarboxylation occurs most readily when the a carbon carries a 
strongly electron-attracting group (i.e., — / substituent), as in the following 
examples : 



NC-CH 2 -COOH 

CH 3 CO-CH 2 -COOH 

CCI,CO,H 



nitroacetic acid 
malonic acid 
cyanoacetic acid 
acetoacetic acid 
trichloracetic acid 



decarboxylation occurs 
readily at 100°-150° 



The mechanisms of thermal decarboxylation are probably not the same in 
all cases, but decarboxylation of acids having a /?-carbonyl group is prob- 
ably a cyclic process of elimination in which hydrogen bonding plays an 
important role : 



'i) tf 



CH 2 

malonic acid 



OH 
/ 
HO-C 
\ 
CH 2 

enol form of 
acetic acid 



O 

II 

c 

II 

O 



CHjCOOH + C0 2 



acetic acid 



chap 13 carboxylic acids and derivatives 342 

Stepwise decarboxylation also occurs, particularly in reactions in which the 
carboxylate radical (RC0 2 • ) is formed. This radical can decompose further to 
a hydrocarbon radical R- and C0 2 . The overall decarboxylation product is 
determined by what R- reacts with: If a good hydrogen atom donor is 
present, RH is formed; if a halogen donor such as Br 2 is present, RBr is 
formed. 

RC0 2 - ► R-+C0 2 

R- + RH > RH + R'. 

R- + Br 2 > RBr + Br- 

Carboxylate radicals can be generated several ways. One is the thermal 
decomposition of diacyl peroxides, which are compounds with a rather weak 
O— O bond: 

O O O 

II i II II 

R— C— O+O-C- R > 2R— C— O 

Another method involves electrolysis of sodium or potassium carboxylate 
solutions, known as Kolbe electrolysis, in which carboxylate radicals are 
formed by transfer of an electron from the carboxylate ion to the anode. 
Decarboxylation may occur simultaneously with, or subsequent to, the forma- 
tion of carboxylate radicals, leading to hydrocarbon radicals, which sub- 
sequently dimerize. 

RC0 2 e ► RC0 2 -+e anode reaction 

K.® + e + H 2 > KOH + i H 2 cathode reaction 

RC0 2 - > R-+C0 2 

R-+R- > RR 

In the Hunsdiecker reaction, an alkyl bromide is formed when a silver salt 
of a carboxylic acid is treated with bromine in the absence of water. Carboxyl- 
ate radicals are probably involved. This reaction provides a means for remov- 
ing a carbon from the end of a chain with retention of a functional group in 
the product. 



-AgBr 




RBr + CO, 



13-7 reactions at the 2 position of carboxylic acids 



HALOGENATION 



Bromine reacts smoothly with carboxylic acids in the presence of small 
quantities of phosphorus to form 2-bromo acids. The reaction is slow in the 



RCH,C0 2 H + Br 2 — ^— > RCHBrC0 2 H + HBr 



sec 13.7 reactions at the 2 position of carboxylic acids 343 

absence of phosphorus, whose function appears to be to form phosphorus 



O 



// 
tribromide which reacts, with the acid to give the acid bromide, — C a 

Br 

compound known to be substituted readily by bromine. Substitution occurs 
exclusively at the 2 position (a position) and is therefore limited to carboxylic 
acids with a hydrogens. Chlorine with a trace of phosphorus reacts similarly 
but with less overall specificity. Concurrent radical chlorination can 
occur at all positions along the chain (as in hydrocarbon halogenation; see 
Section 3-3B). 



CH 3 CH 2 C0 2 H , 



ci 2 , p 



> CH 3 CHC0 2 H 

CI 



CI 2 , ku 



■> CH 3 CHC0 2 H + CICH 2 CH 2 C0 2 H 

CI 

2-chIoropropanoic 3-chloropropanoic 
acid acid 



B. SUBSTITUTION REACTIONS OF 2-HALO ACIDS 

The halogen of a 2-halo acid is activated by the adjacent electron-withdraw- 
ing carboxyl group and is readily replaced by nucleophilic reagents such as 
CN e , OH e , I e , and NH 3 . Thus, a variety of 2-substituted carboxylic acids 
may be prepared by reactions that are analogous to S N 2 substitutions of 
alkyl halides (Scheme II). 



CH 3 CHC0 2 H 

I 
I 

2-iodopropanoic 
acid 





CH 3 CHC0 2 e 
I 
OH 



-> CH 3 CHC0 2 H 



OH 



OH 

lactic acid 



CH 3 CHCO,H eXCeSS > CH,CHC0,NH 4 HS > CH 3 CHC0 2 H 



I 
Br 

2-bromopropanoic acid 

CN e 



CH 3 CHC0 2 H 

I 

CN 

2-cyanopropanoic acid 



H 2 



NH, 



* CH 3 CHCO ? H 

C0 2 H 

methylmalonic acid 



NH 2 
alanine 



SCHEME II 



chap 13 carboxylic acids and derivatives 344 

functional derivatives of carboxylic acids 

A functional derivative of a carboxylic acid is a substance formed by 
replacement of the hydroxyl group of the acid by some other group, X, that 
can be hydrolyzed back to the parent acid according toEquation 13-2. By this 
definition, an amide, RCONH 2 , but not a ketone, RCOCH 3 , is a functional 

o o 

R— C + H 2 ► R-C + HX (13-2) 

X OH 

derivative of a carboxylic acid. A number of types of acid derivatives are given 
in Table 13-3. 

The common structural feature of the compounds listed in Table 13-3 is the 

P 

acyl group R— C . Nitriles, RC=N, however, are often considered to be 

acid derivatives, even though the acyl group is not present as such, because 
hydrolysis of nitriles leads to carboxylic acids. The chemistry of nitriles is dis- 
cussed in Chapter 16. 

CH 3 C=N " ' H2 °> CH3COOH 

acetonitrile acetic acid 

The carbonyl group plays a dominant role in the reactions of acid deriva- 
tives, just as it does for the parent acids. The two main types of reactions of 
acid derivatives with which we shall be concerned are the replacement of X by 
attack of a nucleophile : Nu e at the carbonyl carbon with subsequent cleavage 
of the C— X bond (Equation 13-3), and substitution at the 2-carbon facilita- 
ted by the carbonyl group (Equation 13-4). 

O 9 o 

R _ c > + :N e > R _^/x ► R-C + :X e (13-3) 

\ I \ 

X Nu Nu 



O O 

RCH 2 -C + YZ ► RCH-C + HZ (13-4) 

\ I \ 

X Y * 



13- 8 displacement reactions of acid derivatives 

The following are the more important displacement reactions : 

1 . Acid derivatives are hydrolyzed to the parent acids. These reactions are 
commonly acid and base catalyzed, but acid chlorides usually hydrolyze 
rapidly without the agency of an acid or base catalyst : 



sec 13.8 displacement reactions of acid derivatives 345 

O P 

// H e or OH e " 

R-C + H 2 ° » R-C^ + HX 
\ OH 

X = -OR (ester), halogen (acid halide), -NH 2 (amide), and -0 2 CR (acid anhydride) 

2. Acid or base catalysts are usually required for ester interchange. 

o p 

/ H® or 9 OR / 

CH 3 -C + CH 3 CH;,OH .... CH 3 -C + CH3OH 

OCH 3 OCH 2 CH 3 

methyl acetate ' ethanol ethyl acetate methanol 

3. Esters are formed from acid chlorides and anhydrides. 



O 

// 



o 



R-C + R'OH — > R-C + HC1 

CI OR' 



O 



R-C 

\ 



O 



o 



R-C. 



O + R'OH 



-► R-C + R-C 

W OH 



O 



4. Amides are formed from esters, acid chlorides, and anhydrides. 



R-C 



? 



OR' 



// 



R— C 



\ 
CI 



o 



R-C 



R-C 



O 



NH 3 



R-C 



+ R'OH 



NH 2 



R-C 



R-C 

\ 



NH, 



P 



NH, 



e 



NH, ffi Cl 



RCO, s NHi 



All of these reactions are rather closely related, and we shall illustrate the 
principles involved mostly by the reactions of esters, since these have been 
particularly well studied. Acid-catalyzed hydrolysis of esters is the reverse of 
acid-catalyzed esterification discussed previously (Section 10-4C). In contrast, 
base-induced hydrolysis (saponification) is, in effect, an irreversible reaction. 
The initial step is the attack of hydroxide ion at the electron-deficient carbonyl 
carbon; the intermediate anion [3] so formed then has the choice of losing 
OH e and reverting to the original ester, or of losing CH 3 O e to form the 



chap 13 carboxylic acids and derivatives 346 
Table 13*3 Functional derivatives of carboxylic acids 



derivative 


structure 


exa 


mple 






structure 


name 


esters 


' P 

R-C 

OR 


P 
CH 3 -C x 

OC 2 H 5 


ethyl 
acetate 


acid halides 


P 
R-C x 


of 


benzoyl 


(acyl halides) 


X 

X = F, CI, Br, I 


^- / Br 


bromide 




P 
R-C 


P 
CH 3 -C 




anhydrides 




\ 
O 

CH 3 -< 
O 


acetic 
anhydride 


amides 
(primary) 


P 
R-C x 

NH 2 


NH 2 


benzamide 


amides 


O 

II 


O 

// 


N-methyl- 


(secondary) 


RCNHR' 


CH 3 -C x 

NHCH 3 

O 


acetamide 


amides 


O 

II 


// 
H — C 


N,N-dimethyl- 


(tertiary) 


RCNR'R' 


N(CH 3 ) 2 


formamide 



sec 13.8 displacement reactions of acid derivatives 347 
Table 13-3 Functional derivatives of carboxylic acids (continued) 



derivative 



structure 



example 



structure 



itnides 



acyl azides 



hydrazides 



hydroxamic 
acids 



O 



R-C 



NH 



R-C. 





II 

H 2 C-- C / 



NH 



succinimide 



O 



P 



R— C 



P 



N, 



acetyl 
azide 



P 



R— C 

\ 



C,H,— C 



NHNH, 



O 

propanoyl 
\ hydrazide 

NHNH, 



P 



p 



R-C. 



\ 
NHPH 



C1CH,C 



\ 
NHOH 



chloroacetyl- 
hydroxamic 
acid 



lactones 
(cyclic esters) 


O 

II 

(CH 2 )„| 

most stable 
with n — 3, 4 


O 
II 
H 2 C^ C \ 

1 o 

H * C ^CH 2 


y-butyrolactone 


lactams 
(cyclic amides) 


P 

II 

^"l 
(CH 2 )„ 1 

V_NH 
most stable 


O 

II 
H 2/ C-C x 

HaC^ NH 

H 2 C — CH 

1 


S-caprolactam 




with n — 3, 4 


CH 3 





chap 13 carboxylic acids and derivatives 348 

acid. The overall reaction is irreversible since, once the acid is formed, it is 
immediately converted to the carboxylate anion, which is not further attacked 



o 



f OCH 3 



o e 

I 

CH 3 -C-OCH ; 
OH 

[3] 



by base. As a result, the reaction goes to completion in the direction of 
hydrolysis. 



CH 3 — C— OCHj 



OH 

[3] 



CH 3 -C + CH 3 O e 

\ 
OH 



-» CHjCOf + HOCH3 



Base-catalyzed ester interchange is analogous to the saponification reaction, 
except that an alkoxide base is used in catalytic amounts in place of hydroxide. 
The equilibrium constant is much nearer to unity, however, than for saponi- 
fication, because the salt of the acid is not formed. 



o 



,0 



CH 3 -C 



\ 
OCH, 



RO e II 

+ CH 3 CH 2 OH , CH 3 C 



+ CH3OH 



OCH,CH 3 



The mechanism is as shown in Equation 13-5. Either methoxide or ethoxide 



O 
II 
CH 3 -C 

\)CH 3 

methyl acetate 



O a 
I 
CH 3 -C-OCH 2 CH 3 

OCHj 



O 
II 
CH 3 -C 

3 \ 



+ CH,O e 



OCH 2 CH 3 

ethyl acetate 



(13-5) 



ion can be used as the catalyst since the equilibrium of Equation 13-6 is 
rapidly established. 



CH 3 O e + CH 3 CH 2 OH 



CH 3 CH 2 O e + CH3OH 



(13-6) 



Acid-catalyzed ester interchange is entirely analogous to acid-catalyzed 
esterification and hydrolysis and requires no further discussion. 



sec 13.8 displacement reactions of acid derivatives 349 

The reactions of a number of carboxylic-acid derivatives with organo- 
magnesium and organolithium compounds were described in Chapter 9 
(Section 9-9C). 

Esters, acid chlorides, and anhydrides are reduced by lithium aluminum 
hydride in the same general way as described for the parent acids (Section 
13-5B), the difference being that no hydrogen is evolved. The products are 
primary alcohols. 



o 



R-C 



1. LiAIH„ 



\ 2. H®, H 2 

z 



* RCH 2 OH Z = CI, OR, RC0 2 



Nitriles can be reduced to amines by lithium aluminum hydride. An imine 
salt is an intermediate product; if the reaction is carried out under the proper 
conditions, this salt is the major product and provides an aldehyde on 
hydrolysis. 



R _ C = N M^Ei, R-CH=N e Li ffi 
(imine salt) 



H 2 



LiAlH* H®, H 2 



RC 



Amides can be reduced to primary amines, and N-substituted amides to 
secondary and tertiary amines. 



o 

II 

RC— NH 2 

O 

RC— NHR' 

O 
II 
RC— NR,' 



LiAIH 4 | 
H®, H 2 



RCH 2 NH 2 

RCH 2 NHR 

RCH 2 NR 2 ' 



Although lithium aluminum hydride is a very useful reagent, it is sometimes 
too expensive to be used on a large scale. Other methods of reduction may then 
be necessary. Of these, the most important are reduction of esters with sodium 
and ethanol (acids do not reduce readily) and high-pressure hydrogenation 
over a copper chromite catalyst. 

C 2 H 5 OH 



RCO,R' + 4Na + 4C 2 H 5 OH 



-> RCH 2 OH + R'OH + 4C 2 H 5 O e Na s 



RC0 2 R' + 2H 2 



200° 



Cu(Cr) 



♦ RCH,OH + R'OH 



chap 13 carboxylic acids and derivatives 350 

13-9 reactions at the 2 position (a position) 
of carboxylic acid derivatives 

A. THE ACIDIC PROPERTIES OF ESTERS WITH a HYDROGENS 

Many important synthetic reactions in which C— C bonds are formed 
involve esters and are brought about by basic reagents. This is possible because 
the a hydrogens of an ester such as RCH 2 C0 2 C 2 H 5 are weakly acidic, and a 
strong base, such as sodium ethoxide, can produce a significant concentration 
of the ester anion at equilibrium. 



i 2 <~w 2 >~2ii5 -r L 2 n s i 



RCHCO,C 2 H, + C 2 H s OH 



The acidity of a hydrogens is attributed partly to the — / inductive effects of 
the ester oxygens, and partly to resonance stabilization of the resulting anion. 



O 



e // 

RCH-C 

\ 



OC,H 5 



/ 
-♦ RCH=C 

\ 



When the 2 position of the ester carries a second strongly electron-attracting 



Figure 13*5 Nuclear magnetic resonance spectrum of ethyl acetoacetate at 
60 MHz; calibrations are relative to tetramethylsilane at 0.00 ppm. Peaks 
marked a, b, and c, are assigned respectively to the protons of the enol form, 
whereas peaks d and e are assigned to the a-CH 2 and methyl protons, respec- 
tively, of the keto form. The quartet of lines at 4.2 ppm and the triplet at 1.3 
ppm result from the ethyl groups of both keto and enol forms. 



600 



400 



200 



11/ 



12.0 



10.0 



8.0 



6.0 



III 



4.0 



2~0 



\ f 

piliiiiiliigiiil) 
ppm 



sec 13.9 reactions at the 2 position of carboxylic acid derivatives 351 

group, the acidity of an a hydrogen is greatly enhanced. Examples of such 
compounds follow: 



C 2 H 5 2 CCH 2 C0 2 C 2 H 5 

NCCH 2 C0 2 C 2 H 5 

CH 3 COCH 2 C0 2 C 2 H s 



ethyl nitroacetate 
diethyl malonate 
ethyl cyanoacetate 
ethyl acetoacetate 



The stabilization of the anions of these specially activated esters is greater 
than for simple esters because the negative charge can be distributed over 
more than two centers. Thus, for the anion of ethyl acetoacetate, we can 
regard all three of the resonance structures [4a] through [4c] as important con- 

O o 

ii r 

CH 3 -C-CH 2 -C 

\ 
OC,H, 



-H® 



o 

II e // 

CHj-C-CH-C 



[4a] 



P 



O e o 

I / 

CH 3 -C=CH-C 

\ 



OC 2 H 5 



[4b] 
or 



O 



O 



CH 3 -C=-CH--C 



[4] 



O 



/ 



<— ► CH,-C— CH=C 



OC,H, 



[4c] 



tributors to the hybrid [4]. Since the anion [4] is relatively stable, the K nh of 
ethyl acetoacetate is about 10 _u in water solution. Although this compound 
is about 10 5 times as strong an acid as ethanol it is much more sluggish in its 
reaction with bases. Removal of a proton from carbon is a process with a 
finite energy of activation and only a small fraction of the collisions of such a 
molecule with hydroxide ion result in proton transfer. On the other hand, 
acids that have their ionizable protons attached to oxygen (even feeble acids 
such as ethanol) transfer their protons to strong bases on almost every colli- 
sion. 

Ethyl acetoacetate, like 2,4-pentanedione (Section 12-6), ordinarily exists 
at room temperature as an equilibrium mixture of keto and enol tautomers in 
the ratio of 92.5 to 7.5. This can be shown by rapid titration with bromine but 
is more clearly evident from the nmr spectrum (Figure 13-5), which shows 



chap 13 carboxylic acids and derivatives 352 

absorptions of the hydroxyl, vinyl, and methyl protons of the enol form, in 
addition to the absorptions expected for the keto form. 

o o o o 

II II I II 

CH 3 -C. .C-OC 2 H 5 ' -^ CH 3 -C^ ,C-OC 2 H 5 

CH 2 CH 

keto form, 92.5 % enol form, 7.5 % 

Interconversion of the enol and keto forms of ethyl acetoacetate is power- 
fully catalyzed by bases through the anion [4] and less so by acids through the 
conjugate acid of the keto form with a proton adding to the ketone oxygen. 




If contact with acidic and basic substances is rigidly excluded (to the extent 
of using quartz equipment in place of glass, which normally has a slightly 
alkaline surface), then interconversion is slow enough to enable separating 
the lower-boiling enol from the keto form by fractional distillation under 
reduced pressure. The separated tautomers are indefinitely stable when stored 
at —80° in quartz vessels. 



B. THE CLAISEN CONDENSATION 

One of the most useful of the base-induced reactions of esters is illustrated 
by the self-condensation of ethyl acetate under the influence of sodium ethox- 
ide to give ethyl acetoacetate. 

o 

CH 3 cioC 2 H 5 + H-fCH 2 C0 2 C 2 H 5 Na ° C2H5 » CH 3 COCH 2 C0 2 C 2 H 5 + C 2 H 5 OH 



This reaction is called the Claisen condensation and its mechanism has some 
of the flavor of both the aldol addition (Section 12-2A) and the nucleophilic 
reactions of acid derivatives discussed earlier (Section 13-5). The first step, 
as shown in Equation 13-7, is the formation of the anion of ethyl acetate, 
which, being a powerful nucleophile, attacks the carbonyl carbon of a second 
ester molecule (Equation 13-8). Elimination of ethoxide ion then leads to the 
/?-keto ester, ethyl acetoacetate (Equation 13-9). 



sec 13.9 reactions at the 2 position of carboxylic acid derivatives 353 

C 2 H 5 O e + CH 3 C0 2 C 2 H 5 . e CH 2 C0 2 C 2 H 5 + C 2 H 5 OH (137) 

I 
- CH 3 -C-CH 2 C0 2 C 2 H 5 

OC,H, 




(13-8) 



o 

II 

CH 3 -C— CH 2 C0 2 C 2 H 5 + C 2 H 5 O e 



(13-9) 



The sum of these steps represents an unfavorable equilibrium, and satisfactory 
yields of the /?-keto ester are obtained only if the equilibrium can be shifted by 
removal of one of the products. One simple way of doing this is to remove the 
ethyl alcohol by distillation as it is formed; this may be difficult, however, to 
carry to completion and, in any case, is self-defeating if the starting ester is 
low boiling. Alternatively, one can use a large excess of sodium ethoxide. 
This is helpful because ethanol is a weaker acid than the /?-keto ester, and 
excess ethoxide shifts the equilibrium to the right through conversion of the 
ester to the enolate salt. 

O O 

II II e 

CH 3 -C-CH 2 O0 2 C 2 H 5 + C 2 H 5 O e , ' CH 3 -C-CHC0 2 C 2 H 5 + C 2 H 5 OH 

Obviously the condensation product must be recovered from the enol salt 
and isolated under conditions that avoid reversion to starting materials. The 
best procedure is to quench the reaction mixture by pouring it into an excess 
of cold, dilute acid. 

Claisen condensations can be carried out between two different esters but, 
since there are four possible products, serious mixtures often result. This 
objection is obviated if one of the esters has no a hydrogen and reacts readily 
with a carbanion according to Equations 13-8 and 13-9. The reaction then has 
considerable resemblance to the mixed aldol additions, discussed in Section 
12-2A. Among the useful esters without a hydrogens and with the requisite 
electrophilic reactivity are those of benzoic, formic, oxalic, and carbonic 
acids. Two practical examples of mixed Claisen condensations are shown. 

C 6 H 5 C0 2 C 2 H 5 + CH 3 C0 2 C 2 H 5 '' ^ > C 6 H 5 COCH 2 C0 2 C 2 H 5 + C 2 H 5 OH 
ethyl benzoate ethyl benzoylacetate 



55% 



HC0 2 C 2 H 5 + C 6 H 5 CH 2 C0 2 C 2 H 5 

ethyl ethyl 

formate phenylacetate 



1. C 2 H 5 O e 



2. H® 



CHO 

ethyl formylphenylacetate 
90% 



II 

CHOH 



chap 13 carboxylic acids and derivatives 354 

An important variation on the Claisen condensation is to use a ketone as 
the anionic reagent. This often works well because ketones are usually more 
acidic than simple esters and the base-induced self-condensation of ketones 
(aldol addition) is thermodynamically unfavorable (Section 12-2A). A typical 
example is the condensation of cylclohexanone with ethyl oxalate. 

o o 

6JI COC0 2 C 2 H 5 

ethyl oxalate cyclohexanone 2-(ethyl oxalyl)- 

cyclohexanone 



C. ALKYLATION OF ACETOACETIC AND MALONIC ESTERS 

Alkylation of the anions of esters such as ethyl acetoacetate and diethyl 
malonate is a useful way of synthesizing carboxylic acids and ketones. The 
ester is converted by a strong base to the enolate anion (Equation 13-10), 
and this is then alkylated by an S N 2 attack on the alkyl halide (Equation 
13-11). Usually, C-alkylation predominates. 

o o 

II II e (13-10) 

t 2 C0 2 C 2 H 5 + C 2 H 5 O e , CH 3 C-CHC0 2 C 2 H 5 + C 2 H 5 OH v ' 



O O 

CH 3 I + CH 3 C-CHC0 2 C 2 H 5 ► CH 3 C-CHC0 2 C 2 H 5 + l e (13-11) 

CH 3 
Esters of malonic acid can be alkylated similarly. 

C0 2 C 2 H 5 C0 2 C 2 H 5 

/ NaOQH^ CH3CH.Br, CH / 

2 \ \ 

C0 2 C 2 H 5 C0 2 C 2 H 5 

Alkylacetoacetic and alkylmalonic esters can be hydrolyzed under acidic 
conditions to the corresponding acids and, when these are heated, they 
readily decarboxylate (see Section 13-6). Alkylacetoacetic esters thus yield 
methyl alkyl ketones, while alkylmalonic esters produce carboxylic acids. 

O CH 3 O CH 3 O 

B I H®. HjO II I heat « 

CH 3 C-CHC0 2 C 2 H 5 ILiUZU CH 3 C-CHC0 2 H J^ > CH 3 C-CH 2 CH 3 

methyl ethyl ketone 

C0 2 C 2 H 5 ^ C0 2 H 

CH 3 CH 2 CH H<B ' H2 °» CH 3 CH 2 CH ^^ CH 3 CH 2 CH 2 C0 2 H 

C0 2 C 2 H 5 C0 2 H 

butanoic acid 



sec 13.10 reactions of unsaturated carboxylic acids 355 

13-10 reactions of unsaturated carboxylic acids 
and their derivatives 

Unsaturated carboxylic acids of the type RCH=CH(CH 2 )„COOH usually 
exhibit the properties characteristic of isolated double bonds and isolated 
carboxyl groups when n is large and the functional groups are far apart. As 
expected, exceptional behavior is most commonly found when the groups are 
sufficiently close together to interact strongly, as in 2-alkenoic acids. These 

compounds are invariably called a,/?-unsaturated acids, RCH=CHC0 2 H, 
and we shall use this term herein. 



A. HYDRATION AND HYDROGEN BROMIDE ADDITION 

Like alkenes, the double bonds of a,/?-unsaturated acids can be brominated, 
hydroxylated, hydrated, and hydrobrominated, although the reactions are 
often relatively slow. With unsymmetrical addends, the direction of addition 
is opposite to that observed for alkenes (anti-Markownikoff). Thus pro- 
penoic acid (acrylic acid) adds hydrogen bromide and water so that 3-bromo- 
and 3-hydroxypropanoic acids are formed. These additions are closely anal- 
ogous to the addition of halogen acids to propenal (Section 12-3). 



HBr 

CH 2 =CHCOOH 
propenoic acid n. jj 2 q 



BrCH 2 CH 2 COOH 

3-broraopropanoic acid 



(acrylic acid) h® 



OH 

3-hydroxypropanoic acid 



B. LACTONE FORMATION 

When the double bond of an unsaturated acid lies farther down the carbon 
chain than between the a and /? positions, conjugate addition is not possible. 
Nonetheless, the double bond and carboxyl group frequently interact in the 
presence of acid catalysts because the carbonium ion that results from addition 
of a proton to the double bond has a built-in nucleophile (the carboxyl group), 
which may attack the cationic center to form a cyclic ester (i.e., a lactone). 
Lactone formation usually occurs readily by this mechanism only when a 
five- or six-membered ring can be formed. 

CH 3 
O ,o \ 

// h® ® t -H® HC-CH 2 

CH 2 =CHCH 2 CH 2 C — S — ► CH 3 -CHCH 2 CH 2 C — ^ / \ 

X OH OH \/ C " 2 

II 
O 

4-pentenoic acid (y-valerolactone) 



(allylacetic acid) 



chap 13 carboxylic acids and derivatives 356 

Five- and six-membered lactones are also formed by internal esterification 
when either 4- or 5-hydroxy acids are heated. Under similar conditions, 
3-hydroxy acids are dehydrated to a,/?-unsaturated acids, while 2-hydroxy 
acids undergo bimolecular esterification to substances with six-membered 
dilactone rings called lactides. 



H0CH 2 CH 2 CH 2 C0 2 H > O CH 2 + H 2 

C 

II 

o 

4-hydroxybutanoic acid y-butyrolactone 

(y-hydroxybutyric acid) 

CH 3 CHCH 2 C0 2 H heat > CH 3 CH=CHC0 2 H + H 2 
OH 

3-hydroxybutanoic acid 2-butenoic acid 

A J° 

heat CH 3 HC C 

2CH 3 CHC0 2 H ► I I + 2H 2 

I ,^-L ^CHCH 3 

oh or o^ 

2-hydroxypropanoic acid lactide 

(lactic acid) 



13-11 dicarb oxyli c aci ds 



Acids in which there are two carboxyl groups separated by a chain of more 
than five carbon atoms (n > 5) have, for the most part, unexceptional proper- 
ties, the carboxyl groups behaving more or less independently of one another. 

C0 2 H 
(CH 2 ), 
C0 2 H 

When the carboxyl groups are closer together, however, the possibilities for 
interaction increase; we shall be primarily concerned with such acids. A 
number of important dicarboxylic acids are listed in Table 13-4. 



A. ACIDIC PROPERTIES OF DICARBOXYLIC ACIDS 

The inductive effect of one carboxyl group is expected to enhance the acidity 
of the other and, from Table 13-4, we see that the acid strength of the dicar- 
boxylic acids, as measured by the first acid-dissociation constant, K u is 
higher than that of acetic acid (K HA =1.8 x 10~ 5 ) and falls off with increasing 
distance between the two carboxyl groups. Two other factors operate to 
raise K x in comparison to the K HA of acetic acid. First, the statistical factor: 



sec 13.11 dicarboxylic acids 357 



Table 13*4 Dicarboxylic acids 



acid 


formula 


mp, 

°C 


K ± x 10 5 

at 25° 


K 2 x 10 5 

at 25° 


oxalic 


C0 2 H 


189 


3500 


5.3 


(ethanedioic) 


C0 2 H 








malonic 


p0 2 H 


136 


171 


0.22 


(propanedioic) 


CH 2 
C0 2 H 

C0 2 H 


dec. 






succinic 


(C X H 2 ) 2 
CO,H 


185 


6.6 


0.25 


(butanedioic) 








glutaric 
(pentanedioic) 


C0 2 H 

/ 
(C^ 2 ) 3 
CO,H 

7 C0 2 H 
(CH 2 ) 4 
C0 2 H 


98 


4.7 


0.29 


adipic 
(hexanedioic) 


152 


3.7 


0.24 




7 C0 2 H 
(CH 2 ) 5 








pimelic 


105 


3.4 


0.26 


(heptanedioic) 


C0 2 H 








maleic (cis- 


HCC0 2 H 

II 


130 


1170 


0.026 


butenedioic) 


HCC0 2 H 








fumaric (trans- 


HCC0 2 H 

II 


sub. 


93 


2.9 


butenedioic) 


H0 2 CCH 


200 






phthalic (benzene- 


|f^C0 2 H 


231 


130 


0.39 18 


1 ,2-dicarboxylic) 


l^C0 2 H 









there are two carboxyl groups per molecule instead of one. Second, there can 
be stabilization of the monoanion by internal hydrogen bonding, when the 
geometry of the molecule allows it. 




chap 13 carboxylic acids and derivatives 358 

The second acid-dissociation constant, K 2 , is smaller than K HA for acetic 
acid in most cases, and this must also be largely due to the hydrogen-bonding 
effect. Comparison of K 1 and K 2 for maleic and fumaric acids is especially 
instructive. Only the cis monoanion can be stabilized by internal hydrogen 
bonding and we find K± to be larger and K 2 smaller for the cis acid. 



B. THERMAL BEHAVIOR OF DICARBOXYLIC ACIDS 

The reactions that occur when diacids are heated depend critically upon the 
chain length separating the carboxyl groups. Cyclization is usually favored if a 
strainless five- or six-membered ring can be formed. Thus adipic and pimelic 
acids cyclize and decarboxylate to give cyclopentanone and cyclohexanone, 
respectively. 



COOH 

(CH 2 ) 4 - 
COOH 

adipic acid 



300° 



H,C 



XH 2 
\ _ 

~CH, 



O + CO, 



cyclopentanone 



COOH 
(CH 2 ) 5 
COOH 

pimelic acid 



300° 



H,C— CH, 
/ \ 
H 2 C C=0 + C0 2 + H 2 

H 2 C— CH 2 

cyclohexanone 



Succinic and glutaric acids take a different course. Rather than form the 
strained cyclic ketones — cyclopropanone and cyclobutanone — both acids 
form cyclic anhydrides — succinic and glutaric anhydrides — having five- and 
six-membered rings, respectively. Pbthalic and maleic acids behave similarly 
giving five-membered cyclic anhydrides. 



COOH 

(CH 2 ) 2 
\ 
COOH 



acid 



COOH 
(CH 2 ) 3 
COOH 



;lutanc 
acid 



300° 



-H,0 



300° 
-H,0 



P 



\ 



H 2 C 
H 2 C^ C / 



O 



o 



anhydride 

O 

H 2 C O 

\ _/ 



O 



glutaric 
anhydride 




CO,H 



CO,H 



230° 



-H 2 



phthalic 
acid 



^^co 



phthalic 
anhydride 

P 



O 



HCC0 2 H 200 Q ) HC' 
HCC0 2 H - H ^ 0> HC-./" 
O 



maleic 
acid 



maleic 
anhydride 



summary 359 

Malonic and oxalic acids behave still differently, each undergoing decar- 
boxylation when heated (Section 13-6). 

COOH 

/ 140°-160° 
CH 2 ► CH3COOH + C0 2 

COOH 

malonic acid 

COOH 



COOH 

oxalic acid 



160M80 °. CO, + HCOOH 



summary 

Carboxylic acids, such as (CH 3 ) 2 CHC0 2 H (2-methylpropanoic acid or 
isobutyric acid), have higher melting and boiling points than alcohols of the 
same molecular weight. They ionize weakly in aqueous solution (K HX ~ 10" 5 ) 
but associate as dimers in hydrocarbon solvents. 

Carboxylate salts (RCOf Na®) are quite soluble in water and this fact 
permits carboxylic acids to be separated from other organic compounds by 
extraction with appropriately basic solutions. Acids with long alkyl or alkenyl 
groups are called fatty acids. Their salts form colloidal particles in water 
called micelles in which the charged carboxyl end groups point outward and 
the hydrocarbon chains point inward. The mechanism of soap action is 
related to the ability of micelles to dissolve nonpolar substances ; the tendency 
of salts of fatty acids to concentrate at interfaces is also a factor. 

The C 4 to C 8 carboxylic acids have strong unpleasant odors. The infrared 
spectra of carboxylic acids shows broad absorption near 3000 cm" 1 (hydrogen- 
bonded O— H) and near 1750 cm" 1 (C=0 stretch). The carboxyl proton in 
the nmr absorbs at very low field (> 10 ppm from TMS). 

Some methods of preparing carboxylic acids are illustrated here. 

RMgX 



-► RCHO ► RC0 2 H < RCOZ Z = CI, OR, NH 2 



RCH=CHR RC^N RCHOHCHOHR 

In addition to these methods RX can be converted to RCH 2 C0 2 H by the 
malonic ester synthesis (see below). 
The acidity of the carboxyl group can be attributed to the inductive effects 



chap 13 carboxylic acids and derivatives 360 

of the carbonyl group and resonance stabilization of the ion RCOf. Electron- 
withdrawing substituents such as CI in R increase the acid strength by induc- 
tive withdrawal of charge from the carboxyl group (-/effect). 

Carboxyl groups can be protonated by strongly acidic systems to give 



OH 



cations of formula 
reaction. 



R-C 



\ 



These are intermediates in the esteriflcation 



ie 



OH 



In summary, the reactions of the carboxyl group are : 



RCO,H 



O 
« 
RC 

\ 
CI 

RCH 2 OH 
RH + C0 2 

R— R 
R— Br 



(with PC1 3 , PC1 5 , or SOCl 2 ) 

(with LiAlH 4 ) 

(occurs readily only if C-2 
carries a. —I substituent) 

(Kolbe electrolysis) 

(Hunsdiecker reaction) 



The 2 position of carboxylic acids can be halogenated via the acid halide 
(RCH 2 C0 2 H -*RCH 2 COX -» RCHXCOX -» RCHXC0 2 H). A halogen at 
the 2 position suffers ready displacement by nucleophiles. 

Derivatives of carboxylic acids include esters (RC0 2 R), acid halides 
(RCOX), anhydrides ((RCO) 2 0), and amides (RCONH 2 ). They can all be 
hydrolyzed to carboxylic acids and the first three also react with alcohols or 
amines to give esters and amides. 





II 




RC0 2 H 


R-C-X 




O 


O 




II 


II 




RC-OR" 


R-C- OR 




O 


(RCO) 2 




RC-NH 


*(est 


er interchang 


e in 



case of 



O 



O 



RC-OR - RC-OR) 

Lithium aluminum hydride reduction of carboxylic acid derivatives gives 
primary alcohols or amines. 

O 

II 
RC— NH 2 ^ 




RCH,OH 



RCH,NH, 



RC=N 



exercises 361 

Esters with anion-stabilizing substituents are moderately acidic and some 
of these, especially the /?-keto esters, undergo a number of useful reactions. 
They can be prepared by the Claisen condensation between two molecules of 
ester. 

O 

II 
2RCH 2 C0 2 C 2 H 5 ► RCH 2 CCHC0 2 C 2 H 5 

R 

Acetoacetic and malonic esters may be alkylated via their anions; acid 
hydrolysis then gives ketones and acids, respectively (acetoacetic and malonic 

ester syntheses). 

o o o 

II II II 

CH 3 CCH 2 C0 2 C 2 H 5 ► CH 3 CCHC0 2 C 2 H 5 ► CH 3 CCH 2 R 

R 
CH 2 (C0 2 C 2 H 5 ) 2 ► RCH(C0 2 C 2 H 5 ) 2 > RCH 2 CO a H 

a,/?-Unsaturated acids (double-bond between C-2 and C-3) undergo typical 
alkene-addition reactions except that unsymmetrical addends add in the anti- 
Markownikoff manner. y,5- or <5,e-Unsaturated acids (4- or 5-alkenoic acids) 
form lactones readily. 

Dicarboxylic acids have exalted values of K 1 and depressed values of K 2 
relative to monocarboxylic acids and this is mainly due to internal hydrogen 
bonding in the monoanion. 

Heating of dicarboxylic acids gives five- or six-membered rings if this is 
possible (either ketones by decarboxylation and dehydration or anhydrides by 
dehydration). The C 2 and C 3 dicarboxylic acids tend to undergo simple 
decarboxylation. 

o 

II II /^\ ^c^ 

HO— C C— OH ► C=0 or ^O or RC0 2 H 




o 



exercises 



13-1 Explain why the chemical shift of the acidic proton of a carboxylic acid, dis- 
solved in a nonpolar solvent like carbon tetrachloride, varies less with con- 
centration than that of the OH proton of an alcohol under the same condi- 
tions (see Section 13-1). 

13-2 A white solid contains ammonium octanoate mixed with naphthalene 
(CioHs) and sodium sulfate. Describe the exact procedure you would follow 
to obtain pure octanoic acid from this mixture. 

s> 
13-3 Would you expect the compound Ci 5 H3iN(CH 3 )3Cl e to form micelles in 

water ? 



chap 13 carboxylic acids and derivatives 362 

13-4 Write equations for a practical laboratory synthesis of each of the following 
substances from the indicated starting materials (several steps may be 
required). Give reagents and conditions. 

a. butanoic acid (w-butyric acid) from 1-propanol 

b. trimethylacetic acid from f-butyl chloride 

c. 2-methylpropanoic acid (isobutyric acid) from 2-methylpropene 

d. 2-bromo-3,3-dimethylbutanoic acid from r-butyl chloride 

e. /3-chloroethyl bromoacetate from ethanol and (or) acetic acid 
/. 2-methoxypentanoic acid from pentanoic acid 

g. 3,5,5-trimethyl-3-hexanol from 2,4,4-trimethyl-l-pentene (commer- 
cially available) 
h. 3,3-dimethylbutanal from 3,3-dimethylbutanoic acid 
i. 2,3,3-trimethyl-2-butanol from 2,3-dimethyl-2-butene 
j. cyclopentane from hexanedioic acid (adipic acid) 
k. cyclopropyl bromide from cyclopropanecarboxylic acid 



13.5 Give for each of the following pairs of compounds a chemical test, preferably 
a test tube reaction with a visible result, that will distinguish between the two 
substances. Write an equation for each reaction. 

a. HC0 2 H and CH 3 C0 2 H 

b. CH 3 C0 2 C 2 H 5 and CH 3 OCH 2 C0 2 H 

c. CH 2 =CHC0 2 HandCH 3 CH 2 C0 2 H 

d. CH 3 COBr and BrCH 2 C0 2 H 

H 2 C-CH 2 
I \ 

e. (CH 3 CH 2 CO) 2 and 0=C C=0 

O 

CO,H C0 2 H C0 2 H H 

\ / _, \ / 

/. C=C and C=C 

/ \ / \ 

H H H C0 2 H 

g. HC^CC0 2 CH 3 andCH 2 =CHC0 2 CH 3 
h. CH 3 C0 2 NH 4 and CH 3 CONH 2 
i. (CH 3 CO) 2 and CH 3 C0 2 CH 2 CH 3 

13-6 Explain how you could distinguish between the pairs of compounds listed in 
Exercise 13-5 by spectroscopic means. 

13-7 Write structural formulas for all 13 carboxylic acids and esters of formula 
CsHi O 2 . Provide the IUPAC name and, if possible, one other suitable name 
for each compound. 

13-8 Only two different products are obtained when the following four compounds 
are reduced with an excess of lithium aluminum hydride; what are they? 
(CH 3 ) 2 CHC0 2 H (CH 3 ) 2 CHC0 2 CH 3 (CH 3 ) 2 CHCONH 2 (CH 3 ) 2 CHC=N 

13-9 At high current densities, electrolysis of salts of carboxylic acids in hydroxylic 
solvents produce (at the anode) alcohols and esters of the type ROH and 
RC0 2 R. Explain. 



exercises 363 

13-10 Name each of the following substances by an accepted system: 



H 2/ C-CH 2 C0 2 C 2 H 3 
[ 2 C c 

H 2 C-CH 2 N C0 2 C 2 H 3 



b. CH 3 COCH[CH(CH 3 ) 2 ]C0 2 C 2 H 5 

c. C 6 H 5 CH 2 CH 2 COCl 

d. HOCH=CHC0 2 C 2 H 5 

e. C 2 H 5 OCOCOCH 2 C0 2 C 2 H 5 
/. HC0 2 C 2 F 5 



H,r c " 2 



g. | C=0 

n ^"-C-C0 2 CH 3 
I 
CH, 



CH 2 =CH— CH 2 

h. Yhco 2 h 



i. CH 3 CH 2 COCHCH 3 

I 

CH 2 CH 2 CH(CH 3 ) 2 
/. CH 3 CH 2 CH(CH=CH 2 )C0 2 H 
k. CH 3 COCH(C0 2 C 2 H 5 ) 2 

13-11 Write equations for the effect of heat on the 2, 3, 4, and 5 isomers of hydroxy- 
pentanoic acid. 

13-12 Predict the relative volatility of acetic acid, acetyl fluoride, and methyl 
acetate. Give your reasoning. 

13-13 By analogy with ester hydrolysis, propose a mechanism for each of the 
following reactions : 

a. C 6 H 5 C0 2 CH 3 + C 2 H 5 OH > C 6 H 5 C0 2 C 2 H 5 + CH 3 OH 

b. CH3COCI + CH 3 CH 2 OH * CH 3 C0 2 CH 2 CH 3 + HC1 

c. (CH 3 CO) 2 + CH 3 OH > CH 3 C0 2 CH 3 + CH 3 C0 2 H 

d. CH 3 CONH 2 + H 3 O ffi * CH 3 C0 2 H + NH 4 ® 

e. CH 3 CONH 2 + e OH * CH 3 C0 2 e + NH 3 

/. CH 3 COCl + 2NH 3 > CH 3 CONH 2 + NH4CI 

13-14 Why is a carboxylate anion more resistant to attack by nucleophilic agents, 
such as CH 3 O e , than the corresponding ester? 



chap 13 carboxylic acids and derivatives 364 

13-15 What can you conclude about the mechanism of acid-catalyzed hydrolysis 
of /3-butyrolactone from the following equation : 

18 OH 
H 2 C— C=0 h 2 18 Q, H® H 2C — Q 
H 2 C-0 H 2 COH ° 

O 

II 
13-16 Amides of the type R— C— NH 2 are much weaker bases (and stronger 

acids) than amines. Why? 
13-17 Write a plausible mechanism for the following reaction: 

P 

CH 3 C 

OC(CH 3 ) 3 + CH 3 OH — =_> CH 3 C0 2 H + CH 3 OC(CH 3 ) 3 
13-18 Other conceivable products of the Claisen condensation of ethyl acetate are 

O 

O O OCCH, 

II II / 3 

CH 3 CCHCOC 2 H 5 and CH 2 =C 

I \ 

CH 3 C=0 OC 2 H 5 

Explain how these products might be formed and why they are not formed 
in significant amounts. 

13-19 Suggest a reason why 2,4-pentanedione (acetylacetone) contains much more 
enol at equilibrium than ethyl acetoacetate. How much enol would you 
expect to find in diethyl malonate? In butan-3-on-l-al (acetylacetaldehyde) ? 
Explain. 

13-20 Write structures for all of the Claisen condensation products that may 
reasonably be expected to be formed from the following mixtures of sub- 
stances and sodium ethoxide : 

a. ethyl acetate and ethyl propanoate 

b. ethyl carbonate and acetone 

c. ethyl oxalate and ethyl trimethylacetate 

13-21 Show how the substances below may be synthesized by Claisen-type con- 
densations based on the indicated starting materials. Specify the reagents 
and reaction conditions as closely as possible. 

a. ethyl 2-propanoylpropanoate from ethyl propanoate 

b. CH 3 COCH 2 COC0 2 C 2 H 5 from acetone 

c. diethyl phenylmalonate from ethyl phenylacetate 

d. 2,4-pentanedione from acetone 

e. 2,2,6,6-tetramethyl-3,5-heptanedione from pinacolone Obutyl methyl 
ketone). 



exercises 365 

13-22 Why does the following reaction fail to give ethyl propanoate? 



CH 3 C0 2 C 2 H 5 + CH 3 I a °/ 2 5 > CH 3 CH 2 C0 2 C 2 H 5 

13-23 Show how one could prepare cyclobutanecarboxylic acid starting from 
diethyl malonate and a suitable dihalide. 

13-24 Would you expect vinylacetic acid to form a lactone when heated with a 
catalytic amount of sulfuric acid ? 

13-25 Fumaric and maleic acids give the same anhydride on heating, but fumaric 
acid must be heated to much higher temperatures than maleic acid to effect 
the same change. Explain. Write reasonable mechanisms for both reactions. 

13-26 7-Butyl acetate is converted to methyl acetate by sodium methoxide in 
methanol about one-tenth as fast as ethyl acetate is converted to methyl 
acetate under the same conditions. With dilute hydrogen chloride in metha- 
nol, ?-butyl acetate is rapidly converted to f-butyl methyl ether and acetic 
acid, whereas ethyl acetate goes more slowly to ethanol and methyl acetate. 

a. Write reasonable mechanisms for each of the reactions and show how 
the relative-rate data agree with your mechanisms. 

b. How could one use ls O as a tracer to substantiate your mechanistic 
picture? 



chapter 14 
i Optical isomerism^ enantiomers 

and diastereomers 



chap 14 optical isomerism, enantiomers and diastereomers 369 

Isomers are compounds that have the same molecular formula (e.g., 
C 4 H 1Q 0) but differ in the way in which the constituent atoms are joined 
together. The simplest form of isomerism is structural isomerism, in which the 
bonding sequence differs. Two of the structural isomers of C 4 H 10 O are 
1-butanol, CH 3 CH 2 CH 2 CH 2 OH, and 2-butanol, CH 3 CH 2 CHOHCH 3 . 
Stereoisomerism is the isomerism of compounds having the same structural 
formula but different arrangements of groups in space. We have already met 
one of the two forms of stereoisomerism, called geometrical (cis-trans) iso- 
merism, and we will now encounter the other, optical isomerism. We shall 
see that 2-butanol, but not 1-butanol, exhibits this rather subtle form of 
isomerism. 

Isomers, whether structural, geometrical, or optical are generally long-lived 
and isolable because isomerization usually requires that bonds be broken. In 
this way they differ from conformers, the many different spatial arrangements 
that result from rotations about single bonds (Section 2-2). It should be under- 
stood that isomers and conformers are not mutually exclusive. Thus while 2- 
butanol will be seen to have two stable optical isomers, each of these optical 
isomers exists as a dynamic mixture of conformations. We shall return to 
this point frequently in subsequent discussions. 

What are optical isomers? They are stereoisomers, some or all of which 
have the ability to rotate the plane of polarized light, that is, exhibit optical 
activity. What quality do optically active molecules possess that causes them 
to affect polarized light this way ? We shall see that it is actually a property they 
lack that is responsible for their optical activity and that this property is 
symmetry. Before going further, however, we shall examine the phenomenon 
of the polarization of light. 



14-1 plane-polarized light and the 
origin of optical rotation 

Electromagnetic radiation, as the name implies, involves the propagation 
of both electric and magnetic forces. At each point in a light beam, there is a 
component electric field and a component magnetic field which are perpendi- 
cular to each other and which oscillate in all directions perpendicular to the 
direction in which the beam is propagated. In plane-polarized light, the 
oscillation of the electric field is restricted to a single plane, the plane of 
polarization, while the magnetic field of necessity oscillates at right angles to 
that plane. Passing ordinary light through a split prism of calcite (a form of 
CaC0 3 ) known as a Nicol prism resolves the light into two beams, each of 
which is polarized and has half of the intensity of the original beam. (A sheet 
of Polaroid can also be used.) Light polarized by passage through one 
Nicol prism will not pass through a second Nicol prism set at right angles to 
the first one. Now if a transparent sample (usually a solution) of an optically 
active substance is placed between the two prisms, any change in the angle 



chap 14 optical isomerism, enantiomers and diastereomers 370 

of the plane of polarization in the solution can be detected (Figure 14-1), 
because the second prism will have to be rotated a certain number of degrees 
to be at right angles to the new plane of polarization and stop the light from 
coming through. An instrument that measures optical rotation this way is 
called a polarimeter (Figure 14-2). 

A clockwise rotation of the prism to produce extinction, as the observer 
looks toward the beam, defines the substance as dextrorotatory (rotates to the 
right), and we say that it has a positive ( + ) rotation. If the rotation is counter- 
clockwise, the substance is Ievorotatory (rotates to the left) and the compound 
has a negative ( — ) rotation. The angle of rotation is designated a. 

The question naturally arises as to why compounds whose molecules lack 
symmetry interact with polarized light in this manner. We shall oversimplify 
the explanation since the subject is best treated rigorously with rather complex 
mathematics. However, it is not difficult to understand that the electric forces 
in a light beam impinging on a molecule will interact to some extent with the 
electrons within the molecule. Although radiant energy may not actually be 
absorbed by the molecule to promote it to higher excited electronic-energy 
states (see Chapter 7), a perturbation of the electronic configuration of the 
molecule can occur. One can visualize this process as a polarization of the 
electrons brought about by the oscillating electric field associated with the 
radiation. This kind of interaction is important to us here because it causes 
the electric field of the radiation to change its direction of oscillation. The 
effect produced by any one molecule is extremely small, but in the aggregate 
may be measurable as a net rotation of the plane-polarized light. Molecules 
such as methane, ethane, and acetone, which have enough symmetry so that 
each is identical with its reflection, do not give a net rotation of plane- 
polarized light. This is because the symmetry of each is such that every optical 
rotation in one direction is canceled by an equal rotation in the opposite 
direction. However, a molecule with its atoms so disposed in space that it is 



Figure 14-1 Schematic representation of the vibrations of (a) ordinary 
light, and (b) plane-polarized light that is being rotated by interaction with 
an optically active substance. 





(b) 



sec 14.2 specific rotation 371 




ordinary 
light Y" 




polarized light 
(with new angle 
of polarization) 



V 



A- 



V 




adjustable 
Nicol prism 



■ tube containing solution of 
optically active substance 



Figure 14-2 Schematic diagram of a polarimeter. The plane of polarization has 
been rotated by passage through the solution of the optically active substance, 
and extinction of the beam of polarized light can only be restored by rotating 
the adjustable Nicol prism. 



not symmetrical to the degree of being superimposable on its mirror image will 
have a net effect on the incident polarized light. The electromagnetic inter- 
actions do not average to zero and such substances we characterize as being 
optically active. The structural characteristics of optically active molecules 
will be discussed beginning in Section 14-3. 



14-2 specific rotation 



The angle of rotation of the plane of polarized light a depends on the 
number and kind of molecules the light encounters — it is found that a varies 
with the concentration of a solution (or the density of a pure liquid) and on 
the distance through which the light travels in the sample. A third important 
variable is the wavelength of the incident light, which must always be specified 
even though the sodium D line (5893 A) is commonly used. (See also Section 
14-10.) To a lesser extent, a varies with the temperature and with the solvent 
(if used), which also should be specified. Thus, the specific rotation, [a], of a 
substance is generally expressed by the following formulas : 
for solutions, 



MS = 



100 a 
T~c~ 



for neat liquids, 



™-— t 



where a is measured degree of rotation; f is temperature; X is wavelength of 



chap 14 optical isomerism, enantiomers and diastereomers 372 

light; / is length in decimeters of light path through the solution; c is con- 
centration in grams of sample per 100 ml of solution; and d is density of 
liquid in grams per milliliter. 

For example, when a compound is reported as having [a]o 5 ° = — 100 
(c = 2.5, chloroform), this means that it has a specific levorotation of 100 
degrees at a concentration of 2.5 g per 100 ml of chloroform solution at 25°C 
when contained in a tube 1 decimeter long, the rotation being measured with 
sodium D light, which has a wavelength of 5893 A. 

Frequently, molecular rotation [M] is used in preference to specific rotation. 
It is related to specific rotation by the following equation : 

r^f Kft ' M 

M = Hiob- 

where M is the molecular weight of the optically active compound. Expressed 
in this form, optical rotations of different compounds can be compared 
directly because differences in rotation arising from differences in molecular 
weight are taken into account. 



14-3 optically active compounds with 
asymmetric carbon atoms 



A. ONE ASYMMETRIC CARBON 

Having discussed how optical activity is measured experimentally, we shall 
now consider the conditions of asymmetry (lack of symmetry) which are 
necessary for a compound to be optically active. The inflexible condition for 
optical activity is : the geometric structure of a molecule must be such that it is 
nonsuperimposable on its mirror image. Unless this condition holds, the mole- 
cule cannot exist in optically active forms. Asymmetry is, of course, a property 
of many objects you see around you. Each of your hands is asymmetric, that 
is, your right hand cannot occupy the same space that its mirror image (your 
left hand) fits into, as can be seen by trying to put your right hand into a glove 
that fits your left hand. The term chirality 1 is often applied to asymmetric 
objects (or molecules); it refers to their "handedness." At the molecular 
level, there are a number of types of structural elements that can make a 
molecule asymmetric and nonidentical with its mirror image. The most common 
and important of these is the asymmetrically substituted carbon atom, which 
is a carbon atom bonded to four different atoms or groups. If a molecule has 
such an asymmetrically substituted atom, the molecule will be nonidentical 
with its mirror image and will therefore be optically active. A simple example 
is 2-butanol [1 ], the C-2 of which is said to be asymmetric since it carries four 
different groups, hydrogen, hydroxyl, methyl, and ethyl. 

1 Pronounced ki'ralsd-e. 



sec 14.3 optically active compounds with asymmetric carbon atoms 373 



CH 3 H 


\V 


C 


/ \ 


CH 3 CH 2 OH 



[1] 

However, 2-butanol, as prepared from optically inactive materials (e.g., by 
the reduction of 2-butanone with lithium aluminum hydride), is optically 
inactive : 

CH 3 COCH 2 CH 3 — > CH 3 CH(OH)CH 2 CH 3 

It is a mixture of two isomeric forms, [2] and [3], which are mirror images of 
one another. 












The chemical and physical properties of the two forms are identical, 
except that they rotate the plane of plane-polarized light equally but in oppo- 
site directions. Mirror image forms of the same compound, such as [2] and 
[3], are called enantiomers. The 2-butanol, as prepared by the reduction of 
inactive 2-butanone, is optically inactive because it is a mixture of equal 
numbers of molecules of each enantiomer; the net optical rotation is there- 
fore zero. Such a mixture is described as racemic and is often designated by the 
symbol (+). 

Separation of the enantiomers in a racemic mixture is known as resolution, 
and conversion of the molecules of one enantiomer into a racemic mixture of 
both is called racemization. 



chap 14 optical isomerism, enantiomers and diastereomers 374 
Table 14-1 Physical properties of tartaric acids 



specific specific solubility 

rotation, [a]^ melting gravity in H 2 0, 

tartaric acid in H 2 point, °C of solid g/100gat25° 



meso 
(-) 
( + ) 
(±) 



-11.98° 
+ 11.98° 



140 
170 
170 
205 



1.666 
1.760 
1.760 
1.687 



]20 (i5-> 

147 
147 

25 



" meso-Tartaric acid is discussed in Section 14-3C. 

Although all the physical properties of pure enantiomers (apart from their 
optical properties) are identical, their melting points, solubilities, and any 
other properties involving the solid state are usually different from those of 
the racemic mixture. This is because a mixture of enantiomers packs differently 
in a crystal lattice (often more efficiently) than either enantiomer in the pure 
form. Tartaric acid is one such compound, and the physical properties of its 
various forms are given in Table 14T. 



B. PROJECTION FORMULAS 

We have distinguished the two enantiomers of 2-butanol, [2] and [3], with 
a picture of a three-dimensional model to show the tetrahedral arrangement of 
the groups about the asymmetric carbon atom. Clearly it will be inconvenient 
to do this in every case, particularly for more complex examples. It is necessary 
therefore to have a simpler convention for distinguishing between optical 
isomers. The so-called Fischer projection formulas are widely used for this 
purpose and, by their use, the enantiomers of 2-butanol are represented by 
[4] and [5]. 



CH 3 
H— C-OH 

I 

CH 2 CH 3 

[4] 



CH 3 

I 
HO— C— H 
I 
CH 2 CH 3 

[5] 



North 



The convention of the Fischer projections is such that the east and west bonds 
of the asymmetric carbon are considered to extend in front of the plane of 
the paper and the north and south bonds extend behind the, plane of the paper. 
This is shown more explicitly in structures [6] and [7]. Structures [2], [4], 
and [6] all correspond to one enantiomer while [3], [5], and [7] correspond to 
the other. 



CH 3 
H— C— OH 
CH 2 CH 3 

[6] 



CH 3 
HO— C— H 
CH 2 CH 3 

[7] 



sec 14.3 optically active compounds with asymmetric carbon atoms 37S 

With projection formulas, configuration is inverted (i.e., one enantiomer is 
changed into the other) each time two atoms or groups about the asymmetric 
atom (or asymmetric center, as it is often called) are interchanged. Clearly, if 
we interchange hydrogen for hydroxyl in [4], we have the enantiomer [5]. Less 
obvious is the interchange of methyl and hydrogen in [4] to give [8], which by 
our convention is equivalent to [5]. (Inspection of models may be helpful if 
the results of these operations are not clear.) 

H 

I 

CH 3 -C— OH 

CH 2 CH 3 

[8] 

It should be noted that [8] is not strictly a Fischer projection formula 
because, by the Fischer convention, the carbon chain is always written verti- 
cally. Note also that rotation of a Fischer projection formula 1 80° (but not 
90° or 270°) in the plane of the paper leaves the configuration unchanged. 



CH 3 CH 2 

I I 

-C— OH = HO— C — 

I I 

CH2CH3 CH 3 



For sake of uniformity, we normally show the carbon chain vertically with 
C-l at the top. 



C. COMPOUNDS WITH TWO ASYMMETRIC CARBON ATOMS. 
DIASTEREOMERS 

We have considered how a compound with one asymmetric carbon atom 
can exist in two optically active forms. However, it would be incorrect to 
infer from this that asymmetric carbon is a necessary condition for optical 
activity, because many compounds are known that have no asymmetric atoms 
but still exhibit optical isomerism (Section 14-4). It would be no more correct 
to infer that, because a molecule has two or more asymmetric carbons, it will 
necessarily be optically active. 

For example, consider a molecule with structure, configuration, and con- 
formation as in [9] with two asymmetric atoms, 1 and 2, the groups attached 
to atom 1 being the same as for atom 2, and the structure being so oriented 
that atom 1 is toward the front and atom 2 toward the rear. It can be seen 
that [9] is identical with its mirror image [10] by simply turning [9] end for 
end to give [11], which, when rotated 180° about the 1,2 bond axis as shown 
by the arrow, superimposes on [10]. A molecule with this configuration will 
be optically inactive, despite the presence of two asymmetric carbons, because 
it is superimposable on its mirror image. 



chap 14 optical isomerism, enantiomers and diastereomers 376 



mirror plane 




One could argue, however, that this condition does not hold for all con- 
formations of [9]. Certainly, the conformation [12] is not superimposable on 
its mirror image [13]. 



mirror plane 




Nevertheless, the molecule represented by [12] will not be optically active, 
because the asymmetric conformation [12] is rapidly converted into its 
enantiomer [13] by rotation about the bond joining the two asymmetric 
centers. Optical activity would be possible only if it were possible to " freeze " 
the molecule in one of the asymmetric conformations. 

It will be seen that the eclipsed conformation [14] of the same system has a 
plane of symmetry that bisects the molecule into two halves, each half being 
the mirror image of the other. Molecules that possess this sort of symmetry 
are frequently said to be internally compensated, since the optical rotations 
contributed by each asymmetric half are equal in degree but opposite in sign; 
the net rotation is therefore zero. 



sec 14.3 optically active compounds with asymmetric carbon atoms 377 




[14] 



This is perhaps more easily visualized from the Fischer projection formula 
of the conformations [9] through [14]. 







(+) 



— > 



zero net rotation 



(-) 



"W 



To generalize, if a substance has a plane of symmetry in at least one con- 
formation, it will not be resolvable into optically active forms. 

Among the many examples of compounds that have two asymmetric carbon 
atoms but are optically inactive because their molecules are free to pass 
through a conformation with a plane of symmetry are meso-taitatic acid [15]; 
m&yo-2,3-dibromobutane [16]; and cw-cyclohexane-l,2-dicarboxylic acid 
[17]. The asterisks in the formulas denote the asymmetric carbons. 



C0 2 H 


1* 

c- 


-OH 


I* 
c- 

1 


-OH 


1 

cc 


2 H 



[15] 



CH 3 

I. 
H-C-Br 

I* 
H-C-Br 

I 
CH 3 

[16] 



CH 2 — CH.H 
CH H C* 

N CH 2 — C* C0 2 H 



C0 2 H 



[17] 



The prefix meso denotes an optically inactive optical isomer of a compound 
that can exist in other optically active modifications. The meso form is con- 
sidered to be an optical isomer because it is a stereoisomer of the optically 
active forms and its lack of optical activity does not of itself change the type 
of isomerism involved. This will be clearer when we discuss erythro and threo 
forms later in this section. 

Tartaric acid and 2,3-dibromobutane each have a total of three optical 



chap 14 optical isomerism, enantiomers and diastereomers 378 

isomers, two optically active enantiomers, and one optically inactive meso 
form. These are shown in structures [18a], [18b], and [1 5] for tartaric acid, 
and [19a], [19b], and [16] for 2,3-dibromobutane. 

C0 2 H C0 2 H CH 3 CH 3 

I I I I 

H— C— OH HO— C-H H— C— Br Br— C-H 

I I I I 

HO— C-H H— C-OH Br-C-H H— C-Br 
I I I I 

C0 2 H CO z H CH 3 CH 3 

[18a] [18b] [19a] [19b] 

Stereoisomeric structures that are not enantiomers and thus not mirror 
images are called diastereomers. meso-Tartaric acid [15] and either one of the 
optically active tartaric acids [18a or 18b] are therefore diastereomers. Only 
[18a] and [18b] are enantiomers. Diastereomers usually have substantially 
different physical properties, whereas enantiomers have identical properties 
apart from the direction in which they rotate the plane of polarized light. This 
is illustrated in Table 14-1 for the tartaric acids. 

The reason for the difference in physical properties between diastereomers 
can be seen very simply for a substance with two asymmetric centers by 
noting that a right shoe on a right foot can be a mirror image and non- 
identical with a left shoe on a left foot, but is not expected to be a mirror 
image or have the same physical properties as a left shoe on a right foot or a 
right shoe on a left foot. In chemical terms, a pair of diastereomers have 
different internal distances in their molecules. If the groups about the central 
bonds in [15] and [18a] are rotated so that, for example, the two hydrogen 
atoms are at their closest, then the two hydroxy! groups will be separated by 
different distances in the two isomers. Although free rotation can take place 
about any of the single bonds in [15] and [18a], there are no conformations 
for which all the internal distances correspond. Enantiomers such as [18a] and 
[18b], on the other hand, can be arranged so that all the internal distances in 
one are identical with those in the other. 

Many compounds have two asymmetric carbons but cannot exist in an 
internally compensated meso form. This situation occurs when the two 
asymmetric carbons are differently substituted so that there is no conforma- 
tion in which any of the isomers has a plane of symmetry. As an example, 
consider the possible diastereomers of 2,3,4-trihydroxybutanal, [20] and [21], 
which are commonly known as erythrose and threose, respectively. They are 
represented below by the projection formulas [20a] and [21a], as well as by 
the conformational, or so-called "sawhorse," structures, [20b] and [21b]: 

CHO H CHO OH 

H-C-OH OHC^OH HO-C-H OHCf^H 

| or / | or / 

H-C-OH H^OH H-C-OH H^-OH 

CH 2 OH CH 2 OH CH 2 OH CH 2 OH 

erythrose threose 

[20a] [20b] [21a] [21b] 

Neither diastereomer has a plane of symmetry in any conformation, and 



sec 14.4 optically active compounds having no asymmetric carbon atoms 379 

consequently there are two pairs of enantiomers, or a total of four optical 
isomers. In general, a compound with n asymmetric carbons will have 2" 
possible optical isomers, provided there are no isomers with a plane of sym- 
metry arising from similarly substituted asymmetric carbons as in meso forms. 
The structures assigned to erythrose and threose have been established 
beyond question by the finding that ( — )-erythrose, which is a naturally 
occurring sugar, on oxidation with nitric acid gives wejo-tartaric acid, 
whereas ( + )-threose, which does not occur naturally, gives (-l-)-tartaric acid. 



CHO 
I 
H-C-OH 

I 
H-C-OH 
I 
CH 2 OH 

( — )-erythrose 



hno 3 



C0 2 H 
-C— OH 



I 



H- 

H-C-OH 
I 
CO,H 



oieso-tartaric acid 



CHO 
I 
H— C— OH 
I 
HO— C— H 
I 
CH 2 OH 

( + )-threose 



HNOj 



C0 2 H 

H— C— OH 
I 
HO— C— H 

I 



( + )-tartaric acid 



The terms erythro and threo are often used to designate pairs of diastereo- 
mers related to threose and erythrose. The erythro isomers, like ( + ) or ( — ) 
erythrose, are the ones which give a meso compound when the end groups are 
made identical while the ( + ) or ( — ) threo isomers are converted to optically 
active forms by the same transformation. 

14-4 optically active compounds having no 
asymmetric carbon atoms 

A. ALLENES AND SPIRANES 

Many compounds have no asymmetric carbon atoms, yet exhibit optical 
isomerism. This condition sometimes exists when there is a possibility for 
restricted rotation. For instance, in a molecule of allene, the two planes that 
contain the terminal methylene (CH 2 ) groups are mutually perpendicular 
because of the rigidity and directional character of the two cumulated double 
bonds. 




Consequently, an allene of the type XYC=C=CXY, in which X and Y are 
different, is asymmetric and can exist in two, nonsuperimposable, optically 
active forms. 



C = C=C 



-X 

-Y 



J^C-C- 



mirror plane 



chap 14 optical isomerism, enantiomers and diastereomers 380 

The optical isomerism of allenes was predicted by van't Hoff some 60 years 
before experiment proved him to be correct. The delay was mainly caused by 
practical difficulties in resolving asymmetric allenes. The first successful resolu- 
tions were achieved with the allenes [22] and [23]. 





^ 







c = c=cQ ! *^, c=c=c: 

"CO,CH,COiH 




[22J [231 

Other structures that can have the same type of asymmetry are the spiranes, 
which are bicyclic compounds with one atom (and only one) common to both 
rings. The simplest example is spiro[2.2]pentane [24]. 

CH 2 .CH 2 

CH 2 CH 2 

[24] 

The two rings of [24], or any spirane, cannot lie in a common plane; hence, 
provided that each ring is substituted so that it has no plane of symmetry, the 
substance can exist as one or the other of a pair of optically active enantio- 
mers. An example of a spirane that has been resolved is [25]. 



H.. / H \ ..CH 2v /" H 2 R / H \ ,.CH 2v /" 

H * N W CH ' X NH 2 H \h 2 / CH ' X NH 2 

[25] 



B. OPTICALLY ACTIVE BIPHENYLS 

In contrast to the asymmetric allenes, optical activity in the biphenyl series 
was discovered before it could be explained adequately. In fact, when the 
first biphenyl derivative was resolved (1922), there was considerable confusion 
as to the correct structure of biphenyl compounds, and the source of asym- 
metry was not known. It was subsequently established by dipole-moment and 
X-ray diffraction data that the benzene rings in biphenyls are coaxial. 

-?*5>-.- common axis 



With this information, the existence of stable optical isomers of o,o'- 
dinitrodiphenic acid [26] can be explained only if rotation about the central 
bond does not occur and, in addition, the two rings lie in different planes. A 
molecule of [26], which fulfills these requirements, is not superimposable on 
its mirror image, and will therefore be optically active. 



sec 14.5 absolute and relative configuration 381 
C ° 2H N0 2 




The lack of rotation about the pivot bond is caused by steric hindrance 
between the bulky ortho substituents. Evidence for this stems from the failure 
to resolve any biphenyl derivatives that are not substituted in ortho positions. 
Also, resolution of ortAo-substituted derivatives can be achieved only if the 
ortho substituents are sufficiently large. Thus, o,o'-difluorodiphenic acid [27], 
like the corresponding dinitro derivative [26], is resolvable but is more easily 
racemized than [26]. This is due to the smaller size of the fluorine atom relative 
to the nitro group, and with less interference from the ortho substituents the 
rings can more easily pivot about the central bond. Once they reach the 
planar conformation [27b] the asymmetry is lost and racemization results. 





HO,C F 



[27a] planar 

[27b] 




Racemization of hindered biphenyls does not involve bond breaking and 
hence these compounds are exceptions to the statement made earlier (Section 
2-6B) about the clear distinction between stereoisomers and conformers. 



14-5 absolute and relative configuration 

The sign of rotation of plane-polarized light by an enantiomer is not easily 
related to either its absolute or relative configuration. This is true even for 
substances with very similar structures, and we find that an optically active 
acid derivative having the same sign of rotation as the parent acid need not 
have the same configuration. Thus, given lactic acid (2-hydroxypropanoic 
acid), CH 3 CHOHC0 2 H, with a specific rotation + 3.82°, and methyl lactate, 
CH 3 CHOHC0 2 CH 3 , with a specific rotation — 8.25°, we cannot tell from the 
rotations alone whether the acid and ester have the same or a different arrange- 
ment of groups about the asymmetric center. The relative configurations have 
to be obtained by other means. 

If we convert ( + )-lactic acid into its methyl ester, we can be reasonably 
certain that the ester will be related in configuration to the acid, because 
esterification should not affect the configuration about the asymmetric 
carbon atom. It happens that the methyl ester so obtained is levorotatory so 
we know that (+)-lactic acid and (— )-methyl lactate have the same relative 
configuration at the asymmetric carbon, even if they possess opposite signs 
of optical rotation. However, we still do not know the absolute configuration; 
that is, we are unable to tell which of the two possible configurations of 



chap 14 optical isomerism, enantiomers and diastereomers 382 

lactic acid, [28a] or [28b], corresponds to the dextro or ( + ) acid and which 
to the levo or ( — ) acid. The term chirality, whose meaning was given as 
" handedness " in Section 14-3A, is often used for what we are calling absolute 
configuration. Chirality is the more general term — molecules and environ- 
ments can both possess chirality. 



C0 2 H 


C0 2 H 


— C-OH 


HO— C — H 


CH 3 


CH 3 


[28a] 


[28b] 



Until recently, the absolute configuration of no optically active compound 
was known with certainty. Instead, configurations were assigned relative to a 
standard, glyceraldehyde, which was originally chosen for the purpose of 
correlating the configurations of carbohydrates but has also been related to 
many other classes of compounds, including a-amino acids, terpenes, and 
steroids, and other biochemically important substances. Dextrorotatory 
glyceraldehyde was arbitrarily assigned the configuration [29a] and is known 
as d-( + )-glyceraldehyde. The levorotatory enantiomer [29b] is designated as 
L-( — )-glyceraldehyde. In these names, the sign in parentheses refers only to 
the direction of rotation, while the small capital letter d or L denotes the 
absolute configuration. The sign of rotation is sometimes written as dior ( + ) 
and / for ( — ), or dl for (±). 

CHO CHO 



H— C— OH 


HO— C— H 


CH 2 OH 


CH 2 OH 


[29a] 


[29b] 


D-( + )-glyceraldehyde 


l-( — )-gylceraldehyde 



At the time the choice of absolute configuration for glyceraldehyde was 
made, there was no way of knowing whether the configuration of ( + )- 
glyceraldehyde was in reality [29a] or [29b]. However, the choice had a 50% 
chance of being correct and we now know that [29a], the d configuration, is in 
fact the correct configuration of (+)-glyceraldehyde. This was established 
through use of a special X-ray crystallographic technique, which permitted 
determination of the absolute disposition of the atoms in space of sodium 
rubidium ( + )-tartrate. The configuration of ( + )-tartaric acid [18a] had been 
previously shown chemically to be the same as that of ( + )-glyceraldehyde. 
Consequently, the absolute configuration of any compound is now known 
once it has been correlated directly or indirectly with glyceraldehyde. 2 

2 Relating a compound's configuration to that of glyceraldehyde works well when the 
compound shares two or three of the same groups with the standard (H, OH, CHO, and 
CH 2 OH), but is rather arbitrary otherwise. For this reason a new procedure has been 
introduced, the Cahn-Ingold-Prelog system, in which a series of sequence rules determines 
configuration. The symbols R and s are used to designate configurations by this system 
instead of d and l. In the case of glyceraldehyde and closely related compounds d corres- 
ponds to r, and l corresponds to s. The r and s system has been extended to chiral mole- 
cules in general and it is possible, for example , to designate the optical isomers of [27] 
specifically as r and s. 



sec 14.5 absolute and relative configuration 383 



C T° 2H OHe T^ 2 " mo 9°> H C0 2 H 

CH 3 CH 3 in, ^ 



CH 3 



L-( + )-lactic ( + )-2-bromo- / ,\ i ■ 

acid propanoic acid ( + )-alanme 



Figure 14-3 Chemical tranformations showing how the configuration of 
natural (+)-alanine has been related to L-(+)-lactic acid and hence to 
L-(— )-glyceraldehyde. The transformations shown involve two S N 2 reactions 
which are stereospecific and invert configuration (Section 14-9). Reduction of 
the azide group leaves the configuration unchanged. 



In general, the absolute configuration of a substituent at an asymmetric center 
is specified by writing a projection formula with the carbon chain vertical 
and the lowest numbered carbon at the top. The D configuration is then the 
one that has a specified substituent on the bond extending to the right of the 
asymmetric carbon, while the l configuration has the substituent on the left. 

R, R, 

R 2 — C— X X— C-R 2 

I I 

R 3 R 3 

D configuration L configuration 

Compounds whose configurations are related to D-( + )-glyceraldehyde 
belong to the d series, and those related to l-( — )-glyceraldehyde belong to 
the l series. 

Many of the naturally occurring a-amino acids have been correlated with 
glyceraldehyde by the type of transformations shown in Figure 14-3. Here, 
natural alanine (2-aminopropanoic acid) is related to L-( + )-lactic acid and 
hence to L-( — )-glyceraldehyde. Alanine therefore belongs to the L series, and 
by similar correlations it has been shown that all of the a-amino acids which 
are constituents of proteins are L-amino acids. Many D-amino acids are com- 
ponents of other biologically important substances. 

When there are several asymmetric carbon atoms in a molecule, the configu- 
ration at one center is usually related directly or indirectly to glyceraldehyde, 
and the configurations at the other centers are determined relative to the 
first. Thus in the aldehyde form of the important sugar, ( + )-glucose, there 
are four different asymmetric centers, and so there are 2 4 = 16 possible 
stereoisomers. The projection formula of the isomer which corresponds to 
natural glucose is [30]. By convention for sugars, the configuration of the 
highest numbered asymmetric carbon is referred to glyceraldehyde to deter- 
mine the overall configuration of the molecule. For glucose, this atom is C-5, 
next to the CH 2 OH group, and has the hydroxyl on the right. Therefore, 
naturally occurring glucose belongs to the d series and is properly called 
D-glucose (see also Section 15-2). 



chap 14 optical isomerism, enantiomers and diastereomers 384 



'CHO 

2 I * 

H-C-OH 

si. 
HO— C— H 



H 



-C*-OH 



51* 
H— C— OH 

-~vr 

CH 2 OH 

[30] 



On the other hand, the configurations of a-amino acids possessing more 
than one asymmetric center are determined by the lowest numbered asym- 
metric carbon, which is the carbon alpha to the carboxyl group. Thus, even 
though the natural a-amino acid threonine has exactly the same kind of 
arrangement of substituents as the natural sugar threose, threonine by the 
amino acid convention belongs to the l series, while threose by the sugar con- 
vention belongs to the d series. 



C0 2 H 

~~-t-~~~ 

H 2 N— C — H l 

------I- 

H— C— OH 
I 
CH 3 

L-threonine 



CHO 

HO-C-H 

r -» - 

: H— C— OH d 

L 1-- 

CH 2 OH 

d-( — )-threose 



14-6 separation or resolution of enantiomers 

Since the physical properties of enantiomers are identical, they cannot 
usually be separated by physical methods such as fractional crystallization or 
distillation. It is only in the presence of another optically active substance that 
the enantiomers behave differently, and almost all methods of resolution (and 
asymmetric synthesis) are based on this fact. 

Perhaps the most general resolution procedure is to convert enantiomers to 
diastereomers, whose physical properties are not identical. For instance, if a 
racemic or d,l mixture of an acid is converted to a salt with an optically 
active base of the d configuration, the salt will be a mixture of two diastereo- 
mers, (d acid + d base) and (l acid + d base). These diastereomeric salts are 
not identical and not mirror images and will therefore differ in their physical 
properties. Hence, separation by physical means, such as crystallization, is in 
principle possible. Once separated, the acid regenerated from each salt will be 
either the pure d or l enantiomer : 



d,l acid 



d base 



(d acid + d base) 



-> (l acid + d base) 



d acid 



l acid 



sec 14.6 separation or resolution of enantiomers 38S 

Resolution of optically active acids through formation of diastereomeric 
salts requires adequate supplies of suitable optically active bases. Brucine 
[31a], strychnine [31b], and quinine [32] are most frequently used because 
they are readily available, naturally occurring, optically active bases. 




brucine, R=OCH 3 [31a] 
strychnine, R=H [31b] 



quinine 

[32] 



CH=CH 2 



For the resolution of a racemic base, optically active acids are used, such as 
( + )-tartaric acid, ( — )-malic acid, ( — )-mandelic acid, and ( + )-camphor-10- 
sulfonic acid. 



CH,CO,H 

I 
CHOH 

I 
C0 2 H 

malic acid 



C 6 H 5 

CHOH 
I 
C0 2 H 

mandelic acid 



H 3 C CH 3 

A,CH 2 S0 3 H 

AT 

camphor- 10-sulfonic acid 



To resolve an alcohol, an optically active acid may be used to convert the 
alcohol to a mixture of diastereomeric esters. High-molecular-weight acids 
(~400) are advantageous since they are likely to give crystalline esters, and 
these may usually be separated by fractional crystallization. 

Other, more specialized methods of resolution are also available. One 
procedure, which is excellent when applicable, takes advantage of differences 
in reaction rates of enantiomers with optically active substances. One enan- 
tiomer may react more rapidly, leaving an excess of the other enantiomer 
behind. As one example, racemic tartaric acid can be resolved with the aid of 
certain penicillin molds that consume the dextrorotatory enantiomer faster 
than the levorotatory enantiomer; as a result, almost pure ( — )-tartaric acid 
can be recovered from the mixture. 



(±)-tartaric acid + mold 



( — )-tartaric acid + more mold 



A crystallization procedure was employed by Louis Pasteur for his classical 
resolution of D,L-tartaric acid, but this technique is limited to very few cases. 
It depends upon the formation of individual crystals of each enantiomer. 
Thus, if the crystallization of sodium ammonium tartrate is carried out below 
27°, the usual racemate salt does not form; a mixture of crystals of the D and 
L salts forms instead. The two different kinds of crystals, which are mirror 



chap 14 optical isomerism, enantiomers and diastereomers 386 

images of one another, can be separated manually with the aid of a micro- 
scope and may be subsequently converted to the tartaric acid enantiomers by 
strong acid. 

Optical activity had been observed in quartz crystals and in solutions of 
various natural substances such as sugars well before Pasteur effected the 
first resolution in 1848. The explanation of optical activity on a molecular 
basis had to await the development of structural organic chemistry in the 
years following 1860 (Chapter 1). In 1874, Le Bel and van't Hoff showed that 
all known examples of optically active organic compounds possess asym- 
metrically substituted carbon atoms and, if the valences of carbon are 
tetrahedral, then each of the known optically active compounds would 
have to be able to exist in two mirror-image forms. This not only explained 
the existence of optically active organic compounds, it strengthened belief in 
the tetrahedral character of carbon and nurtured the growth of structural 
organic chemistry. 



14-7 asymmetric synthesis and asymmetric induction 

If you could prepare 2-hydroxypropanonitrile from acetaldehyde and 
hydrogen cyanide in the absence of any optically active substance and 
produce an excess of one enantiomer over the other, this would constitute an 
absolute asymmetric synthesis — that is, creation of an optically active com- 
pound in a symmetrical environment from symmetrical reagents. 



CN / f— ; 7 CN 

H~C^OH < HS / h" / -■■"-■* HO~C-H 




CH, I i / CH 



D-2-hydroxypropanonitrile L-2-hydroxypropanonitrile 

(product of attack (product of attack 

of CN e from above) of CN e from below) 

This is obviously unlikely for the given example because there is no reason 
for cyanide ion to have anything other than an exactly equal chance of attack- 
ing above or below the plane of the acetaldehyde molecule, thus producing 
equal numbers of molecules of each enantiomer. 

However, when an asymmetric center is already present and a second 
center is created, an exactly 1 :1 mixture of the two possible isomers (which are 
now diastereomers) is not expected because the transition states are diastereo- 
meric also and the reaction rates will differ. The optically active aldehyde [33] 
and methylmagnesium iodide react, for example, to give the two diastereo- 
meric alcohols [34a] and [34b] in a 2 : 1 ratio. 



sec 14.7 asymmetric synthesis and asymmetric induction 387 

O 

II 
C-H 

H 3 C ^ /h 




[33] 



1. CH 3 MgI 
.2. H® 

OH 




CH 



[34a] 

The formation of unequal amounts of diastereomers when a second 
asymmetric center is created in the presence of a first is called asymmetric 
induction. The degree of stereochemical control displayed by the first asym- 
metric center usually depends on how close it is to the second. The more 
widely separated they are, the less steric control there is. Another factor is the 
degree of asymmetry at the first asymmetric center. If all the groups there are 
very nearly the same electrically and sterically, not much stereochemical 
control is to be expected. 

Even when the asymmetric centers are close neighbors, asymmetric induc- 
tion is seldom 100% efficient in simple molecules. In biochemical systems, 
however, asymmetric synthesis is highly efficient. The photosynthesis of 
glucose [30] by plants from carbon dioxide and water gives the d enantiomer 
only, which means a completely specific synthesis at each of four asymmetric 
carbons. The l enantiomer is "unnatural" and, furthermore, is not even 
metabolized by animals. Similarly, all of the cc-amino acids which can be 
asymmetric and are constituents of proteins have the l configuration. 

CO a H 
NH,— C— H 

I 

R 

L-ct-amino acid 

The stereospecificity of biochemical reactions is a consequence of their 
being catalyzed in every case by enzymes, which are large protein molecules 
that possess many asymmetric centers and hence are themselves highly 
asymmetric. The stereospecificity of living organisms is related to their 
efficient operation since no organism could deal with all of the possible 
isomers of molecules with many asymmetric centers. Thus, if a protein mole- 
cule has 100 different asymmetric centers (a not uncommon or, in fact, large 
number), it will have 2 100 or 10 30 possible optical isomers. A vessel with a 
capacity of about 10 7 liters would be required to hold all of the possible 
stereoisomeric molecules of this structure if no two were identical. An organism 
so constituted as to be able to deal specifically with each one of these isomers 
would be very large indeed. 



chap 14 optical isomerism, enantiomers and diastereomers 388 

14-8 racemization 

Racemization, which is loss of optical activity, occurs with optically active 
biphenyls (Section 14-4B) if the two aromatic rings at any time pass through a 
coplanar configuration by rotation about the central bond. This can be brought 
about by heat, unless the ortho substituents are very large. 

The way in which other compounds — for example, those with asymmetric 
carbon atoms — are racemized is more complicated. Optically active carbonyl 

I I 
compounds of the type — CH— C=0, in which the a carbon (C-2) is asym- 
metric, are racemized by both acids and bases. From the discussion in Section 
12-1 on halogenation of carbonyl compounds, this is surely related to enoliza- 
tion. Formation of either the enol or the enolate anion will destroy the asym- 
metry at the a carbon so that, even if only trace amounts of enol are present 
at any given time, eventually all of the compound will be racemized. 

Base-catalyzed enolization : 



HO R 2 so HO 

\ / (-H®), \ ./ (+H®) \ // 

D ..c-c . c=-c , D --c-c 

R, R R, R R 2 R 



\ 


// 


.<:- 


-c 


i/ 


\ 


R 2 


R 



Acid-catalyzed enolization: 



H O R 2 OH 

\ // \ / 

_...c-c ^ ' c=c 

R 2 / \ / \ 

R, R R, R 



The racemization of an optically active secondary halide (e.g., 2-chloro- 
butane) may occur in solution. Usually, the more polar and better ionizing 
the solvent is, the more readily the substance is racemized. Ionization of the 
halide by an S N 1 process is probably responsible, and this would certainly 
be promoted by polar solvents (see Section 8-1 ID). All indications are that 
an alkyl carbonium ion once dissociated from its accompanying anion is 
planar; and, when such an ion recombines with the anion, it has equal prob- 
ability of forming the D and L enantiomers. 



I -a 9 

H-C-Cl ;= 

I 



A 

H X CH,CH, 



CH 3 

Cl e 



Cl-C-H 
I 
CH 2 CH 3 



D L 



Asymmetric alcohols are often racemized by strong acids. Undoubtedly, 
ionization takes place, and recombination of the carbonium ion with water 
leads to either enantiomer. 



sec 14.10 optical rotatory dispersion 389 



CH 3 
I 
H— C-OH 
I 
CH 2 CH 3 

D 

CH 3 

I 
HO-C-H 

I 
CH 2 CH 3 

L 



CH 3 

H ffl I e 

=± H-C-OH 2 , u _ 

CH 2 CH 3 ^^ 



CH 3 H^ 

(_H®) e I ^y^ 

. H 2 0-C-H ' 

I 



CH 3 
H CH 2 CH 3 



14-9 inversion of configuration 



In discussing the mechanism of S N 2 displacements we pointed out that 
backside attack would reduce electrostatic repulsions in the transition state 
to a minimum. Such a mechanism would cause inversion of configuration at 



HO e + /C-Cl 



[8e I Sel 
HO-C-Cl] 



-»" HO-C^ + Cl e 



the central carbon atom. When an optically active compound such as d-(+)- 
2-chlorobutane is converted to 2-butanol by aqueous base the product is, 
indeed, found to have an inverted configuration : 



CH 3 
H— C^Cl 
CH 2 CH 3 

d isomer 



HO e 



CH 3 
HO— C— H 
CH 2 CH 3 



The rotation of L-2-butanol is negative and that of D-2-chlorobutane is 
positive, but these relative rotations do not themselves prove that inversion 
occurs. The relative configurations of reactant and product must be deter- 
mined according to the principles discussed earlier in Section 14-5. Because 
secondary alkyl halides can also react, though slowly, by an S N 1 mechanism 
(Section 8-11 A) a certain amount of racemization usually accompanies this 
reaction. 



14-10 optical rotatory dispersion 

The sodium D line (5893 A) is used as the light source in most polarimeters. 
If we use light of a different wavelength can we expect a, the measured rota- 
tion, to change? Yes, and in many cases drastic changes occur even to the 
extent of a change in sign of rotation. The variation of optical activity with 
wavelength is known as optical rotatory dispersion (ORD). 



chap 14 optical isomerism, enantiomers and diastereomers 390 




Figure 14-4 Rotatory dispersion curves for ft-ans-lO-methyl-2-decalone [35] 
and cw-10-methyl-2-decalone [36]. (By permission from C. Djerassi, Optical 
Rotatory Dispersion, McGraw-Hill, New York, 1960.) 



Figure 14-4 shows a as a function of wavelength for the two ketones [35] 
and [36], the trans and cis isomers of 10-methyl-2-decalone, which have quite 
different stable conformations. 



O 



CH, 




CH, 




O 

[35] 



O 

[36] 



CH, 




CH, 




It can be seen that the effect of [35] on the polarization angle of the light 
increases as the wavelength of the light decreases, reaching a maximum just 
above 3000 A. A drastic drop then occurs followed by a change in sign. This 
type of behavior is called a positive Cotton effect. With some compounds, for 
example [36], a decrease in wavelength causes a change of sign to occur before 
the maximum is reached; this is the negative Cotton effect. 
The change in sign of the optical rotation can be linked to the presence of 



summary 391 

the carbonyl chromophore in [35] and [36] since the « ->• tt* absorption 
maximum of nonconjugated ketones occurs near 3000 A (Section 1 1 -2B). 
Light absorption and rotatory dispersion are hence associated phenomena. 

We can expect that any molecular property that is highly dependent on 
structure and configuration will find analytical application and ORD is no 
exception. The quite different ORD curves that are often obtained for various 
members of a set of diastereomers can be highly useful in assigning config- 
urations. 



summary 

Optical activity, the ability to rotate plane-polarized light, is exhibited by 

compounds lacking molecular symmetry and is detected and measured by an 

instrument called a polarimeter. The angle of rotation of a liquid or solute is 

expressed either in terms of a, the measured degree of rotation; [a], the specific 

rotation (adjusted for concentration and tube length); or molecular rotation 

[M] (adjusted for molecular weight). 

At the molecular level, the structural feature most commonly associated 

with optical activity is the asymmetrically substituted carbon atom, a carbon 

x 
I 
atom with four different groups attached, w— C— y. A molecule with a single 

z 
such structural unit is not superimposable on its mirror image and there are 
two possible configurations, called enantiomers. 

Optical isomers, like geometrical isomers, have the same bond structures 
and differ only in the spatial arrangements of the atoms and hence are stereo- 
isomers. Unlike cis-trans pairs, enantiomers have identical internal distances 
and hence identical melting points, solubilities, and so on. A 50 : 50 mixture 
of enantiomers, which is called a racemic mixture and does not rotate the 
plane of polarized light, may have different melting or solubility characteristics 
than either of its components because of crystal-packing effects. Resolution 
is the process of separation of a racemic mixture into its components and 
racemization is the reverse process. 

Fischer projection formulas are used to represent in two dimensions the 
three-dimensional configurations of enantiomers, such as 2-butanol. 

CH 3 CH 3 

H-C-OH = H— C-OH 
I ! 

CH 2 CH 3 CH 2 CH 3 

Compounds with two identical asymmetric carbon atoms exist in three 
forms, two optically active enantiomers and an optically inactive meso form, 
which possesses a plane of symmetry. In addition, the two enantiomers in 
equal amount form a racemic mixture. If a compound possesses two non- 
identical asymmetric carbon atoms, four optical isomers are possible — two 



chap 14 optical isomerism, enantiomers and diastereomers 392 

pairs of enantiomers. Molecules of different enantiomeric pairs are called 
diastereomers of one another and have different physical properties. (Dia- 
stereomers are stereoisomers that are not mirror images and hence a meso 
compound and one of the enantiomers of the same structure are also dia- 
stereomers.) Thus, 2,3-butanediol [1] exists in two enantiomeric forms and one 
meso form whereas 2,3-pentanediol [2] exists in four optically active forms 
which constitute two pairs of enantiomers. 

CH3CHOHCHOHCH3 CH 3 CHOHCHOHCH 2 CH 3 

[1] [2] 

Optical activity results from molecular asymmetry and certain allenes, 
spiranes, and hindered biphenyls that do not possess asymmetric carbon 
atoms can exist in optically active forms. 

The sign of rotation of an enantiomer, ( + ) or d, ( — ) or /, reveals neither the 
molecule's absolute configuration nor its configuration relative to some other 
compound. Chemical means, however, can be used to determine the latter 
and this knowledge, combined with X-ray diffraction studies, has enabled 
many absolute configurations to be determined. The symbols D and l are 
used to designate absolute configuration and configurations are assigned 
relative to glyceraldehyde. ( + )-Glyceraldehyde is known to be the D isomer. 

CHO 

H-C-OH 

CH 2 OH 

d-( + )-glyceraldehy de 

The number of optical isomers possible for a compound with n different 
asymmetric carbon atoms is 2", which corresponds to 2"/2 pairs of enantio- 
mers. Enantiomers A D and A h can be separated by appropriate combination 
of the racemic mixture with a second optically active compound, B D , which 
thus produces a pair of diastereomers, A D B D and A L B D . Because the members 
of the pair of diastereomers possess different physical properties, they can be 
separated and converted to the separate optical isomers. 

. B D A D B D ► A D 

A D A L mixture ► 

A L B D ► A L 

Occasionally a racemic mixture will crystallize to give a mixture of d crystals 
and L crystals, which can be separated mechanically (method of Pasteur). 

A chemical reaction that creates an asymmetric center in a molecule will 
produce equal amounts of the two enantiomers unless asymmetry is already 
present in the molecule, in which case unequal amounts of two diastereoiso- 
mers are expected by asymmetric induction. 

Racemization occurs when a symmetrical substance is produced at any 
stage in a reaction undergone by an optically active compound. Inversion of 
configuration normally occurs during S N 2 reactions as a result of backside 
attack by the nucleophile. 



exercises 393 

When the wavelength of polarized light is varied the amount of rotation 
caused by an optically active substance changes (optical rotatory dispersion), 
and may even undergo a change in sign. The sign change is normally associated 
with the absorption of visible or ultraviolet light by a chromophore in the 
molecule. 



exercises 

14-1 Many familiar objects, such as gloves, and screws, are nonidentical with their 
mirror images. Is it reasonable to expect such objects to be optically active 
provided they are transparent to light, or are further conditions necessary? 

14-2 Which of the following compounds exist in optically active forms? Identify 
those that have a meso form. 

a. 3-heptanol g. bis(4-chlorophenyl)carbinol 

b. 4-heptanol h. 1,1-dimethylcyclobutane 

c. 3,4-dibromohexane i. 1,2-dimethylcyclobutane 

d. 1,6-dimethoxy-l-hexene j. 1,3-dimethylcyclobutane 

e. diphenylcarbinol k. 2,3-dimethylbutanoic acid 
/. 4-chlorodiphenylcarbinol (Section 20-1A) 

14-3 Identify the compounds listed in Exercise 14-2 that exhibit geometrical 
isomerism. 

14-4 Citric acid (2-hydroxy-l,2,3-propanetricarboxylic acid) is the principal acid in 
citrus fruits. Can citric acid or any of its methyl esters exhibit optical activity ? 

14-5 The compound l,l,l-trifluoro-2-bromo-2-chloroethane is the well-known 
anesthetic halothane. Can it exist in stable optically active forms? Make 
sawhorse drawings of the different optically active conformations of this 
substance. 

14-6 Draw the chair form of cw-cyclohexane-l,2-dicarboxylic acid. Is it identical 
with its mirror image? Why is the compound not resolvable? How many 
stereoisomers are possible for cyclohexane-l,3-dicarboxylic acid and, of these, 
which are optical isomers and which are geometric isomers ? 

14-7 Draw structures similar to [9] through [13] for all the possible different 
staggered conformations of (+)-tartaric acid [18]. Are any of these identical 
with their mirror images ? How many optically active forms of tartaric acid 
could there be altogether if rotation were not possible about the 2,3 bond 
and only the staggered conformations were allowed ? 

14-8 Write projection formulas for (+)-erythrose and (— )-threose. Which tartaric 
acids will they give on oxidation? 

14'9 Draw projection formulas for all of the possible optical isomers of 2,3,4- 
trihydroxypentanoic acid. 



chap 14 optical isomerism, enantiomers and diastereomers 394 

14-10 Camphor has two asymmetric carbons, but only two optical isomers are 
known. Explain. (Models may be helpful.) 



CH, 



CH, 




O 



camphor 

14-11 Would the following structures be resolvable into optically active isomers? 
Show the structures of the possible isomers. 



w c H,C-CH 2 H 

3 ^\ 7 \ / 

a. ,C C=C 

f/ \ / \ 



H 



H,C— CH, 



CO,H 



H,C H 2, c_ CH 2 



X c=c=c c, 

H H 2 C-CH 2 H 2 C-CH 2 



14-12 Which of the following biphenyl derivatives would you expect might give 
stable enantiomers ? Show your reasoning. 




b. 




d. 




C0 2 H 



14-13 Which of the following projection formulas represent the same optical 
isomers ? Write each in its proper form as a Fischer projection formula of 
3-amino-2-butanol. 

H 



CH 3 

I 
H— C— OH 

I 
H-C— NH 2 

I 
CH 3 

H 

I 

CH 3 -C-NH 2 

I 

H— C— OH 

I 

CH 3 



CH 3 — C— OH 

I 
CH 3 -C-NH 2 

I 
H 

CH 3 
I 
HO— C— H 

I 
H— C— NH 2 
I 
CH 3 



OH 

I 
CH 3 -C-H 

I 
C 

I 
CH 3 

H 

I 
CH 3 -C— OH 

I 
H 2 N-C-CH 3 

I 
H 



exercises 395 

14-14 Draw Fischer projection formulas for all the possible different optical 
isomers of the following substances : 

a. 1,2,3,4-tetrachlorobutane 

b. methylethylpropylboron 

c. 2,3-dibromopropanoic acid 

d. triisopropylmethane 

e. 3-bromo-2,5-hexanediol 

/ C0 2 CH 3 

CHO 

CH 2 
/ 
CHO 

C0 2 CH 3 

g. methyl hydrogen tartrate 
h, s-butyl lactate 

1445 Predict the stereochemical configuration of the products from each of the 
following reactions. Write projection formulas for the starting materials and 
products. 

a. D-2-butanol with acetic anhydride 

b. D-methylethylisobutylcarbinol with hydrochloric acid 

c. D-glycerol monoacetate with aqueous sodium hydroxide 

d. D-s-butyl ?-butyl ketone with bromine and dilute base 

14-16 Explain how one could determine experimentally whether hydrogen peroxide 
in formic acid adds cis or trans to cyclopentene, assuming the possible 
addition products to be unknown. 

14-17 Devise a reaction scheme for relating configuration of (+)-2-butanol to 
glyceraldehyde. Think carefully about the reaction mechanisms involved in 
devising your scheme. 

14-18 Discuss possible procedures for resolution of ethyl D,L-lactate (ethyl 
2-hydroxypropanoate, bp 155°) into ethyl D-lactate and ethyl L-lactate. 

14-19 How could you tell whether a chloroform solution of an optically active 
compound showing a rotation of —100° was actually levorotatpry by —100° 
or dextrorotatory by +260°? 

14-20 Solutions of optically active 2,2'-diiodobiphenyl-5,5'-dicarboxylic acid race- 
mize at a measurable rate on heating. Racemization of active 2,3,2'3'- 
tetraiodobiphenyl-5,5'-dicarboxylic acid goes many thousand times more 
slowly. (For nomenclature see Section 20-1 A.) Make a scale drawing of the 
transition state (planar) for racemization; deduce from it the reason for 
the very slow racemization of the tetraiodo diacid. Use the following bond 
distances (note that the benzene ring is a regular hexagon) : 



chap 14 optical isomerism, enantiomers and diastereomers 396 

C—C (benzene ring) =1.40 A 
C— C (between rings) = 1.47 A 
C-H =1.07 A 

C-I = 1.63 A 

The interference radii of iodine and hydrogen are 2.15 and 1.20 A, respec- 
tively. 

14-21 Compounds of the type shown below have been found to be resolvable into 
two optically active forms. Explain. 




-C0 2 H 

H 3 C CH 3 

14-22 Can the structures CH 2 =C=CBr 2 and BrHC=C=C=CHBr be optically 
active ? Explain. 

14-23 Write Fischer projection formulas for each of the following substances, 
remembering, where necessary, that t> and l isomers of substances with 
more than one asymmetric carbon are always enantiomers, not diastereo- 
mers. 

a. L-alanine (L-2-aminopropa- e. D-?Areo-2,3-dihydroxybutanoic 

noic acid) acid 

b. D-2,3-butanediol / L-ery?/i/"0-2,3-butanediol mono- 

c. D-threonine methyl ether 

d. L-glucose 



chap IS carbohydrates 399 

Carbohydrates are a major class of naturally occurring organic compounds 
which came by their name because they usually have the general formula 
C x (H 2 0) y . Among the more well-known carbohydrates are the various 
sugars, the starches, and cellulose, all of which are important for the 
maintenance of life in both plants and animals. 

Carbohydrates are formed in green plants as the result of photosynthesis, 
which is the chemical combination or " fixation " of carbon dioxide and water 
by utilization of energy gained through absorption of visible light : 

xC0 2 +xH 2 — > (CH 2 0), + x0 2 

green plants 

carbohydrate 

Although many aspects of photosynthesis are not yet well understood, the 
primary process is clearly the excitation of the green plant pigment chloro- 
phyll-a (Figure 15-1) by absorption of light; the energy of the resulting 
activated chlorophyll-a molecules is used to oxidize water to oxygen and to 
reduce carbon dioxide. One of the first-formed products in the fixation of 
carbon dioxide is believed to be D-glyceric acid. 

C0 2 H 
I 
H-C-OH 

CH 2 OH 

D-glyceric acid 

From this compound, the plant carries out a series of enzyme-catalyzed 
reactions which result in the synthesis of simple sugars, such as glucose 
(C 6 H 12 6 ), and polymeric substances, such as the starches and cellulose 
(C 6 H 10 O 5 )„, with«> 1000. 



Figure 15-1 The structure of chlorophyll-a. The coordination of the un- 
shared electron pairs on two of the nitrogens to the magnesium is indicated 
by dashed lines. Other resonance structures can also be written leading to the 
conclusion that all four bonds to magnesium are equivalent. 




H .-■ 

CH 2 C0 2 CH£> 

I 
CO 2 C 20 H39 



chap 15 carbohydrates 400 
Table 15*1 Some classes and examples of naturally occurring carbohydrates 





MONOSACCHARIDES 


Pentoses (CsHjoOs) 


Hexoses (C 6 H 12 6 ) 


L-arabinose 




D-glucose 


D-ribose 




D-fructose 


D-xylose 




D-galactose 
D-mannose 




OLIGOSACCHARIDES 


Disaccharides (C12H22OU) 




Trisaccharides (C 18 H 3 20 16 ) 


sucrose (D-glucose + D-fructose) 




raffinose (D-glucose + D-fructose 


maltose (D-glucose + D-glucose) 




+ D-galactose) 


lactose (D-galactose + D-glucose) 






POLYSACCHARIDES 


(C6H 10 O 5 )„ 


Plants 




Animals 


starch 




glycogen 


cellulose 







15-1 classification of carbohydrates 



The simplest sugars, called monosaccharides, are polyhydroxy aldehydes 
or ketones, usually containing five or six carbon atoms. If several of these are 
joined together by acetal-type linkages the molecule is called an oligosac- 
charide. If 10 or more are joined this way the resulting polymer is called a 
polysaccharide. The names of all individual monosaccharides and oligosac- 
charides (sugars) end in -ose. Table 15T shows some of the important natur- 
ally occurring carbohydrates. 

The aldopentoses are five-carbon sugars containing an aldehyde function. 
The most abundant of these are L-arabinose, D-ribose, D-xylose, and 2-deoxy- 
D-ribose. (Deoxy means without oxygen; thus, 2-deoxy-D-ribose is D-ribose 
in which the 2-hydroxyl group is replaced with hydrogen.) Their structures 
and configurations are given in Table 15-2. We shall see later that they exist 
predominantly in a cyclic form involving combination of the carbonyl group 
and a hydroxyl group further down the chain. D-Ribose is a component of 
ribonucleic acid (RNA) and vitamin B 12 and some coenzymes. 2-Deoxy-D- 
ribose is part of the deoxyribonucleic acid(DNA) chain. Both RNAandDNA 
will be met again later in the book. 

There are three important aldohexoses: D-glucose, D-mannose, and 
D-galactose. Of these D-glucose is by far the most abundant in nature and we 
shall examine its chemistry in detail in the next section. 

There is one other important monosaccharide, D-fructose. This is a keto- 
hexose which is found in fruit juices and in honey. Linked to glucose it forms 
the disaccharide sucrose, which is common table sugar. The structures and 



sec IS. 1 classification of carbohydrates 401 
Table IS -2 Typical pentoses and hexoses 



pentoses 


hexoses 


CHO 

1 


CHO 

| 


H-C-OH 

j 


H— C— OH 

I 


HO-C-H 

I 


HO— C— H 

| 


HO— C— H 

| 


H— C— OH 

1 


CH 2 OH 


H— C— OH 

j 




CH 2 OH 


L-arabinose 


D-glucose 


CHO 

1 


CH 2 OH 

t 


H— C— OH 

| 


c=o 

j 


H— C— OH 

1 


HO— C— H 

j 


H— C— OH 

I 


H— C— OH 

1 


CH 2 OH 


H— C— OH 

J 


D-ribose 


CH 2 OH 




D-fructose 


CHO 

1 


CHO 

1 


CH 2 

| 


HO— C— H 

1 


H— C— OH 

1 


HO— C— H 

1 


H— C— OH 

| 


H— C— OH 

1 


CH 2 OH 


H— C— OH 

| 


2-deoxy-D-ribose 


CH 2 OH 




D-mannose 


CHO 

1 


CHO 

1 


H— C— OH 

1 


H— C— OH 

1 


HO— C— H 

| 


HO— C— H 

1 


H— C— OH 

i 


HO— C— H 

1 


CH 2 OH 


H— C—OH 

J 


D-xylose 


CH 2 OH 




D-galactose 



configurations of fructose and the three important aldohexoses are given in 
Table 15-2. As with the aldopentoses, these molecules exist largely in the 
cyclic form. 
Some carbohydrates found in nature contain an amino group in place of 



chap 1 S carbohydrates 402 

an hydroxyl group. The antibiotic streptomycin, for example, contains an 
amino derivative of glucose, as do other compounds of structural and immu- 
nological importance (Sections 15-8 and 15-9). 



IS- 2 glucose 



D-Glucose, the most abundant monosaccharide, occurs free in fruits, plants, 
and honey, and in the blood and urine of animals, and combined in many 
oligosaccharides and polysaccharides. It is one of the optical isomers of the 
aldohexoses, all of which have structure [1]. 

CHO 

*CHOH 

*CHOH 

*CHOH 

*CHOH 
I 
CH 2 OH 

[1] 

The carbons labeled with an asterisk in [1] are asymmetric, and there are 
therefore 2 4 , or 16, possible optically active forms. All are known — some 
occur naturally and others have been synthesized. The problem of identifying 
glucose as a particular one of the 16 possibilities was solved by Emil Fischer 
during the latter part of the nineteenth century. The configurations he deduced 
for each of the asymmetric carbons (C-2 to C-5) are shown in the projection 
formula [2]. (Remember that the horizontal bonds in Fischer projection 
formulas such as [2] are understood to extend forward out of the plane of the 
paper and the vertical bonds extend back behind the paper.) 

'CHO 

HO— 3 C— H 

J 
H-C-OH 

'? 

°CH 2 OH 

[2] 

Fischer was well aware that natural glucose could be the enantiomer of 
[2]. His original guess proved to be correct, because the configuration at C-5 
is that of the simplest "sugar," D-(+)-glyceraldehyde. This was arbitrarily 
assigned as [3] and later shown to be correct (Section 14-5). Therefore, 
natural glucose is specifically D-glucose. 



sec IS. 2 glucose 403 

CHO 
H— C— OH 
CH 2 OH 

[3] 

The relationship between naturally occurring glucose, mannose, and 
fructose was uncovered by Fischer with the help of the phenylhydrazine reac- 
tion. Phenylhydrazine is one of the reagents used to characterize aldehydes 
and ketones (Table 11-4), and even though only a small fraction of the sugar 
molecules are in the noncyclic carbonyl form, hydrazone formation occurs and 
eventually all the sugar is converted to a phenylhydrazone. With 2-hydroxyal- 
dehydes or ketones the reaction does not stop at this stage, and the adjacent 
alcohol group is oxidized by a second molecule of phenylhydrazine; the 
resulting carbonyl group then reacts with a third molecule of reagent. The 
reaction stops at this stage giving a crystalline derivative called an osazone. 
With glucose, the product would be named glucose phenylosazone. Although 



CHO C 6 H 3 NHNH 2 CH=NNHC 6 H 5 CH=NNHC 6 H 5 

| ► | ► | 

CHOH CHOH C=0 

I I I 



glucose 



glucose 
phenylhydrazone 



CH=NNHC 6 H 5 
I 
C=NNHC 6 H 5 

I 

glucose 
phenylosazone 



glucose, mannose, and fructose all give different phenylhydrazones, they give 
the same phenylosazone. 



CHO 

(CHOH), C * HsNHNH - 
I * excess 

CH,OH 



CH=NNHC 6 H 5 

C=NNHC 6 H 5 

(CHOH), 
I 
CH 2 OH 



C<,H 5 NHNH 2 



glucose and phenylosazone from glucose, 

mannose mannose, and fructose 



CH 2 OH 

I 

c=o 

I 

(CHOH) 3 
CH 2 OH 

fructose 



This shows that the configuration at C-3, C-4, and C-5 must be the same for 
all three sugars. Further, glucose and mannose differ only in configuration at 
C-2 ; they are said to be epimers. Fischer used many pieces of evidence such as 
this to deduce the configuration of each asymmetric carbon atom in each of 
the 16 optical isomers of the aldohexoses (an amazing accomplishment in the 
days when instrumental analysis, other than polarimetry, was unknown). 

Glucose and related sugars have the properties expected of compounds con- 
taining several alcoholic hydroxyl groups. They tend to have high water 
solubilities, their hydroxyl groups can be oxidized to give aldehydes and 
ketones, and they form esters ; being glycols they are cleaved by periodic acid, 
and so on. 

Glucose is also an aldehyde and although it has many of the reactions of 
aldehydes it lacks others. For example, it forms certain carbonyl derivatives 



chap 15 carbohydrates 404 

(e.g., oxime, phenylhydrazone, and cyanohydrin) ; it can be reduced to the 
hexahydric alcohol sorbitol, and oxidized with bromine to gluconic acid (a 
monocarboxylic acid). With nitric acid, oxidation proceeds further to give the 
dicarboxylic acid, D-glucaric acid. 



CH 2 OH 

H— C-OH 
I 
HO-C-H 

H-C-OH 

I 
H-C-OH 

I 
CH 2 OH 

sorbitol 



CHO 
I 
H-C-OH 

Na-Hg HO-C-H 

H-C-OH 

I 
H-C-OH 

I 
CH 2 OH 

D-glucose 



Br 2 



I 
H-C-OH 

I 
HO-C-H 

I 
H-C-OH 

I 
H-C— OH 

I 



D-gluconic acid 



HN0 3 



C0 2 H 

-C 

I 
HO-C-H 



H- 



OH 



H-C— OH 
H-C-OH 
C0 2 H 

D-glucaric acid 



Glucose will also reduce Fehling's solution (Cu'^Cu 1 ) and Tollen's 
reagent (Ag'-^Ag ). However, it fails to give a bisulfite addition com- 
pound and it forms two different monomethyl derivatives (called methyl 
a-D-glucoside and methyl /?-D-glucoside) under conditions which normally 
convert an aldehyde to a dimethyl acetal. 



H 

/ 

R-C + 2CH 3 OH 
O 

C 6 H 12 6 + CH 3 OH 



H 



H a> 



-► R-C-OCH3 + H,0 
I 
OCH3 

(C 6 H u 5 )OCH 3 + H 2 

methyl a- and /?-D-glucoside 

The above behavior suggests that the carbonyl group is not free in glucose 
but is tied up in combination with one of the hydroxyl groups, which turns 
out to be the one at C-5, to form a cyclic hemiacetal represented by [4]. 



,1 
CHOH 

2 I 
H— C— OH 

,1 
HO— C— H 

H-C-OH 

,1 
H-C-O— 



[4] 

A new asymmetric center is created at C-l by hemiacetal formation, and 
there are therefore two stereoisomeric forms of D-glucose, a-D-glucose and 
/?-D-glucose. 



H-C— OH 



a-D-glucose 



,1 
HO— C — H 

I 

/?-D-glucose 



sec 15.3 cyclic structures 405 

A specific term is used to describe carbohydrate stereoisomers differing only 
in configuration at the hemiacetal carbon; they are called anomers. Although 
Fischer projection formulas are useful for indicating configuration in open- 
chain molecules they are less suitable for cyclic structures such as [4] because 
they do not clearly show the spatial relationships of the molecules. 



15-3 cyclic structures 



Because glucose and most other sugars exist predominantly (>99%) in the 
cyclic hemiacetal structure we will want to draw the structures and configura- 
tions as clearly as possible. The Haworth projection formulas [5] and [7] were 
adopted before chemists became aware of the conformational mobility of six- 
membered rings. They are still used extensively but are rather less informative 
than conformational formulas such as [6] and [8]. 




H CH,OH 



HO 
OH HO 



H OH 

[5] 

Haworth projection 

formula 




H" OH 

[6] 

conformational 

formula 



a-D-glucose 

X-Ray diffraction studies have shown that crystalline a-D-glucose has the 
chair conformation [6]. The Haworth and conformational formulas can be 
related to the chainlike Fischer projection formulas in the following way. 
A group that extends to the right in a Fischer projection formula extends 
downward in a Haworth projection or conformational formula. Furthermore, 
for D sugars the OH at C-l will be down in the a form and up in the fi form. 

/9-D-Glucose differs from a-D-glucose only in the configuration at C-l and 
its Haworth and conformational formulas are shown in [7] and [8]. Note that 
all the OH and CH 2 OH groups in this molecule occupy equatorial positions. 
This suggests that the /? anomer should be of slightly lower energy because the 
hydroxyl at C-l in the a anomer will suffer 1,3 interactions with the hydrogens 
at C-3 and C-5. 



CH 2 OH 

HON^H 
H OH 

[7] 

Haworth projection 

formula 



HCH,OH 



HO 




[81 

conformational 

formula 



/?-D-glucose 



chap IS carbohydrates 406 

Since the oxide ring is six membered in some sugars and five membered in 
others, it is helpful to use names that indicate the ring size. The five- and 
six-membered oxide rings bear a formal relationship to the cyclic ethers, 
furan and pyran. Hence, the terms furanose and pyranose have been coined 

II II HC* CH 

pyran furan 

to denote five- and six-membered rings, respectively, in cyclic sugars. The 
two forms of glucose are appropriately identified by the names a-D-gluco- 
pyranose and /?-D-glucopyranose. Likewise, L-arabinose, D-xylose, D-galactose, 
and D-mannose occur naturally as pyranoses — but D-ribose and D-fructose 
usually occur as furanoses. 



IS- 4 mutarotation 

Although the crystalline forms of a- and /?-D-glucose are quite stable, in 
solution each form slowly changes into an equilibrium mixture of both. The 
process can be readily observed as a decrease in specific optical rotation 
from that of the a anomer (+ 1 12°) or an increase from that of the fi anomer 
(+18.7°) to the equilibrium value of +52.5°. The phenomenon is known as 
mutarotation (Figure 15-2) and is commonly observed for reducing sugars 
(i.e., sugars with their carbonyl function in the form of a hemiacetal). Mutaro- 
tation is catalyzed by both acids and bases and by molecules that can change 
from one tautomeric form to another (Section 18-1E). 

At equilibrium, there is present some 64 % of the ji anomer and 36 % of 
the a anomer. The amount of the free aldehyde form present at equilibrium 
is very small (0.024%). Preponderance of the fi anomer is to be expected 
because the hydroxyl substituent at C-l is equatorial in the fS anomer [8] and 



Figure 15-2 Mutarotation of D-glucose in solution via the open-chain 
aldehyde form. 




°- H 



OH 



\c?> -=- -^y « 



H 



[6] [7] [8] 

a-D-glucose open-chain form /!-D-glucose 

(chair conformation) (one of many conformations) (chair conformation) 



sec 15.5 glycosides 407 

axial in the a anomer [6]. Oddly enough, compounds such as those described 
in the next section having a methoxy group at C-l tend to have the methoxy 
group axial (anomeric effect). 



15-5 glycosides 





Although abundant quantities of glucose and fructose occur in the free state 
in nature, they and the less common sugars are also found combined with 
various hydroxy compounds. The generic name for these combinations is 
glycoside, or, more specifically, O-glycoside, to denote that the linkage to the 
hydroxy compound is through oxygen. The simplest glycosides are those 
formed by the acid-catalyzed reaction of methanol with a molecule such as 
jS-D-glucose ; this reaction is shown in Equation 15-1. (Since the a and P forms 
of glucose are in equilibrium the a-glucoside will also be formed.) The cyclic 
sugar is a hemiacetal and the glycoside is an acetal. The additional group in 
the glycoside (methyl in Equation 15-1) is called the aglycone group. This 

H CH 2 OH ^ H CH2 0H 

H® ^ 

OH + CH 3 OH ► HO hoXJ-^-V OCH3 + H2 ° 



methyl ^-D-glucoside 

(plus a anomer) (15T) 

group is another sugar in many compounds found in nature, in which case the 
molecule is a disaccharide. Polysaccharides are formed by the union of many 
sugars by such linkages. 

Of particular importance biologically are the N-glycosides, in which the 
sugar residue is D-ribose and the aglycone is attached to the sugar by a C-N 
bond to a nitrogen base, usually a pyrimidine or purine derivative. These 
glycosides are better known as ribonucleosides and deoxyribonucleosides. 



N 

J 
N 

purine 



Furanose rings are usually shown by a Haworth projection formula because 
there is less need to indicate conformational mobility than with the six- 
membered pyranose rings. The latter are much like cyclohexane except that 
the ring is not puckered to quite the same extent. The furanose rings are much 
like cyclopentane, one of the atoms being slightly out of the plane of the 
other four. 

Adenosine is one example of a purine ribonucleoside, and when esterified 



HOH 2 C/°^N^ 


HOH 2 C/°\Xj^ 


rT^N 


,N. 


ANMJ/ 


IfajA 


V 


< 


i r 


T—T 


N 


N' 


OH OH 


OH H 




H 


ribonucleoside 


deoxyribonucleoside 


pyrimidine 


P 


(partial structure) 


(partial structure) 







chap IS carbohydrates 408 

at the 5' position with the triphosphoryl group (a phosphoric acid anhydride) 

OH OH OH 

I I I 

— P— O— P— O— P— OH 

II II II 

o o o 

it is known as adenosine triphosphate (ATP), a so-called "energy-rich" 
compound present in muscle tissue. 



5'-v 





adenosine adenosine triphosphate (ATP) 



15-6 disaccharides 

The simplest and most important oligosaccharides are the disaccharides. 
These, on acid or enzymic hydrolysis, give the component monosaccharides, 
which are frequently hexoses. The bond between the hexoses is an O-glycoside 
linkage, but only one hemiacetal hydroxyl need be involved. In fact, most 
disaccharides have reducing properties indicating that one of the sugar 
residues has the easily opened hemiacetal function. However, when both 
hexoses are joined through their anomeric carbons, as in sucrose, the sugar 
is an acetal (like a methyl glycoside) and has no reducing properties, and 
forms no phenylosazone or other carbonyl derivative (provided that the 
experimental conditions do not effect hydrolysis of the acetal function). 

Among the more important disaccharides are sucrose, maltose, cellobiose, 
and lactose (Figure 15-3). Sucrose and lactose occur widely as the free sugars, 
lactose in the milk of mammals and sucrose in plants, fruit, and honey 
(principally in sugar cane and sugar beet). Maltose is the product of enzymic 
hydrolysis of starch, and cellobiose is a product of hydrolysis of cellulose. 

To fully establish the structure of a disaccharide we must know (1) the 
identity of the component monosaccharides ; (2) the ring size (furanose or 
pyranose) in each monosaccharide, as it exists in the disaccharide: (3) the 
positions which link one monosaccharide with the other; and (4) the anomeric 
configuration (a or j6) of this linkage. 

The hydrolysis products of these four disaccharides are shown here. 



sec 15.6 disaccharides 409 



CH,OH 



HO HO 




H CH,OH 

H OH 



maltose 
[9] 



CH,OH 



HO HO 




'OH cellobiose 
[10] 



HO CH 2 OH x HO 

rO OHO/ ' 





CH 

I 
OH 



O 



OH lactose 
[11] 



CH,OH 



HO'ho 




O v 
HO 



CH,OH 



O 



■0> 
HO 




CH,OH 



sucrose 
[12] 



HO 



Figure 15-3 Conformational formulas for maltose [9], cellobiose [10], lactose 
[11], and sucrose [12]. For maltose, cellobiose, and lactose only one of the 
anomeric forms of the right-hand ring is shown. In the case of cellobiose and 
lactose the right-hand rings have been turned upside down to permit reason- 
able bond angles to be shown at the oxygens between the rings. 



C 12 H 22 11 + H 2 



H« 



- C 6 H 12 6 + C 6 H 12 6 



(or enzymes) 

maltose ► D-( + )-glucose + D-( + )-glucose 

cellobiose ► D-( + )-glucose + D-( + )-glucose 

lactose ► D-( + )-glucose + D-( + )-galactose 

sucrose ► D-( + )-glucose + d-( — )-fructose 



The hydrolysis products of maltose and cellobiose are identical — two mole- 
cules of glucose — because these two disaccharides differ only in the configura- 
tion at the anomeric linkage (see Figure 15-3). Although both disaccharides 
can be hydrolyzed with acid, they are much more discriminating in their 



chap IS carbohydrates 410 

behavior toward enzymes. Hydrolysis of maltose [9] and many other a-linked 
oligosaccharides is catalyzed by the action of an enzyme called maltase 
whereas hydrolysis of cellobiose [10] and other /Winked compounds is cata- 
lyzed by the enzyme emulsin. 

Sucrose [12] is a nonreducing nonmutarotating sugar because its compon- 
ent hexose units are joined at their anomeric carbons. Sucrose has a positive 
sign of rotation ([a] = +66.5°) as does D-glucose ([a] = +52°). D-Fructose, 
on the other hand, has a large negative rotation, [a] = — 92°. The complete 
hydrolysis of sucrose to glucose and fructose thus causes an inversion of the 
sign of rotation and the product is known as invert sugar. Honey is chiefly 
invert sugar because bees contain the enzyme invertase which catalyzes the 
hydrolysis. Sucrose has an extraordinarily high solubility in water — 1.5 g of 
sucrose will dissolve in 1 g of water at room temperature. 

The sign of rotation of naturally occurring glucose is revealed by one of the 
common names for this sugar : dextrose. 



15-7 polysaccharides 



The fibrous tissue in the cell walls of plants and trees contains the poly- 
saccharide cellulose, which consists of long chains of glucose units, each of 
which is combined by a jS-glucoside link to the C-4 hydroxyl of another 
glucose unit, as in the disaccharide cellobiose. 




Indeed, enzymic hydrolysis of cellulose leads to cellobiose. The molecular 
weight of cellulose varies with the source but is usually high; cotton cellulose, 
for example, consists of some 3000 glucose units per molecule. 

Cellulose is the natural fiber obtained from cotton, wood, flax, hemp, and 
jute and is used in the manufacture of textiles and paper. In addition to its 
use as a natural fiber, cellulose is used to make cellulose acetate (for making 
rayon acetate yarn, photographic film, and cellulose acetate plastics), cellulose 
nitrate (gun cotton and celluloid), and cellulose xanthate (for making viscose 
rayon fibers). The process by which viscose rayon is manufactured involves 
converting wood pulp or cotton linters into cellulose xanthate by treatment 
first with sodium hydroxide and then with carbon disulfide. 



-C-OH 



NaOH 



CS 2 



-C-0-C-S e N 

I 
cellulose xanthate 



The degree of polymerization of the original cellulose generally falls to 
around 300 monomer units in this process. At this degree of polymerization 



sec IS. 7 polysaccharides 411 

the cellulose is regenerated in the form of fine filaments by forcing the 
xanthate solution through a spinneret into an acid bath. 

S 

I II H ® I 

-C-O-C-S 6 H > -C-OH + CS 2 



Wood owes its strength to the presence of lignin, a highly cross-linked 
polymeric substance that binds the cellulose fibers together in a rigid mass. 
Cellulose is obtained from wood by pulping it with various chemicals that will 
dissolve the lignin and leave the cellulose fibers behind. Most woods are about 
50% cellulose and about 30% lignin, the remainder being other carbohy- 
drates, fats, and terpenoid resins (Section 29-3). 

Unlike cellulose there is no general agreement about the exact structure of 
lignin, possibly because it seems to lack the structural regularity of cellulose. 
There is no doubt, however, that a dioxyphenylpropyl moiety is an important 
part of the polymer. 

\ 
o 




-o^y-c-c-c 

a segment of lignin 

Lignin can be dissolved from wood in a number of ways but the method 
that least damages the cellulose fiber and can be applied to most varieties of 
wood involves digesting the wood with a solution of sodium hydroxide and 
sodium sulfide. This is called the kraft process because of the strength of the 
cellulose fibers that remain. (Kraft means strength in Swedish and in German.) 
The length of the cellulose fiber varies greatly from one wood to another and 
this is important in determining the strength of paper produced from it. 
Douglas fir has the longest fiber of the common woods — about 4 mm. 

The chemistry of lignin degradation that occurs during kraft pulping is not 
completely understood. A black liquor containing a wide variety of organo- 
sulfur compounds, some of them volatile and evil smelling, is formed. This 
liquor is burned to provide power and to enable the inorganic pulping chemi- 
cals to be recovered. In most other pulping methods, the waste liquor cannot 
be recycled. 

The principal alternative to kraft pulping involves dissolving the lignin 
through the action of sulfite in acid solution. The waste sulfite liquor that 
remains after the cellulose has been removed contains pulping chemicals, 
degraded lignin, and other substances. These were formerly discharged directly 
into rivers or inlets with drastic ecological effects. Largely as a result of anti- 
pollution legislation but partly to obtain useful organic products the waste 
sulfite liquors are now generally treated before being discharged. 

The per capita consumption of paper each year in North America is more 
than a quarter of a ton. The reader may wish to ponder the benefits and the 
drawbacks of this enormous rate of consumption. 

A few creatures (e.g., ruminants and termites) are able to metabolize 
cellulose with the aid of appropriate microorganisms in their intestinal tracts; 



chap IS carbohydrates 412 

but man cannot utilize cellulose as food because he lacks the necessary 
hydrolytic enzymes. However, such enzymes are widely distributed in nature 
and cause cellulosic materials, either textiles, paper, or wood, to deteriorate 
slowly. 

The second very widely distributed polysaccharide is starch, which is stored 
in the seeds, roots, and fibers of plants as a food reserve — a potential source 
of glucose. The number of glucose units in starch varies with the source, but 
in any one starch there "are two structurally different polysaccharides. Both 
consist entirely of glucose units, but one is a linear structure (amylose) and the 
other is a branched structure (amylopectin). 

The amylose form of starch consists of repeating 1,4-glucopyranose links 
as in cellulose, but unlike cellulose the linkage is a rather than /?. Hydrolysis 
by the enzyme diastase leads to maltose. 




O 




maltose unit 



o s 




o 




o' 



In amylopectin, amylose-like chains are apparently branched by 1,6 
linkages. 



O' 




O 





O 



O 

/ 
CH 




O* 




O' 



Animals also store glucose in the form of starchlike substances called 
glycogens. These resemble amylopectin more than amylose in that they are 
branched chains of glucose units with 3,4- and 1,6-glucoside links. 

A polysaccharide which resembles cellulose in structure and stability is 
chitin, which forms the hard shells of crustaceans and insects. Chitin is a 
polymer of N-acetyl-D-glucosamine; that is, the —OH groups at C-2 of the 

O 

II 
glucose units of cellulose are replaced by — NHCCH 3 groups. Crab shells 

boiled with hydrochloric acid produce D-glucosamine (2-amino-2-deoxy-D- 

glucose) [13]. 



sec 1S.9 immunologically important carbohydrates 413 




[13] 
D-glucosamine (/? anomer) 



15-8 vitamin C 

The " antiscorbutic " factor of fresh fruits which prevents the development 
of the typical symptoms of scurvy in man is a carbohydrate derivative known 
as vitamin C or ascorbic acid. This substance is not a carboxylic acid but a 
lactone and owes its acidic properties (and ease of oxidation) to the presence 
of an enediol grouping conjugate to a carbonyl group. It belongs to the l 

HO /° H 

I ,c=o 

H-CV 

HO— C— H 
I 
CH 2 OH 

L-ascorbic acid 

series by the glyceraldehyde convention. L-Ascorbic acid is formed from 
D-glucose in certain plants and in the liver of most animals (see Exercise 
15-10). 



15-9 immunologically important carbohydrates 

One of the most active areas of carbohydrate chemistry at the present is the 
study of the polysaccharides present in microorganisms and their influence 
on immunological specificity. The production of antibodies in the bloodstream 
when foreign bacteria are introduced is a well-known phenomenon, and the 
extraordinarily high degree of specificity of these antibodies can be traced to 
the chemical composition of that part of the cell wall of bacteria called the 
antigen. The antigen is a complex combination of polysaccharides, lipids 
(esters of fatty acids), and proteins. The antigen can be separated into these 
three fractions and the carbohydrate fraction degraded. 

The Salmonella bacteria, for example, have been intensively studied, and 
it has been shown that all of the 100 or more different cell- wall antigens that 
have been characterized contain five sugars: D-glucose; D-galactose; D-glucos- 
amine [13]; a C 7 sugar (a heptose); and a C 8 sugar (an octose). A polysac- 
charide of these five sugars makes the basal chain to which are attached side 



chap IS carbohydrates 414 

chains containing a number of other pentoses and hexoses. The way in which 
these sugar units are arranged is evidently the source of the specificity of the 
antigen-antibody interaction. 



summary 

Carbohydrates are polyhydroxyaldehydes and polyhydroxyketones — 
monosaccharides — or substances that yield these on hydrolysis — oligosac- 
charides (two to nine components) or polysaccharides (10 or more com- 
ponents). The monosaccharides are further classified according to the number 
of carbon atoms in the chain and to whether they are aldehydes or ketones ; 
that is, aldopentoses, ketohexoses, and so on. Most carbohydrates exist pre- 
dominantly in cyclic hemiacetal or hemiketal forms. 

There are four nonidentical asymmetric centers in the open-chain form of 
an aldohexose which means that there are 2 4 or 16 possible optical isomers 
(eight pairs of enantiomers), all of which are known. The most important of 
these is D-(+)-glucose (also called dextrose), whose Fischer projection for- 
mula is shown. 

CHO 

H-C-OH 
I 
HO-C— H 
I 
H-C-OH 

I 
H-C-OH 
I 
CH 2 OH 

D-( + )-glucose 

The relation of D-glucose to another of the aldohexoses (D-mannose) and 
one of the eight ketohexoses (D-fructose) was revealed by the fact that all three 
gave the same phenylosazone on treatment with phenylhydrazine. The con- 
figurations at C-3, C-4, and C-5 must be identical in all three compounds. 
The two aldohexoses are called epimers of one another since they differ only 
in configuration at C-2. 

CH=NNHC 6 H 5 

I 

C = NNHC 6 H 5 
I 
HO— C— H 

H— C— OH 

I 
H-C— OH 



phenylosazone from D-glucose, 
D-mannose, and D-fructose 

D-Glucose can be reduced to a hexahydric alcohol (sorbitol) or oxidized to 
a monocarboxylic acid (D-gluconic acid) or a dicarboxylic acid (D-glucaric 



summary 415 

acid). It reacts with methanol and acid to form a pair of acetal-like substances 
called methyl glucosides (general term — glycoside) whose structures are 
analogous to the two cyclic forms of D-glucose itself. The latter, designated 
a and /?, are anomers — they differ in configuration only at C-l. Two ways of 
representing the cyclic structure' of one of these anomers (a-D-glucose) and 
its methyl glucoside are shown here. 



CH 2 OH 


H CH 2 OH 


"Jr°^ 


HO \tT\ h 


Ho\° H I jj/0H 

H OH (CH3) 


Ho\i---V--V 

Th oh 1 

° H (CH 3 ) 


Haworth projection formula 


conformational formula 



The equilibrium between the open-chain form and the two cyclic anomers 
of D-glucose strongly favors the cyclic forms. The latter, being diastereomers 
of one another, can be separated by crystallization. When dissolved in 
water, each anomer reverts at a measurable rate to the equilibrium mixture. 
This interconversion process is known as mutarotation. 

The additional group in a glycoside (methyl in the formulas above) is called 
the aglycone group. Two of the important kinds of aglycones are nitrogen 
bases and other monosaccharides. The former leads to compounds such as 
the ribonucleosides — adenosine, for example — and the latter to di- and poly- 
saccharides. 

Four important disaccharides are sucrose (D-glucose and D-fructose joined 
at C-l and C-2, respectively), maltose and cellobiose (two D-glucose units 
joined at C-4 and C-l, respectively), and lactose. The two components of 
sucrose are bound together through their anomeric carbons and hence this 
sugar does not undergo mutarotation and is not oxidized by reagents such as 
Fehling's solution unless the disaccharide linkage is first hydrolyzed. The other 
three disaccharides have one free anomeric carbon and they can undergo 
mutarotation and reduce Fehling's solution. Maltose and cellobiose differ only 
in the configuration at the glycosidic link between the glucose units. Hydroly- 
sis of maltose and some other a-linked compounds is catalyzed by maltase, 
and hydrolysis of cellobiose and some other /Minked compounds is catalyzed 
by emulsin. 

Polysaccharides include cellulose and starch. Cellulose is a polymer of 
D-glucose with all the glycosidic links (S and is the fibrous material in cotton, 
wood, flax, and so on. It can be converted to various esters, some of which are 
useful in their own right (cellulose acetate) and from some of which a shorter- 
chain cellulose can be regenerated (viscose rayon). Starch is a mixture of 
polymers of D-glucose with all the glycosidic links a, one a linear polymer 
(amylose) and the other a branched-chain polymer (amylopectin). 

Vitamin C (ascorbic acid) is a carbohydrate derivative. Other important 
biological substances include the antigens, parts of which are polysaccharide 
in nature. Immunological response is related to the arrangement of monosac- 
charides in the polymer chain. 



chap IS carbohydrates 416 



exercises 



15-1 A naturally occurring optically active pentose (C 5 H 10 Os) reduces Tollen's 
reagent and forms a tetraacetate with acetic anhydride. It gives an optically 
inactive phenylosazone. Write all the possible structures for this pentose 
which are in accord with all the experimental facts. 

15-2 A hexose, C 6 H 12 6 , which we shall call X-ose, on reduction with sodium 
amalgam gives pure D-sorbitol, and, upon treatment with phenylhydrazine 
gives an osazone different from that of D-glucose. Write a projection formula 
for X-ose and equations for its reactions. 



15-3 D-Arabinose and D-ribose give the same phenylosazone. D-Ribose is reduced 
to the optically inactive pentahydric alcohol ribitol. D-Arabinose can be 
degraded by the Ruff method, which involves the following reactions : 



iCHO 

I 
2 CHOH 



Br 2 , H 2 Q , 



iC0 2 H 

I 

,CHOH 



Ca 2 ®. 



iCOf 
2 CHOH 



Ca 



M ^TTT* aCHO + C0 2 3 e 

rt2U2 I 



The tetrose, D-erythrose, so obtained can be oxidized with nitric acid to meso- 
tartaric acid. What are the configurations of D-arabinose, D-ribose, ribitol, 
and D-erythrose ? 

15-4 The logic necessary to solve this problem is essentially that used by Fischer in 
his classic work which established the configurations of glucose, arabinose, 
and mannose. 

a. Write projection formulas for all the theoretically possible D-aldopen- 
toses, HOH 2 C(CHOH) 3 CHO. 

b. One of the D-aldopentoses is the naturally occurring D-arabinose, 
enantiomeric with the more abundant L-arabinose. Oxidation of 
D-arabinose with nitric acid gives an optically active five-carbon 
trihydroxydicarboxylic acid. Which of the D-aldopentoses could be 
D-arabinose ? 

c. D-Arabinose is converted by the following transformations into D-glu- 



CHO 
I 
(CHOH) 3 

CH 2 OH 

D-arabinose 



NaCN 
pH9 



CN 

I 
CHOH 

I 
(CHOH) 3 



CH 2 OH 



o=c 

I 
CHOH 

I 
CHOH 

I 
CH 



O 



CHOH 
I 
CH 2 OH 

y-lactone 



H 2 Q, H®, 



Na-Hg 
pH3 



C0 2 H 
I 
CHOH 

(CHOH) 3 

CH 2 OH 



CHO 

I 
(CHOH) 4 



CH 2 OH 



D-glucose 

+ 
D-mannose 



-H,0. 



exercises 417 

cose and D-mannose. (This is the classic Kiliani-Fischer cyanohydrin 

synthesis of sugars). What do these transformations tell about the 
relationship between the configurations of mannose and glucose? 

d. Oxidation of D-glucose and D-mannose gives the six-carbon, tetrahy- 
droxydicarboxylic acids, glucaric and mannaric acids, respectively. 
Both are optically active. What then are the configurations of the d- 

/ and L-arabinoses ? 

e. D-Glucaric acid can form two different y-monolactones, whereas 
D-mannaric acid can form only one monolactone. What then are the 
configurations of D-glucose and D-mannose ? 

15-5 Deduce possible configurations of natural galactose from the following: 

a. The Wohl degradation is a means of reducing the chain length of a 
sugar by one carbon through the following reaction sequence: 

CHO CH=NOH CN 

I NH 2 OH I (CHjCO) 2 I (-HCN) 

CHOH * CHOH — ( _h 2 o) ' CHOH > CHO 

D-Galactose gives a pentose by one Wohl degradation. This pentose 
on nitric acid oxidation gives an optically active, five-carbon, trihy- 
droxydicarboxylic acid. 

b. The pentose by a second Wohl degradation followed by nitric acid 
oxidation gives D-tartaric acid. 

c. Write reasonable mechanisms for the reactions involved in the Wohl 
degradation. 

15-6 Draw Haworth- and conformation- type formulas for each of the following: 

a. methyl 2,3,4,6-O-tetramethyl-a-D-glucopyranoside 

b. a-L-arabinofuranose 

c. L-sucrose 

15-7 Sugars condense with anhydrous acetone in the presence of an acid catalyst 
to form isopropylidene derivatives. 



CH, 



I 



-C-OH / 3 HS -C-O /CH 3 

I + o=c — jf^ V 

"f-° H X CH 3 2 -A-C^CH, 

cis 

Predict the products of reaction of oc-D-galactopyranose, a-D-glucopyranose, 
and a-D-glucofuranose with acetone and an acid catalyst. 

15-8 How can the /3-D-glucoside units of cellulose produce a polymer with a 
stronger, more compact physical structure than the a-D-glucose units of 
starch ? Models will be helpful. 

15-9 Complete the following sequence of reactions, writing structures for all the 
products, A-I: 

acetone 

a. a-D-glucofuranose — ^j" > A (see Exercise 15-7) 



chap IS carbohydrates 418 



b. 


A 


c. 


B 


d. 


C+ D 


e. 


E+F 


/• 


G + H 



NalO* 

» 

Na 1A CN 

1. H 2 Q,0H 9 

2. H 9 , H 2 ' 

A 
"VH 2 0) ' 

NaBH 4 



H 2 



B 

C+D 

E + F + acetone 

G + H 

I + D-glucose-6- 14 C 



15-10 The following sequence of steps has been worked out for the conversion of 
D-glucose to L-ascorbic acid in nature. 



H CH,OH 



HO 




OH 



H CO,H 



^ HO 




O 

H H 

[14] 



HO 



HC v 




OH 



.OH 
~H 



115] 



HO- 



O 






II 


OH 




c 


HO 1 




/\ HH 






\ _J-L_ ^OH - 


1 c- 


=o 


'V^r^H 


H-CV 




H H 


HO— C— H 




[16] 


1 
CH,OH 





Formulas [14], [15], and [16] are drawn to show stereochemical interrela- 
tionships and not stable conformations. 

a. Identify the type of chemical reaction (reduction, lactone formation, 
etc.) that occurs in each of the steps. 

b. Account for the change from the d series to the l series that accom- 
panies the reaction. 

15-11 Ascorbic acid (see Exercise 15-10) has a p^Tha of 4.2, making it stronger than 
acetic acid, pK a x = 4.7. Identify the ionizable proton in ascorbic acid and 
account for the compound's acidity. 

15-12 A compound called inositol that is widely distributed in nature bears some 
resemblance to simple sugars. Its formula is 1,2,3,4,5,6-hexahydroxycy- 
clohexane and it can exist in nine possible stereoisomeric forms, two of 
which are optically active and seven of which are not. Draw the chair forms 
of the all-m and all-trans isomers of inositol. Which of these isomers will be 
the more stable? Will either of these isomers rotate the plane of polarized 
light? 



_ chapter 16 
organic nitrogen compounds 



chap 16 organic nitrogen compounds 421 

Nitrogen is an element with many oxidation levels, ranging from the most 
reduced state, NH 3 or NH 4 ®, to the most oxidized state, HN0 3 or N 2 5 . 
With a compound at an intermediate oxidation level, particularly one with 
nitrogen-nitrogen bonds, such as HN 3 , it is often not helpful to assign 
numerical oxidation states to the atoms. (The same situation occurs with 
compounds of carbon.) The important inorganic states of nitrogen are listed 
in Table 16T together with their organic analogs. The groupings there are 
made on the basis of bonding arrangements at the nitrogen atoms rather than 
on formal oxidation state. 



16-1 



amines 



A. TYPES AND NOMENCLATURE 

The nomenclature of alkyl-substituted ammonias, or amines, was considered 
briefly in Chapter 8. We shall give a short review here to help focus attention 
on the types of substitution which are commonly encountered. Classification 
is made according to the number of alkyl or aryl groups attached to nitrogen, 
since this number is important in determining the chemical reactions that are 
possible at the nitrogen atom. 



RNH 2 


R 2 NH 


R 3 N 


R 4 NX 


primary 


secondary 


tertiary 


quaternary 


amine 


amine 


amine 


ammonium salt 



Amino compounds can be named either as derivatives of ammonia or as 
amino-substituted compounds. Thus, HOCH 2 CH 2 NH 2 can be named almost 
equally well as 2-hydroxyethylamine or as 2-aminoethanol, although by con- 
vention 2-aminoethanol is favored, since hydroxyl normally takes precedence 
over amino. With halogens the situation is reversed, and 2-chloroethylamine 
is favored over 2-aminoethyl chloride. Some typical amines with the names 
which they have in common use are listed in Table 16-2 together with their 
physical properties. 

Salts of amines with inorganic or organic acids are usually best named as 
substituted ammonium salts. Often, however, the name of the corresponding 



CH3NH3 CI CH 2 =CH-CH 2 — N-CH 3 S 2 CCH 3 

CH 3 

methylammonium chloride allyldimethylammonium 

acetate 

amine is used in conjunction with the name of the acid, for instance, methyl- 
amine hydrochloride for methylammonium chloride. With quaternary salts 



chap 16 organic nitrogen compounds 422 
Table 16-1 Inorganic compounds of nitrogen and their organic analogs" 



inorganic compound 


methyl derivative 


acetyl derivative 


NH 3 


CH 3 NH 2 


CH 3 -C 

NH 2 


ammonia 


methylamine 






NH 4 ® Cl e 


CHN 3 H 3 S C1 9 


acetamide 


ammonium chloride 


methylammonium chloride 




NH 2 NH 2 


CH 3 NHNHCH 3 




hydrazine 


1 ,2-dimethylhydrazine 




HN=NH 


CH 3 N=NCH 3 




diimine 


azome thane 




(unstable) 






e e 




^° 


H-N=N=N 


CH 3 N 3 


CH 3 -C 


hydrazoic acid 


methyl azide 


N N 3 
acetyl azide 


HC=N 


CH 3 C=N 


CH 3 -C 


hydrogen cyanide 


acetonitrile 


acetyl cyanide 


N=COH ±> HN=C=0 


CH 3 N=C=0 




cyanic isocyanic 


methyl isocyanate 




acid acid 






NH 2 OH 


(CH 3 ) 2 NOH 


CH 3 — C 


hydroxylamine 


N,N-dimethylhydroxylamine 


NHOH 

N-acetylhydroxylamine 


a e 


© e 




0=N=N 


CH 2 =N=N 




nitrous oxide 


dizaomethane 




■N=0 






nitric oxide 






HON=0 


CH 3 -N=0 




nitrous acid 


nitrosomethane 
CH 3 0— N=0 
methyl nitrite 




■NO 2 ±5:N0 2 -NO 2 






nitrogen dinitrogen 






dioxide tetroxide 






HO-N0 2 


CH 3 -N0 2 




nitric acid 


nitrome thane 


P 




CH 3 0-N0 2 


CH 3 -C 




methyl nitrate 


ON0 2 

acetyl nitrate 



" Arranged according to the type of bonding at the nitrogen atoms ; the alkyl or acyl group 
replaces a hydrogen atom in the inorganic compound in some cases and a hydroxyl group or an 
oxygen atom in others. 



sec 16.1 amines 423 



Table 16-2 Typical amines and their properties 



amine 


name 


bp, 
°C 


mp, 

°C 


water solubility, 
g/100 ml, 25° 


in water" 


NH 3 


ammonia 


-33 


-77.7 


90 


1.8 x 10- 5 


CH 3 NH 2 


methylamine 


-6.5 


-92.5 


1156 


4.4 x 10" 4 


CH 3 CH 2 NH 2 


ethylamine 


16.6 


-80.6 


00 


5.6 x 10~ 4 


(CH 3 ) 3 CNH 2 


r-butylamine 6 


46 


-67.5 


CO 


2.8 x 10~ 4 


(CH 3 CH 2 ) 2 NH 


diethylamine 


55.5 


-50 


very soluble 


9.6 xlO- 4 


(CH 3 CH 2 ) 3 N 


triethylamine 


89.5 


-115 


1.5 


4.4 x 10" 4 


(CH 3 CH 2 CH 2 CH 2 ) 3 N 


tri-«-butylamine 


214 




slightly soluble 




O h 


piperidine 


106 


-9 


00 


1.6 x 10- 3 


o 


pyridine 


115 


-42 


CO 


1.7 x 10~ 9 


O nh2 


cyclohexylamine 


134 




slightly soluble 


4.4 x 10" 4 


/ Vnh 2 


aniline 


184.4 


-6.2 


3.4 


3.8 x 10- 10 


H 2 NCH 2 CH 2 NH 2 


ethylenediamine 


116 


8.5 


soluble 


8.5 x 10" 5 



' Usually at 20°-25°. 

' Note that f-butylamine is a primary amine. 



an extension of the same system leads to trimethylamine methiodide for 
tetramethylammonium iodide. Whenever possible, however, the substituted 
ammonium ion name should be used because a name such as trimethylamine 
methiodide conceals the fact that the bonds to the four methyl groups are 
identical. 



B. PHYSICAL AND SPECTROSCOPIC PROPERTIES OF AMINES 



The properties of amines depend in an important way on the degree of substi- 
tution at nitrogen. For example, tertiary amines have no N— H bonds and 
are thus unable to form hydrogen bonds of the type N— H---N. In general, 
N— H— N bonds are somewhat weaker than those of corresponding types, 
O— H— O and F— H---F, because the electronegativity of nitrogen is less than 
that of oxygen or fluorine. Even so, association of the molecules of primary 
and secondary amines (but not tertiary amines) through hydrogen bonding 
is significant and decreases their volatility relative to hydrocarbons of similar 
size, weight, and shape, as the accompanying examples show. 



chap 16 organic nitrogen compounds 424 




SS 



nr 1 



m 




wavelength, jJL 

j T 



-i 1 rr 



ch.nh : 




. \ II stretch 



(.'. N siix'tcli 



3600 3200 2800 2400 2000 2000 



Jtfclll-L 




frequency, cm 



Figure 16-1 Infrared spectra of cyclohexylamine and N-methylaniline. 



CH 3 CH 2 CH 2 CH 2 CH 2 NH 2 

rt-pentylamine 
mol. wt. 87; bp 130° 

CH 3 CH 2 NHCH 2 CH 3 

diethylamine 

mol. wt. 73; bp 55.5° 

CH 2 CH 3 
I 
CH 3 CH 2 -N— CH 2 CH 3 

triethylamine 
mol. wt. 101 ; bp 89.5° 



CH 3 CH 2 CH 2 CH 2 CH 2 CH 




hexane 


mol 


wt. 86; bp 69° 


CH 3 CH 2 CH 2 CH 2 CH 3 




pentane 


mol. 


wt. 72; bp 36° 




CH 2 CH 3 



CH 3 CH 2 CHCH 2 CH 3 

3-ethylpentane 
mol. wt. 100; bp 93.3° 



sec 16.1 amines 425 



I 
300 



200 



100 



Hz 



OH -N 

V 



V 



CH-C 



■I V N ~" 



_jy i- pj J iv A ty 



i 

5.0 



: L 



_Lj 



4.0 



3.0 



2.0 



1.0 



ppm 



Figure 16-2 Nuclear magnetic resonance spectrum of diethylamine at 60 
MHz relative to tetramethylsilane, 0.00 ppm. 



The water solubilities of the lower-molecular-weight amines are generally 
greater than those of alcohols of comparable molecular weights. This is the 
result of hydrogen bonding between the amine and water, which leads to 

I 
hydrogen bonds (of considerable strength) of the type — N:— H— O— H. 

I 
A characteristic feature of the infrared spectra of primary and secondary 
amines is the moderately weak absorption at 3500 to 3300 cm -1 , correspond- 
ing to N— H stretching vibrations. Primary amines have two such bands in 
this region, whereas secondary amines generally show only one band. Absorp- 
tion is shifted to lower frequencies on hydrogen bonding of the amine, but 
because NH---N hydrogen bonding is weaker than OH— O hydrogen bonding, 
the shift is not as great and the bands are not as intense as are the absorption 
bands of hydrogen-bonded O— H groups (see Table 7-1). Absorptions corre- 
sponding to'G—N vibrations are less easily identifiable, except in the case of 
aromatic amines, which absorb fairly strongly near 1300 cm -1 . Spectra that 
illustrate these effects are shown in Figure 16-1. 

I 
The nmr spectra of amines show characteristic absorptions for H— C— N 

I 
protons around 2.7 ppm. The resonances of N— H protons are not so easily 
identifiable; considerable variability arises from differences in degree of hydro- 
gen bonding and matters are further complicated when, as with diethylamine 
(Figure 16-2), the N— H resonance has nearly the same chemical shift as the 
resonances of C— CH 3 protons. 



C. STEREOCHEMISTRY OF AMINES 



The bond angles of nitrogen in amines and ammonia are less than tetrahedral 
as a consequence of electrostatic repulsion between the bonding electrons and 



chap 16 organic nitrogen compounds 426 

the unshared electron pair (Section 1-2). The configuration at nitrogen is 
pyramidal and, if three different groups are attached, optically active forms 
are, in principle, possible because there is no plane or center of symmetry in 
the molecules. 



V 



~Ri 



Ri 



,1SL 



-R2 



The resolution of such a mixture of forms has not as yet been achieved 
because they are rapidly interconverted by an inversion process involving a 
planar transition state. 



\ 



— R, 



,-R, 



R2 N- 



Ri 



planar transition 
state 



With ammonia, inversion of this type occurs at about 4 x 10 10 times per 
second at room temperature; with aliphatic tertiary amines, the rate is of the 
order of 10 3 to 10 5 times per second. Such inversion rates are much too great 
to permit separation of amines into stable optical isomers at room tempera- 
ture. There are a few specially substituted amines such as [1] which invert at 
nitrogen sufficiently slowly at room temperature to permit isolation of the 
separate diastereomers [la] and [lb]. 



CI / 



CH, 



\ 



N H 
[la] 



CH, 

CH 2 — C 

N H 
/ 
CI 

[lb] 



D. AMINES AS ACIDS AND BASES 

Although perhaps the most characteristic chemical property of amines is their 
ability to act as bases by accepting protons from a variety of acids, it should 
not be forgotten that primary and secondary amines are also able to act as 
acids, albeit very weak acids. The lithium salts of such amines are readily 
preparable in ether solution by treatment of the amine with phenyllithium : 



ether e ® 
> (C 2 H 5 ) 2 N Li + C 6 H 6 



lithium diethylamide 



sec 16.1 amines 427 

The anions from aliphatic amines, being conjugate bases of very weak acids 
(K HA ~ 10" 33 ), are powerfully basic reagents and will remove protons even 
from many notably weak carbon acids. 

The base strengths of saturated aliphatic amines, as given by K B , are 
usually about 1CT 4 in water solution. 

K B © e 

RNH 2 + H 2 , RNHj + OH 

Ammonia and primary, secondary, and tertiary amines have the same base 
strengths within perhaps a factor of 50, as can be seen from the data in 
Table 16-2. There is no evidence for formation of undissociated amine hy- 
droxides such as R3NHOH, any more than there is for NH 4 OH. 

Unsaturated amines are often very weak bases. An example with C— N 
unsaturation is pyridine, C 5 H 5 N, which is a nitrogen analog of benzene, often 
called a heterocycle because not all of the atoms in the ring are of the same 
element. Pyridine, K B = 1.7 x 10" 9 , is less basic by a factor of about 10 5 than 
aliphatic amines. 



pyridine 

Vinylamines or enamines, RCH=CH— NH 2 , are not usually stable and 
rearrange to imines, RCH 2 CH=NH. An exception of a particular sort is 
aniline (phenylamine, C 6 H 5 NH 2 ), which has an amino group attached to a 
benzene ring. Here, the imine structure is less favorable because of the con- 
siderable stabilization energy of the aromatic ring: 

H 

Q-NH 2 ^_ Q=NH 

The K B of aniline is 10" 10 , which is less by a factor of 10 6 than that of 
cyclohexylamine (and is even less than that of pyridine). The decrease in 
stabilization associated with forming a bond to the unshared electron pair 
of nitrogen accounts for the difference ; it prevents the electron pair from being 
delocalized over the benzene ring as represented by the following structures : 



<^S-NH 2 < > ^\=NH 2 < ► ?/ V=NH 2 < ► / \=NH 2 



Some or all of the low basicity of aniline may also be accounted for by the 
electron-attracting power of unsaturated carbons relative to saturated car- 
bons (see Section 5-5). 



chap 16 organic nitrogen compounds 428 

E. THE PREPARATION OF AMINES 

1 . Alkylation. We discussed the reactions of ammonia and amines with 
alkyl halides in Chapter 8. Such processes would be expected to provide 
straightforward syntheses of amines, at least with those halides which under- 
go S N 2 but not E2 reactions readily. For example, 

CH3I + NH3 ► CH3NH3I 

The methylamine can be recovered by treating the salt with a strong base, 
such as sodium hydroxide. 

e e e e 
CH 3 NH 3 I+NaOH 



In practice, such reactions lead to mixtures of products because of equilibria 
of the following kind and subsequent reactions to give more than mono- 
alkylation. 



CH3NH3I + NH3 ; ' CH 3 NH 2 + NH 4 I 

CH 3 NH 2 + CH 3 I ► (CH 3 ) 2 NH 2 I 

The reaction continues and ultimately gives some tetramethylammonium 
iodide. 

(CH 3 ) 3 N + CH 3 I ► (CH 3 ) 4 NI 

The latter product, of course, is not readily converted to an amine by treat- 
ment with base. 

Nevertheless, the alkylation reaction is by no means hopeless as a practical 
method for the preparation of amines because usually the starting materials 
are readily available and the boiling-point differences between mono-, di-, and 
trialkylamines are sufficiently large to make for easy separations by fractional 
distillation. Separations may also be achieved by chemical means. 

The tetraalkylammonium halide salts formed by exhaustive alkylation of 
amines resemble alkali salts and, with moist silver oxide, can be converted to 
tetraalkylammonium hydroxides. These compounds are strong bases and are 
true ammonium hydroxides. 

(CH 3 ) 4 N e I e Ag2 ° ,H2 °. (CH 3 ) 4 N ffi OH e + Agl(s) 

Tetramethylammonium hydroxide, when heated, decomposes slowly accord- 
ing to the following equation : 

(CH 3 ) 4 N OH > CH3OH + (CH 3 ) 3 N 

With a higher alkylammonium hydroxide, thermal decomposition leads to the 
formation of an alkene. The reaction is a standard method for the prepara- 
tion of alkenes. 



CH 2 N(CH 3 ) 3 OH ( Hz0) > < >=CH 2 + N(CH 3 ) 3 





sec 16.1 amines 429 

Many of the- drugs that have been used in cancer therapy are chloroalkyl- 
amines whose effectiveness appears to be associated with their alkylating 
ability. The so-called nitrogen mustards [2], containing two chloroethyl 
groups, react by a sequence of displacement steps. The chlorine is displaced 

CH 2 CH 2 C1 CH 2 — CH 2 CH 2 CH 2 Z 

/ \ffi/ ZH / 

R— N ► N — > R— N 

K \ (~Cl e ) / \ (-H®) K \ 

CH 2 CH 2 C1 R CH 2 CH 2 C1 CH 2 CH 2 C1 

[2] 

by the neighboring amino group and the resulting ion rapidly alkylates an 
amino or hydroxy lie compound (ZH). Alkylating agents react with many 
cellular constituents but there is evidence to suggest that reaction with 
deoxyribonucleic acid (DNA; Section 17-6) is the biologically important 
process. 

The name nitrogen mustards arises from the structural resemblance of 
these compounds to the toxic agent mustard gas S(CH 2 CH 2 C1) 2 , and some of 
them share its vesicant (blistering) action on the skin. The nitrogen mustards 
contain two chloroethyl groups and both can undergo displacement, each 
reaction being aided by the neighboring amino group. It is significant that 
almost all the alkylating agents that exhibit tumor-inhibiting properties con- 
tain two or more alkyl groups, suggesting that some form of cross-linking 
process takes place within the cell. 

Some of the more successful chemotherapeutic agents are chlorambucil [3], 
used extensively in treating chronic lymphocytic leukemia, and merophan [4]. 

CH,CH 2 C1 
I 

N— CH 2 CH 2 C1 
C,CH 2 CH 2 ^ = ^ /=( 

N-^ J>-CH 2 CH 2 CH 2 C0 2 H <( ^-CH 2 CHC0 2 H 

C1CH 2 CH 2 NH 2 

[3] [4] 

Merophan has been used successfully in the treatment of Burkitt's lymphoma, 
a disease which mainly affects children in parts of Africa and which is often 
curable by chemotherapy alone. 

2. The Beckmann and Related Rearrangements. Rearrangement of oximes of 
ketones (Table 1 1 -4) by the action of concentrated sulfuric acid often pro- 
vides a useful synthesis of amines through formation of intermediates with 
positive nitrogen called nitrenium ions. 1 This reaction is called the Beckmann 
rearrangement. The nitrenium ion has only a fleeting existence at best and, 
indeed, may not be a true intermediate at all. The departure of the water 
molecule from the protonated oxime is probably accompanied by the 1 ,2-shift 



1 A more appropriate name for such an ion might be nitronium ion, analogous to car- 
bonium ion, but this term is already in use for the species N0 2 ®. 



chap 16 organic nitrogen compounds 430 

of the R group from the carbon to nitrogen. The amide is not usually hy dro- 



ll OH R OH 2 

\ / H,S0 4 \ / 

C=N > C=N 

/ / 

R R 



-H 2 



R— CO,H + RNH, 



H 2 Q 

© e 
H(OH) 



o 

II 

R— C-NH— R * 



H 2 



R 



R 



\ e 
C=N: 



a nitrenium 
ion 



R-C=N-R 



lyzed under the conditions of the reaction. If the amine is desired, a separate 
hydrolysis step has to be carried out. The Beckmann rearrangement is often 
conducted using phosphorus pentachloride instead of sulfuric acid. 

Three other reactions that proceed by way of neutral intermediates possess- 
ing electron-deficient nitrogen and also give amines are the Schmidt, Curtius, 
and Hofmann reactions, shown in that order (the influence of late nineteenth 
century German chemists is seen in the number of reactions that bear their 
names). 



R-C 



O 



OH 
O 



R— C 

\ 



HN 3 



H 2 SO* 



NHNH 2 
O 



R— C + Br 2 + NaOH 




HNO 



\ 



NH, 



♦ N 2 + 



O 



R-C 



N: 



-♦ R— N=C=0 



H 2 
RNH 2 + C0 2 



3. Formation of Amines by Reduction. Excellent procedures are available 
for preparation of primary, secondary, and tertiary amines by the reduction 
of a variety of types of nitrogen compounds. Primary amines can be obtained 
by hydrogenation or lithium aluminum hydride reduction of nitro compounds, 
azides, oximes, nitriles, and unsubstituted amides. 



RCH=NOH 
RC=N 

O 

// 
RC 
\ 
NH, 



LiAlH* 



(or cat. H 2 ) 



* RCH,NH 2 



sec 16.1 amines 431 

Some care has to be exercised in the reduction of nitro compounds since 
reduction is highly exothermic. For example, the reaction of 1 mole (61 g) 
of nitromethane with hydrogen to give methylamine liberates sufficient heat 
to raise the temperature of a 25-lb iron bomb by 100°. 

► CH 3 NH, +2H,0 Aff=-85kcal 



Secondary and tertiary amines, particularly those with different R groups, 
are advantageously prepared by lithium aluminum hydride reduction of 
substituted amides (Section 13-8). 



F. REACTIONS OF AMINES 

1. With Acids. The formation of salts from amines and acids is the most 
characteristic reaction of amines, and, since amines are usually soluble in 
organic solvents, amines are often very useful when a mild base is required for 
a base-catalyzed reaction or when it is desirable to tie up an acidic reaction 
product. Pyridine has excellent properties in this regard, being a tertiary amine 
with a K B of about 10~ 9 , reasonably volatile (bp 115°), and soluble in both 
water and hydrocarbons. When a stronger base is required, triethylamine is 
commonly used. The strengths of various amines as bases were considered in 
Section 16- ID. 

2. Acylation of Amines. The unshared electrons on nitrogen play a key role 
in many other reactions of amines besides salt formation. In fact, almost all 
reactions of amines at the nitrogen atom have, as a first step, the formation of 
a bond involving the unshared electron pair on nitrogen. A typical example is 
acylation, wherein an amide is formed by the reaction of an acid chloride, an 
anhydride, or an ester with an amine (Section 13-8). The initial step in these 
reactions is as follows, using benzoyl derivatives and methylamine as illus- 
trative reactants. 

e 
=± C 6 H 5 -C-X 



H,N-CH 



O 

II 
X = halogen, -0-C-C 6 H 5 or -OR 

The reaction is completed by loss of a proton and elimination of X e . 

r o e o 

1 -H® H P --X e II 
: 6 H 5 -C-X ^ C 6 H 5 -C-X ^ - C 6 H 5 -C-NHCH 3 

H 2 N— CH 3 HN— CH, 

A serious disadvantage to the preparation of amides through the reaction 



chap 16 organic nitrogen compounds 432 

of an amine with an acid chloride (or anhydride) is the formation of 1 mole of 
amine salt for each mole of amide. This is especially serious if the amine is the 

O O 

II II H e e 

CH 3 -C— CI + 2 RNH 2 ► CH 3 — C— N— R + RNH 3 CI 

expensive ingredient in the reaction. In such circumstances, the reaction is 
usually carried on in a two-phase system with the acid chloride and amine in 
the nonaqueous phase and sodium hydroxide in the aqueous phase. As the 
amine salt is formed and dissolves in the water, it is converted to amine by the 
sodium hydroxide and extracted back into the nonaqueous phase. This 

© e © e ® e 

RNH 3 Cl + NaOH ► RNH 2 + Na CI + H 2 

procedure requires an excess of acid chloride, since some of it is wasted by 
hydrolysis. 

3. Amines with Nitrous Acid. Nitrous acid reacts with all amines, but the 
nature of the products depends very much on whether the amine is primary, 
secondary, or tertiary. Indeed, aqueous nitrous acid is a useful reagent to 
distinguish these compounds. Thus, with primary amines, nitrous acid 
evolves nitrogen gas; with secondary amines, an insoluble yellow liquid or 
solid N-nitroso compound (R 2 N— N=0) separates; tertiary amines dis- 
solve in and react with aqueous nitrous acid without evolution of nitrogen, 
usually to give complex products. These reactions are of little preparative 
value with many aliphatic amines because mixtures of products result. With 
aromatic amines such as aniline, however, the reaction with nitrous acid is 
extremely useful (Section 22-6). 

4. Halogenation. Primary amines in the presence of base react with hypo- 
chlorous acid [or f-butyl hypochlorite, (CH 3 ) 3 C— OC1] to produce both 
mono- and disubstitution products : 

H CI 

DXIU C ' 2 ' R-N-Cl —^ R-N-Cl 



2 OH e OH 6 

Secondary amines give mono-N-haloamines. These substances are rather 
weak bases, K B ~ 10" 13 ; they are not very stable and are oxidizing agents. 
They hydrolyze in water, particularly in acid solution, to regenerate the 
halogen. 

R 2 NC1 + HC1 ► R 2 NH + C1 2 

5. Oxidation of Amines. Oxidation with permanganate. Potassium per- 
manganate and most other strong oxidants oxidize amines in much the same 
way that they oxidize alcohols, giving aldehydes, ketones, or carboxylic 
acids. It should be remembered that the terms primary, secondary, and 
tertiary mean different things when applied to alcohols and amines. A 



sec 16.1 amines 433 

secondary alcohol, for example, is an alcohol with one hydrogen atom on the 
alcoholic carbon atom (R 2 CHOH); a secondary amine is an amine with one 
hydrogen on the nitrogen atom, R 2 NH. We should focus our attention on the 
alkyl groups in the two cases; a primary alkyl group, whether attached to 
—OH, — NH 2 , — NHR, or — NR 2 , will give an aldehyde or carboxylic acid 
on oxidation: 



[O] 



-► R-C 



primary alcohol RCH 2 OH 

primary amine RCH 2 NH 2 

secondary amine RCH 2 NHCH 3 

tertiary amine RCH 2 N(CH 3 ) 2 

A secondary alkyl group will give a ketone : 

secondary alcohol R 2 CHOH 
primary amine R 2 CHNH 2 

secondary amine R2CHNHCH3 

tertiary amine R 2 CHN(CH 3 ) 



[O] 



■* R-C 



O 



OH 



[O] 



+ R 2 C=0 



The amino nitrogen in the above reactions is converted largely to ammonia. 
A tertiary alcohol is not oxidized unless the conditions are drastic enough 
to cleave a carbon-carbon bond; a tertiary alkyl group in an amine is likewise 
stable. In this case, however, the amino group succumbs slowly and it is often 
possible to obtain high yields of the nitro compound : 



(CH 3 ) 3 C-NH 2 
r-butylamine 



KMnd 



(CH 3 ) 3 C-N0 2 

2-nitro-2-methylpropane 
83% 



A compound such as (CH 3 ) 3 C— N(CH 3 ) 2 will be demethylated by per- 
manganate by rapid oxidation of the two methyl groups, producing even- 
tually two moles of C0 2 . 

Acid helps protect amines from oxidation by tying up the pair of electrons 
on nitrogen and this is a matter of some importance because many common 
oxidizing agents, such as Cr VI , are vigorous oxidants only in acid solution. 
Permanganate oxidations, however, do not require acid catalysis and amines 
are readily oxidized by this reagent in basic solution. 

Oxidation with peracids. Oxidation of tertiary amines by peracids gives 
amine oxides. Thus, triethylamine can be oxidized with peracetic acid to 
triethylamine oxide: 



CH,CH 2 



\ 

— f 
/ 



O 



\ 



CH 3 CH 2 o 

\e e / 

-» CH 3 CH 2 -N-0 + CH 3 C 
/ \ 

CHXH, OH 



CH 3 CH 2 O-OH 

Peracids are excellent donors of oxygen atoms to amines (and alsotoalkenes; 



chap 16 organic nitrogen compounds 434 

see Section 4-4G) because bond breaking in the transition state is accompanied 
by bond formation, and high-energy intermediates are thus avoided: 



Ah / 

-c' | + :N- 



A-H 



(hydrogen bonded) , transition state 



OH 

/ B / 

R-C + e O-N- 

xX o 

With secondary amines, peracids give hydroxylamines, presumably as the 
result of rearrangement of the N-oxide : 

r 2 NH > R 2 NH-0 ► R 2 NOH 

Primary amines react similarly, but with an excess of peracid they go all 
the way to the nitro compound, though the yields are only moderate. 

Unlike amines, amine oxides would not be expected to undergo inversion 
at the nitrogen atom, and the oxides from amines with three different R 
groups should be resolvable into optically active forms. This has been realized 
for several amine oxides, including the one from methylethylallylamine. 

Amine oxides are one of the few types of organic nitrogen compounds that 

do not have analogs in known inorganic nitrogen compounds. The inorganic 

e e 
model would be H 3 N— O, a surely unstable tautomer of hydroxy lamine, 
H 2 NOH. 



16-2 amides 

An understanding of the chemistry of simple amides is particularly important 
because peptides and proteins, substances that are fundamental to all life 
as we know it, are polyamides. 

o o 

// // 

-c -c 

\ \ 

NH 2 NR 2 

simple amide N,N-disubstituted amide 

An important characteristic of the amide group is its planarity — the carbon, 
oxygen, nitrogen, and the first atoms of the R groups of an N,N-disubstituted 
amide all lie in the same plane. This planarity can be regarded as reflecting the 
importance of the following type of resonance in the amide group : 



sec 16.2 amides 435 

o o e 

/ / 

R-C „ < ► R-C 

\ ,R \©/R 

N N 

II 
R R 

Coplanarity is required if the dipolar structure is to be significant. The 
stabilization energy of acetamide is 1 1 kcal/mole, a value typical of most 
amides. 



A. PHYSICAL AND SPECTRAL CHARACTERISTICS OF AMIDES 

The amides have rather high melting and boiling points as a result of extensive 
intermolecular hydrogen bonding. Formamide is the only simple amide that 
is liquid at room temperature. Acetamide (mp 82°, bp 222°) has melting and 
boiling points far above the values for acetic acid (mp 17° and bp 118°), which 
has almost the same molecular weight. Disubstituted amides such as N,N- 
dimethylacetamide have lower boiling and melting points because of the 
absence of hydrogen bonding. 



O o O 

J // // 

H-C CH 3 -C CHj-C 

NH 2 NH 2 N(CH 3 ) 2 

formamide acetamide N,N-dimethylacetamide 

(methanamide) (ethanamide) (N,N-dimethylethanamide) 

mp 2° mp 82° mp -20° 

bp 193° bp222° bp 165" 



Considerable information is available on the infrared spectra of amides. By 
way of example, the spectra of three typical amides with different degrees of 
substitution on nitrogen are shown in Figure 16-3. 

As we might expect, a strong carbonyl absorption is evident in the spectra 
of all amides, although the frequency of absorption varies slightly with the 
structure of the amide. Thus, unsubstituted amides generally absorb near 
1690 cm" 1 , whereas mono-N-substituted and di-N-substituted amides 
absorb at slightly lower frequencies. The N— H stretching frequencies of 
amides are close to those of amines and show shifts of 100 to 200 cm -1 to 
lower frequencies as the result of hydrogen bonding. Unsubstituted amides 
have two N — H bands of medium intensity near 3500 and 3400 cm" 1 , 
whereas monosubstituted amides, to a first approximation, have only one 
N— H band near 3440 cm -1 . 

The nmr resonance of the N — H protons of amides are usually somewhat 
different from any we have discussed so far. Customarily, these will appear as 
a quite broad ragged singlet absorption which may turn to a broad triplet at 
high temperatures. A typical example is shown by propanamide (Figure 
16-4). The reason for this is associated with the special nuclear properties of 
14 N, the abundant isotope of nitrogen. When the 14 N is replaced by 15 N, the 



chap 16 organic nitrogen compounds 436 



-~\ 



* (II CH 4 C— NH 2 




wavelength, jX 

6 7 



Y|[^ h\clnii;i-n- 
T Ixmcl.il N— H 



1 1 i i i ,l ; 

3600 3200 2800 1400 2000 2000 




1800 1600 

frequency, cm 



_1 i lJ 

1400 1200 1000 800 







wavelength, /X 










3 4 5 


5 6 7 


8 
p- 

(Y 
1 1 


9 

~ I — 

v 


10 12 14 




1 1 1 

1 I 


1 1 1 

~Vvv 


If 

H!/ ' 

iji/ 


a 
o 

S 

= 

a i 


I r ° 
J ' CH.,C— N 

/■v. H 

/ ^ I..TC 


■hi 

*=/ Ml 
yen- bonded 


A 
II A 

. ■ ! 
/ li i ' 

1 UK 
1 II I 


HI 

1 ' 


I 




free N-H 


nmMmHH 

=() slM-tdi — n ! Ij 


1 'I 
1 ij 

I 










i i i i i 


1 , - , L .J/ f -^ B j 


1 


( 






3600 3200 2800 2400 2000 


2000 1800 1600 , 1400 

frequency, cm 


1200 




1 000 800 







wavelength, jU 








3 


4 5 


5 6 7. 


8 9 


10 


12 14 


w 

' 1 

1 1 

1 j 

1 1 


gMMMfef 

lllllllilip 


1 1 1 
~**~\1 " — V 

V / 1 M 

\ A 


i ,i 

if- 1 


y^~ 


lliiliiilllllll 

i 


1 

1 11 


ii- 


O GH, A :' 

II / hi 

CH. . • 1 

Mi 


if* 1 ' i 
II i 

11 jl 


f 




1 

1 

1 




. "■ \ L 


a 








(. <) stretch '*— ***!"| ] .■;■• . 1 


. 1 






1 1 1 


• 


W * f .' U t - 4 •' t • ( 


t , 


^i L ' « - 




3600 3200 2800 


2400 2000 


2000 1800 1600 _, 1400 

frequency, cm 


1200 


1000 


800 



sec 16.2 amides 437 



<- Figure 16-3 Infrared spectra of propanamide, acetanilide, and N,N- 
dimethylformamide in chloroform solution. Note the appearance of both free 
NH bands (sharp, 3300-3S00 cm" 1 ) and hydrogen-bonded N-H bands 
(broad, 3100-3300 cm -1 ) for unsubstituted and monosubstituted amides. 



lines become completely sharp. A more detailed explanation of this behavior 
is beyond the scope of this book. 

The ultraviolet spectra of amides generally resemble those of carboxylic 
acids. 

In general, the amide group is reasonably polar and the lower-molecular- 
weight amides are reasonably high melting and water soluble as compared to 
esters, amines, alcohols, and the like. N,N-Dimethylformamide and N- 
methylpyrrolidone have excellent solvent properties for both polar and non- 
polar substances. 

O / CH 3 



H-C— N(CH 3 ) 



3'2 



a 



N, N-dimethylformamide N-methylpyrrolidone 

(N-methyl-y-butyrolactam) 

Amides with N— H bonds are weakly acidic, the usual K HA being about 



10 



-16 



CH 3 -C 

\ 
NH, 



P 

NH NHJ 



Amides, then, are far more acidic than ammonia with K HA ~ 10 33 , and 
this reflects a very substantial degree of stabilization of the amide anion. 



Figure 16-4 Nuclear magnetic resonance spectrum of propanamide, CH 3 CH 2 - 
CONH 2 , in chloroform solution (solvent not shown) at 60 MHz relative to 
TMS at 0.00 ppm. 




ppm 



chap 16 organic nitrogen compounds 438 

However, amides are still very weak acids and, for practical purposes, are to 
be regarded as essentially nonacidic in aqueous solutions. 

The degree of basicity of amides is very much less than that of aliphatic 
amines. For acetamide, K B is about 10~ 15 (the K BHS of the conjugate acid is 
~10) and acetamide is " half-protonated " in 1 M sulfuric acid. 



NH, 



H,0 



o 

// 

CH 3 — C 

\e 
NH 3 


e 
+ OH 


OH 
/ 
CH,— C 


e 
+ OH 



NH 2 

The proton can become attached either to nitrogen or to oxygen. Nitrogen 
is, of course, intrinsically more basic than oxygen, but formation of the 
N-conjugate acid would cause loss of all the resonance stabilization energy of 
the amide. Infrared and nmr studies of amide cations reveal that the O- 
protonated form predominates; the equilibrium between the two is established 
so rapidly, however, that the JV-protonated form appears to serve as an inter- 
mediate in some reactions. 

B. PREPARATION OF AMIDES 

Two of the three reactions given below have been discussed in some detail 
earlier. The third, hydrolysis of nitriles, has been met previously only as a 
means of preparing carboxylic acids. 

1. Reaction of esters, acyl chlorides, or anhydrides with ammonia or 
amines (Section 13-8). 

p o 

R-C^ + NH 3 > R-C + EtOH 

OEt NH 2 

P ° 

R-C + 2NH 3 > R-C + NH?Cl e ' 

CI NH 2 

P 

P 

O + 2NH 3 ► R-C 

/ \ 

R ~ C XX NH 2 

O 

If primary or secondary amines are used in place of ammonia, N-substituted 
amides result. 



p o 

CH 3 -C + 2CH 3 NH 2 CH 3 -C + CH 3 NHf CI e 

CI NHCH 3 

N-methylacetamide 



sec 16.2 amides 439 

2. Beckmann rearrangement of oximes (Section 16TE2). 

R2 C=NOH ^^ RC^R -^ R-/ 

\ 
NHR 

N-Substituted amides are formed ; hydrolysis beyond the amide stage gives an 
amine and a carboxylic acid. 

3. Hydrolysis of nitriles. 

O 
dhs // 

R-C=N + H 2 2 > R-C + J0 2 

NH 2 

Nitriles can be hydrolyzed to amides in strongly acidic or basic solution but a 
better method is to use hydrogen peroxide in mildly basic solution, thus 
avoiding further hydrolysis of the amide to the carboxylic acid. The effective- 
ness of hydrogen peroxide can in part be traced to the very high nucleophilicity 
of the hydroperoxide anion H0 2 e , compared to the hydroxide ion itself 
(see Exercise 16-24). 



C. REACTIONS OF AMIDES 

In general, amides can be hydrolyzed in either acid or base solutions although 
the reactions are usually slow. The mechanisms are much like that of ester 
hydrolysis (Section 13-8). 

Amides can be converted to amines by reduction (Sections 13-8 and 16TE3), 
or by the Hofmann hypobromite reaction (Section 16TE2). In the latter case 
the amine that is formed has one less carbon atom than the amide : 



O UAIH4 .» RCH,NH, 



R-C 

\ 



NH > nSh* RNHa 



NaOH 



D. IMIDES 

Imides are N-acyl amides. Unlike amides they are considerably ionized in 
aqueous solution (p^ H A ~ 9) as a result of effective charge dispersal in the 
anion. 

00 OO 

11 II k ffl I: • il 

R-C^ /C-R « H ffi + R-C- e .X-R 

N ^N 

I 
H 

an imide 
Examples of cyclic imides are succinimide [5] (succinic acid is 1,4-butanedioic 



chap 16 organic nitrogen compounds 440 

acid) and the compound [6]. The common name of [6] is thalidomide, the 
drug which attained notoriety when its unforeseen teratogenic (fetus-deform- 
ing) properties were discovered. 



o 




o 


9 H 


// 




ll 


ll / 


H,C-C 


sO*^ 


c 


C— N 


1 \ 


r 


if x 


/ \ 


NH 




N- 


-HC C=0 


1 / 


^w 


^/ 


V / 


H 2 C-C 


^^ 


C 


H 2 C— CH 2 


\ 




II 




O 




o 





[5] [6] 



16-3 nitriles 

The carbon-nitrogen triple bond differs considerably from the carbon-carbon 
triple bond by being stronger (213 kcal vs. 200 kcal) and much more polar. 
Liquid nitriles have rather high dielectric constants compared to most organic 
liquids and are reasonably soluble in water. 

Nitriles absorb in the infrared in the region 2000 to 2300 cm -1 , owing to 
stretching vibrations of the carbon-nitrogen triple bond. 

The preparation of nitriles by S N 2 reactions of alkyl or allyl halides with 
cyanide ion has been mentioned before (Section 8-7), and this is the method 
of choice where the halide is available and reacts satisfactorily. Other useful 
syntheses involve cyanohydrin formation (Section 11-4A) or dehydration of 
the corresponding amide. 

The reduction of nitriles to amines and the hydrolysis of nitriles to amides 
have been discussed earlier (see Sections 16-1E3, 13-8, and 16-2B). 

Hydrogens on the a carbons (C-2) of nitriles are about as acidic as the 
hydrogens a to carbonyl groups; accordingly, esters of cyanoacetic acid 
undergo many reactions similar to those of the esters of malonic and aceto- 
acetic acids (Section 13-9C). The a positions of nitriles can be alkylated with 
alkyl halides by processes like the following: 



CH 3 CH 3 

H CH 3 

66% 



16'4 nitroso compounds 

Although C-nitroso compounds, R— N=0, have no special synthetic 
importance at present, they do possess some interesting properties. Primary 



sec 16.5 nitro compounds 441 

and secondary nitroso compounds are unstable and rearrange to oximes. 

H C H CH 3 OH 

C — C = N 

H 3 C / N N=0 CH 3 

2-nitrosopropane 

Tertiary and aromatic nitroso compounds are reasonably stable substances, 
which, although usually blue or green in the gas phase or in dilute solution, 
are isolated as colorless or yellow solids or liquids. The color changes are 
due to dimerization. 

e O O 
I II 
2R— N=0 . R— N— N— R 

© 

(green or blue) (yellow or orange) 



16-5 nitro compounds 

Nitro compounds make up a very important class of nitrogen derivatives. 
The nitro group (— N0 2 ), like the carboxylate anion, is well formulated as a 
hybrid of two equivalent resonance structures. 

O O e O ie 

<$>// <&/ mv 

R _ N < > R _ N or R _ N 

The hybrid structure is seen to have a full positive charge on nitrogen and a 
half-negative charge on each oxygen. The polar character of the nitro group 
results in lower volatility of nitro compounds than ketones of about the 
same molecular weight; thus the boiling point of nitromethane (mol. wt. 61) 
is 101°, whereas acetone (mol. wt. 58) has a boiling point of 56°. Surprisingly, 
the water solubility is low; a saturated solution of nitromethane in water is 
less than 10% by weight, whereas acetone is infinitely miscible with water. 

Nitro groups of nitroalkanes can be identified by strong infrared bands at 
about 1580 and 1375 cm -1 , whereas the corresponding bands in the spectra 
of aromatic nitro compounds occur at slightly lower frequencies. A weak 
transition occurs in the electronic spectra of nitroalkanes at around 2700 A; 
aromatic nitro compounds, such as nitrobenzene, have extended conjuga- 
tion and absorb at longer wavelengths (~3300 A). 

Nitro compounds are quite unstable in the thermodynamic sense and the 
heat of decomposition of nitromethane, according to the following stoichio- 
metry, is 67.4 kcal/mole : 

CH 3 N0 2 ► |N 2 + C0 2 +fH 2 AH= -67.4 kcal 

The rate of decomposition under ordinary conditions is immeasurably slow, 
however, and pure nitromethane is a stable, easily handled, compound. On 




chap 16 organic nitrogen compounds 442 

the other hand, some polynitro compounds, such as tetranitromethane, 
explode on shock and require very careful handling. 

2,4,6-Trinitrotoluene (TNT) is not set off easily by simple impact and is 
used in high-explosive shells. However, once set off, TNT explodes violently. 



0,N 



N0 2 
2,4,6-trinitrotoluene 

The characteristics of reasonable handling stability and high thermodyna- 
mic potential make the chemistry of nitro compounds particularly interesting 
and useful. 

Nitro compounds can be prepared in a number of ways, including the direct 
substitution of hydrocarbons with nitric acid. This reaction was discussed 

RH + HON0 2 ► RN0 2 + H 2 

earlier (Section 3-3B) in connection with the nitration of alkanes, and it was 
there noted that reaction is successful only when conducted at high tempera- 
tures in the vapor phase. Mixtures of products are invariably obtained. Direct 
nitration of aromatic compounds such as benzene, in contrast, takes place 
readily in the liquid phase. The characteristics of aromatic nitration are 
discussed in Chapter 22. 

Other routes to aliphatic nitro compounds include the reaction of an 
alkyl halide (of good S N 2 reactivity) with sodium nitrite. Suitable solvents are 
dimethyl sulfoxide and dimethylformamide. As can be seen below, formation 
of the nitrite ester by O- instead of N-alkylation is a competing reaction. 

CH 3 (CH 2 ) 6 Br + NaN0 2 ► CH 3 (CH 2 ) 6 N0 2 + CH 3 (CH 2 ) 6 ONO 

60% 30% 

Tertiary nitro compounds may be prepared by the oxidation of the corre- 
sponding amine with aqueous potassium permanganate solution (Section 
16-1F5). 

Nitro compounds are readily reduced to amines (Section 16TE3) and this 
offers a particularly useful synthesis of aromatic amines, as will be discussed 
in Chapter 22. Most aliphatic amines are more easily prepared other ways. 



16-6 some compounds with nitrogen-nitrogen bonds 

There are a number of important compounds containing nitrogen-nitrogen 
bonds. Among these are hydrazines, azo and diazo compounds, and azides. 



A. HYDRAZINES 

Organic hydrazines are substitution products of NH 2 — NH 2 and have many 
properties similar to those of amines in forming salts and acyl derivatives as 



sec 16.6 some compounds with nitrogen-nitrogen bonds 443 

well as undergoing alkylation and condensations with carbonyl compounds 
(Table 11-4). Unsymmetrical hydrazines can be prepared by careful reduction 
of N-nitrosoamines. 

CH 3 CH 3 CH 3 

\ \ 2 H 2 \ 

NH+HONO > N — NO ► N-NH, 

/ / / 

CH 3 CH 3 CH 3 

Aromatic hydrazines are best prepared by reduction of aromatic diazonium 

salts (Chapter 22). H H 

I I 
Hydrazines of the type R— N— N— R are usually easily oxidized to the 

corresponding azo compounds, R— N=N— R. 



B. AZO COMPOUNDS 

Azo compounds possess the — N=N— grouping. Aliphatic azo compounds 
of the type R— N=N— H are highly unstable and decompose to R — H and 
nitrogen. Derivatives of the type R— N=N— R are much more stable and 
can be prepared as mentioned above by oxidation of the corresponding hydra- 
zines. Aromatic azo compounds are available in profusion from diazonium 
coupling reactions (Chapter 22) and are of commercial importance as dyes 
and coloring materials. 

A prime characteristic of azo compounds is their tendency to decompose 
into organic radicals and liberate nitrogen: 

R-N=N-R ► 2R-+N 2 

The ease of these reactions is usually a fairly reasonable guide to the stabil- 
ities of the radicals that result. For instance, it is found that azomethane 
(CH 3 N=NCH 3 ) is stable to about 400°, and azobenzene (C 6 H 5 N=NC 6 H 5 ) 
is also resistant to thermal decomposition; but, when the azo compound 
decomposes to radicals that are stabilized by resonance, the decomposi- 
tion temperature is greatly reduced. Thus 2,2'-azobis(2-methylpropano- 
nitrile) decomposes to radicals at moderate temperatures (60° to 100°), and 
for this reason is a very useful agent for the initiation of polymerization of 
vinyl compounds. 

CN CN CN 

CH 3 — C — N=N— C— CH 3 6Q °~" )0 °> 2CH 3 — C-+N, 
I I I 

CH 3 CH 3 CH 3 

2,2-azobis(2-methylpropanonitrile) 



C. DIAZO COMPOUNDS 

e e 
The parent of the diazo compounds, diazomethane CH 2 =N=N, has been 
mentioned before in connection with formation of carbenes (Section 9-7). 
It is one of the most versatile and useful reagents in organic chemistry, despite 
the fact that it is highly toxic, dangerously explosive, and cannot be stored 
without decomposition. 



chap 16 organic nitrogen compounds 444 

Diazomethane is an intensely yellow gas, bp —23°, which is customarily 
prepared and used in diethyl ether or methylene chloride solution. It can be 
synthesized a number of ways, such as by the action of base on an N-nitroso- 
N-methylamide. 

o 

R-C + NaOH ether > R-C0 2 Na + CH 2 N 2 + H 2 

N— CH 3 
I 
NO 

As a methylating agent for moderately acidic substances, diazomethane 
has nearly ideal properties. It can be used in organic solvents ; it reacts very 
rapidly without need for a catalyst; the coproduct is nitrogen, which offers 
no separation problem; it gives essentially quantitative yields; and, because 
of its color and the appearance of bubbles of nitrogen, it acts as its own indi- 
cator to show when reaction is complete. With acids it gives esters and with 
enols it gives O-alkylation. 

\S~<\ +CH 2 N 2 ► \/~ C \ +N 2 

X==/ OH X=/ OCH 3 

O O OCH 3 O 

I II I II 

CH,-C ,C-CH, + CH 2 N 2 ► CH 3 — C C— CH 3 + N 2 

CH CH 

Diazomethane was originally believed to possess the three-membered, 
diazirine ring structure, but this was disproved by electron-diffraction studies, 
which show the linear structure to be correct. 



H 2 C II 

N 


CH 2 =N=N 


diazirine 


diazomethane 



Recently, a variety of authentic " cyclodiazomethanes " or, more properly, 
diazirines have been prepared, and these have been found to have very 
different properties from the diazoalkanes. The simple diazirines are colorless 
and do not react with dilute acids, bases, or even bromine. The syntheses of 
these substances are relatively simple. The route shown is one of the several 
possible ones. 

R R NH R N 

\ \ /\ Cl-Oj \ /ll 

C=0 + NH 3 + NH 2 C1 ► C > C 

/ 32 / \l / \ll 

R 7 R NH R N 

an " isohydrazone " 



summary 

Organic nitrogen compounds include amines (RNH 2 ), amine salts 
(RNH 3 ®X e ), amides (RCONH 2 ), nitriles (RC=N), isocyanates (RNCO), 



summary 445 

hydrazines (RNHNHR), azo compounds (RN=NR), diazo compounds 
(RCHN 2 ), azides (RN 3 ), nitroso compounds (RNO), and nitro compounds 
(RN0 2 ). 

Amines are associated by hydrogen bonding to a moderate extent and as a 
result they are more volatile than hydrocarbons but less volatile than alco- 
hols of the same molecular weight. However, amines are more soluble in 
water than alcohols because they form strong hydrogen bonds with the 
protons of hydroxylic compounds. Amines, like ammonia, undergo rapid 
inversion at nitrogen. Aliphatic amines are weak bases (K n ~ 10~ 4 ) and are 
very' feeble acids (K HA ~ 10" 33 ). 

Amines can be prepared as illustrated. 



RNO, 



-» RNH 2 



O 

II 
RCNH, 



RCH,NH, 



RX R 2 C=NOH RC^N 

*Hofmann reaction (see also Schmidt and Curtius reactions) 



RCH=NOH 



Some of the reactions of amines are 



RCH 2 NH 2 



RCH 2 NH 3 ® 
O 



(with acids) 



> RCH 2 NHCCH 3 (with acetyl halides) 

► N 2 + complex products (with nitrous acid) 

> RCHO ► RC0 2 H (with oxidants like Mn™; f-alkyl 

amines give nitro compounds) 

► RCH 2 N0 2 (with peracids; secondary amines give 

oximes, tertiary amines give amine oxides) 

Amides have high melting and boiling points as a result of strong hydrogen 
bonding. They absorb in the infrared near 3400 cm -1 (N— H stretch) and 
near 1690 cm -1 (C=0 stretch). They are weak acids (K HA ~ 10~ 16 ) and 
weak bases (K B ~ 10~ 15 ). Their preparations and reactions are summarized 
here. 




II 
RCOCH3 










II 


N. 


O 

II 
RCNH 2 


^ RC^N 




. 




(RCO) 2 


^ 




^ R 2 C=NOH 






w 

ICNH 2 




RC0 2 H 










y 


RNH 2 






» 


RCH 2 NH 2 



chap 16 organic nitrogen compounds 446 

Nitriles are rather polar compounds with moderate water solubility. The 
cyano group activates an adjacent methylene like a carbonyl group. Nitriles 
absorb in the infrared near 2200 cm -1 (C=N stretch). Some methods of 
preparation and reactions of nitriles are summarized. 

O 

II 
RX > RC^N < RCNH 2 RCHO ► RCHOHC=N 

O 

II 
RC=N ► RCNH 2 ► RCO z H 

^"~-* RCH 2 NH 2 

Nitro compounds absorb strongly in the infrared near 1580 cm -1 and 
1375 cm -1 . Polynitro compounds are often explosive. Nitro compounds can 
be prepared by nitration of hydrocarbons (works best with arenes) or by the 
route RX + N0 2 e -» RN0 2 + X e . 

Azo compounds can be prepared by oxidation of their dihydro derivatives 
(hydrazines). They are used as sources of radicals (RN==NR -+N 2 +2R-). 
Diazo compounds also tend to lose N 2 and the simplest of them, CH 2 N 2 
(diazomethane), is an excellent methylating agent for carboxylic acids and 
enols. 



exercises 

16-1 Provide a suitable name for each of the following compounds : 

a. H 2 N(CH 2 ) 7 NH 2 g. H 2 NCH 2 C0 2 H 

b. (CH 3 ) 2 CHC N h. H 2 NCH 2 CH 2 NH 3 Cl e 



NHCH 



c. C 6 H 5 CHC1CHC1CN /, /"A-j/ 

d. (CH 3 ) 3 CCH 2 N0 2 ~ CH 2 CH 3 

e. CH 2 =CHN(CH 3 ) 3 Br e J- I /^~ CH 3 l ° 



\ 



f. (HOCH 2 CH 2 ) 3 N 



16-2 How could you show with certainty that the peak at 47 Hz with reference to 
tetramethylsilane in the nmr spectrum of diethylamine (Figure 6-2) is 
actually due to the N— H resonance? 

16-3 Show how structures can be deduced for the substances (a) and (b) of 
molecular formulas C 8 HnN, whose nmr and infrared spectra are shown in 
Figure 16-5. 
















... — 


■_ 








o 
- o 

04 • 




tipliiillpii 
1 J 








^^^^^B 








6 








S 






Q 
" 2 


l 








V 



u 
o 



"3 
S 


/^ 

"8 
C 



a 

3 


Q- 

E 
o 
y 




chap 16 organic nitrogen compounds 448 

16-4 Guanidine (K B ~ 1) is a very strong base and an exception to the generaliza- 
tion that unsaturated amines are weaker bases than saturated amines. 
Consider various ways of adding a proton to guanidine and the kind of 
changes in stabilization energies which would be expected for each. 



/ 
HN=C . guanidine 

NH 2 

16-5 Write equations for the reaction that occurs when each of the following 
compounds is refluxed with an excess of aqueous sodium hydroxide. 

a. heptanamide 

b. N-chlorocyclohexylamine 

c. allyl 2,3-dimethylbutanoate 

d. heptanonitrile 

16-6 Write equations for the reaction that occurs when each of the following 
compounds is treated with an excess of lithium aluminum hydride and the 
resulting mixture decomposed with water and acid. 

a. diphenylacetonitrile 

b. 1,2-dinitrocyclopentane 

c. diphenylacetic acid 

d. 4-nitrohexanal 

16-7 Write structures and provide names for the simplest amine and the simplest 
nitrile that can be separated into stable optical isomers at room temperature. 

16-8 Write equations for a practical laboratory synthesis of each of the following 
compounds based on the indicated starting materials. Give reagents and 
conditions. 

a. dimethyl-?-pentylamine from 2-chloro-2-methylbutane 

b. (CH 3 ) 3 CCH 2 NH 2 from (CH 3 ) 3 CC0 2 H 

c. 1,6-diaminohexane from 1,3-butadiene 

d. butanonitrile from 1-butanol . 

e. (CH 3 C0 2 CH 2 ) 3 C-N0 2 from nitromethane 
/. N-;-butylacetamide from f-butyl alcohol 

g. methylethyl-M-butylamine oxide from «-butylamine 

16-9 a. Make a chart of the mp, bp, and solubilities in water, ether, dilute 
acid, and dilute base of each of the following compounds: 

«-octylamine N,N-dimethylacetamide 

di-«-butylamine 1-nitrobutane 
tri-«-propylamine 2-nitro-2-methylbutane 

b. Outline a practical procedure for separation of an equimolal mixture 
of each of the above compounds into the pure components. Note that 
selective reactions are not suitable unless the reaction product can be 
reconverted to the starting material. Fractional distillation will not 
be accepted here as a practical means of separation of compounds 
boiling less than 25° apart. 



exercises 449 

16-10 Give for each of the following pairs of compounds a chemical test, preferably 
a test tube reaction, which will distinguish between the two compounds. 

a. (CH 3 ) 3 CNH 2 and (CH 3 ) 2 NC 2 H 5 

b. CH 3 CH 2 N0 2 and CH 3 CONH 2 

c. CH 3 CH 2 C=N and CHs=C-CH 2 NH 2 

d. CH 3 CH 2 NHC1 and CH 3 CH 2 NH 3 C1 

e. CH 3 OCH 2 CH 2 NH 2 and CH 3 NHCH 2 CH 2 OH 

/. CH 3 CH 2 c' and CH 3 OCH 2 CH 2 NH 2 
NH 2 

16-11 Arrange the following pairs of substances in order of expected base strengths. 
Show your reasoning. 

a. (CH 3 ) 3 N and (CF 3 ) 3 N 



b. V \-CH 2 NH 2 and CH 3 -f V" NH ' 





c. CH 3 C=N: and 4 N: 

NH O 

// // 

H— C and H— C (review Exercise 16-4) 

NH, NH 2 

(review Section 6-7) 




16-12 Using spectroscopic methods, how could you distinguish one isomer from 
the other in the following pairs ? 

CH 3 

a. / V-NH 2 and / V-NHCH 3 

O O 

II II 

b. CH 3 CH 2 -C-NH 2 and H~C-N(CH 3 ) 2 

c. CH 3 CH 2 -N0 2 and CH 3 CH 2 -ONO 

16-13 It is usually possible to obtain two different stable oximes from an unsym- 
metrical ketone. These so-called syn and anti isomers are related to cis-trans 
isomers of alkenes. 

R OH R 

\ / \ 

C=N C=N 

/ / \ 

R' R' OH 

(syn to R, anti to R') (syn to R', anti to R) 



chap 16 organic nitrogen compounds 450 

An important characteristic of these isomers is that they give different 
products in the Beckmann rearrangement. In each case, the group anti to the 
OH group migrates to nitrogen. Decide whether or not the mechanism 
given in Section 16-1E2 for the Beckmann rearrangement can account for 
this fact and, if not, how it might be modified to do so. What products 
might be expected from the syn and anti forms of benzaldehyde oxime ? 

16-14 Show how one could synthesize and resolve methylethylallylamine oxide 
from allylamine with the knowledge that amine oxides are somewhat basic 
substances having K B values of about 10"". 

16-15 The proton nmr spectrum of N,N-dimethylformamide shows a single proton 
resonance at 8.06 ppm and two separate three-proton resonance at 2.78 and 
2.95 ppm at room temperature. At 150°, the two three-proton lines are 
found to have coalesced to a single six-proton line, while the single-proton 
line is unchanged. Explain the nmr spectrum of this compound and its 
behavior with temperature. What would you predict the energy barrier 
would be for the process by which the lines are caused to coalesce at elevat- 
ed temperatures ? 

16-16 Amides with structures like the following are difficult to prepare and are 
relatively unstable. Explain. 




16-17 Benzamidine [7] has a p^ B of 2.4, making it about 10 14 times as strong a 
base as benzamide [8]. 

NH O 

t t 

C 6 H 5 -C^ C 6 H 5 -C^ 

NH 2 NH 2 

[7] [8] 

How do you account for this large difference in view of the fact that pro- 
tonation produces a resonance-stabilized cation in both cases ? 

16-18 Show how structures can be deduced for the two substances with the mo- 
lecular formula C5H9NO3 and Ci H ]3 NO from their infrared and nmr 
spectra, as given in Figure 16-6. 

16-19 Nitriles of the type RCH 2 CN undergo an addition reaction analogous to 

the aldol addition in the presence of strong bases such as lithium amide. 

Hydrolysis of the initial reaction product with dilute acid yields a cyano- 

O CN 

II I 

ketone, RCH 2 — C— CH— R. Show the steps that are involved in the 

mechanism of the overall reaction and outline a scheme for its use to synthe- 
size large-ring ketones of the type (CH 2 )„C=0 from dinitriles of the type 
NC(CH 2 )„CN. 



wavelength, jJL 




8 9 10 12 14 

~i 1 i ^t r 



n \ ! 



M 



*J' 



i i i _j j _l_j 



<- U — ^ 



J . 1 1 ,J L_ 



3600 3200 2800 2400 200O 2000 1800 1600 ,1400 1200 1000 800 



frequency, cm 





i 

400 




1 


i 

oil/ 








IIIIIIJ 


' 




i /-- 


c:..H, 


NO 

III|iIlI|illlll|i|iSlliillllli 


















_J\)\ j 

1 1 


I 


1 


i 


1 




r~ 




f 



i t ~r 



wavele 


ngth, [X 












6 


7 


8 


9 


10 


12 


14 


i 


1 


1 


1 


j- 


1 


1 



( : 1 1 \( ) 



_L_- J I i. 





3600 3200 2800 2400 2000 2000 



1800 1600 ,1400 1200 

frequency, cm 



1 000 800 





'" 1 

400 




1 

200 


i 

:, 11/ 












> 






C H.No 




; — 


, 






/~ 





■ 


J 

II: 










1 




1 


— w 

i 


~ — — 




■ 


»- 



6.0 4.0 2.0 ppm 8.0 6.0 4.0 2.0 

Figure 16-6 Infrared and nmr spectra of a substance C 10 H 13 NO and a substance C 5 H 9 N0 3 . See Exercise 1648. 



o ppm 



chap 16 organic nitrogen compounds 452 

16-20 Show how the following substances may be synthesized from the indicated 
starting materials. 

a. (CH 3 ) 3 CCN from (CH 3 ) 3 CC1 

b. CH 3 CH=CHCN from CH 2 =CHCH 2 Br 

c. CH 2 =CHC0 2 HfromCH 3 CHO 

16-21 Consider possible ways of formulating the electronic structures of nitroso 
dimers with the knowledge that X-ray diffraction studies indicate the presence 
of nitrogen-nitrogen bonds. 

16-22 Show from the electronic structure of nitrite ion how it could react with an 
alkyl halide in the S N 2 manner to give either a nitrite ester or a nitro com- 
pound. What kind of properties would you expect for a nitrite ester and 
how could they be removed from the reaction products ? 

16-23 Arrange the following azo substances in order of their expected rates of 
thermal decomposition to produce nitrogen. Give your reasoning. 

a. I V-CH 2 -N=N-CH 2 ^ \ d. CH 3 -N=N-CH 3 

b. (CH 3 ) 3 C-N=N-C(CH3)3 e - ILJJI 

c. /\\-n=n 

16-24 Nitriles are converted readily to amides with hydrogen peroxide in dilute 
sodium hydroxide solution. The reaction is 

P 
RC=N + 2H 2 2 ° H > RC + 2 + H 2 

NH 2 

The rate equation is 

e 
v = /t[H 2 2 ][OH][RC=N] 

When hydrogen peroxide labeled with ls O (H 2 ,8 2 ) is used in ordinary 
water (H 2 1<s O), the resulting amide is labeled with ls O (RC I8 ONH 2 ). 

Write a mechanism for this reaction which is consistent with all the 
experimental facts. Note that hydrogen peroxide is a weak acid (K A ~ 10~ 12 ) 
and, in the absence of hydrogen peroxide, dilute sodium hydroxide attacks 
nitriles only very slowly. 

16-25 An amine, A, forms a precipitate with chloroplatinic acid of formula 
(AH®) 2 PtCl 6 2e . A sample of 0.1834 g of this salt is found to contain 0.0644 g 
of platinum. On treatment with nitrous acid the amine, A, gives an alcohol 
which gives a positive iodoform test. What is the structure of A ? 

16-26 The oxidation of an aldehyde to an acid and a secondary alkyl amine to a 
ketone can be carried out as shown in the following equations, which ap- 




exercises 453 

pear to involve merely addition or removal of water. How do the oxidations 

arise? 



RCHO + H 2 NOH -^ RCH=NOH — ^+ RC=N -^^* RCONH, 



+ H 2 



RC0 2 H 
HO 

- R 2 C=N^} 



o o 

R 2 CHNH 2 + 0=/ \ ^ H2 °> R.CHNH^ J> 



H 2 



HO 



16-27 Use bond energies (Table 2-1) to evaluate qualitatively the position of the 
following equilibrium. 

CH,=CH-NH 2 < ' CH,-CH=NH 



1 



V:VS 3 



and "nucleic acids 



chap 17 amino acids, proteins, and nucleic acids 457 

The chemistry of life is largely the chemistry of polyfunctional organic com- 
pounds. The functional groups are usually of types which interact rather 
strongly and are often so located with respect to one another that both 
intra- and intermolecular interactions can be important. Carbohydrates offer 
one example, and we have seen how the alcohol and carbonyl functions 
interact in these substances, leading both to the cyclization of the simple 
sugars and to the formation of bonds between sugar molecules to give poly- 
saccharides. This chapter is devoted to the chemistry of amino acids, pro- 
teins, and nucleic acids. 

Apart from water, protein is the principal component of muscle and many 
other kinds of tissue. Nucleic acids, the substances which control heredity, are 
just as widely distributed in living organisms although they are present in 
smaller amounts. We shall see that both proteins and nucleic acids are poly- 
meric materials built up with polyfunctional compounds as units. In proteins, 
these units are amino acids and we shall begin with a discussion of their 
chemistry. 



17-1 amino acids 

All the natural amino acids which occur as constituents of proteins are 
carboxylic acids with an amino group at the a position (C-2), RCHNH 2 C0 2 H. 
All, with the exception of the simplest one (glycine, R = H), have a center of 
asymmetry at the a position and belong to the l series, corresponding to the 
projection formula shown (see also Section 14-3B). 

C0 2 H 

I 
H 2 N— C-H 

I 
R 

L-amino acid 

In the case of two amino acids, proline and hydroxyproline, the amino 
group is secondary by virtue of being part of a ring. The structures and names 
of these and other important a-amino acids are shown in Table 17-1. You 
can see that the names in common use for amino acids are not very descriptive 
of their structural formulas ; but they do have the advantage of being shorter 
than the systematic names. As you will see later, the abbreviations Gly, Glu, 
and so on listed in Table 17-1 are particularly useful in designating the se- 
quences of amino acids in proteins and peptides. Amino acids that have an 
excess of amine over acid groups are called basic amino acids (e.g., lysine and 
arginine) ; those with an excess of acid groups are called acidic amino acids 
(e.g., aspartic and glutamic acids). The additional amino or carboxyl groups 
that these four amino acids possess and the functional groups that amino 
acids such as serine possess are important in providing points of attachment 
for other groups to the protein chain. These may be covalent, ionic, or 
hydrogen-bond links. 

Three of the amino acids listed in Table 17-1, cysteine, cystine, and methio- 
nine, contain sulfur, and the making and breaking of S— S linkages in the 



chap 17 amino acids, proteins, and nucleic acids 458 

interconversion of cysteine and cystine are important processes in the bio- 
chemistry of sulfur-containing peptides and proteins. Further characteristics 
of the general reaction 

[O] 

2RSH . R-S-S-R 

2[H] 

are considered in Chapter 19. 

Organisms differ considerably in their ability to synthesize amino acids. 
The eight acids that are indispensable for human beings, yet which the body 
cannot synthesize, are often called " essential " amino acids and are so marked 
in Table 17T. Of the 24 amino acids listed in the table only 20 are fundamental 
building blocks. The presence in proteins of the remaining four (see Table 
17T) results from in vivo conversions of other amino acids after the protein 
has been synthesized. Protein synthesis, both in vitro and in vivo, is considered 
in more detail later in the chapter. 



A. SYNTHESIS OF a-AMINO ACIDS 

Many of the types of reactions useful in preparing amino acids in the labora- 
tory have been discussed earlier in connection with separate syntheses of 
carboxylic acids (Chapter 13) and amino compounds (Chapter 16). Two 
examples are included here. 

1 . Amination of chloroacetic acid to yield glycine, which works best with 
a large excess of ammonia. 

50° e ffl ® e 

3 NH 3 + C1CH 2 C0 2 H ► NH 2 CH 2 C0 2 NH 4 + NH 4 C1 

H© 

NH 2 CH 2 C0 2 H 

glycine 

2. Strecker synthesis; in its first step, it bears a close relationship to cyano- 
hydrin formation. 

f VCHO + NH 3 + HCN ► / VCH-C = N H* B ,H 2 > / Y-CH— C0 2 H 

NH 2 NH 2 

Many amino acids are very soluble in water and it may be necessary to 
isolate the product either by evaporation of an aqueous solution or by precip- 
itation induced by addition of an organic solvent such as alcohol. Difficul- 
ties are often encountered in obtaining a pure product if inorganic salts are 
coproducts of the synthesis. 



B. THE ACID-BASE PROPERTIES OF AMINO ACIDS 

The behavior of glycine is typical of that of the simple amino acids. Since 
glycine is neither a strong acid nor a strong base, we shall expect a solution of 



sec 17.1 amino acids 4S9 



Table 17-1 Amino acids important as constituents of proteins 



*\9 






73 






3 g © 


«/i 


oo 




<N 


>Ti 


* u W 


<N 


>— •< 








+j 41 o 












£.£ M 







8*9 

o'SI 
.2 a a 



<X 



o 
in 



o o 
\6 \6 



o 



r- 






ON 



ffi 



s 








o 


fa 

a 


o 
o 


d 

r ) 




u 

K 
o 

a 




33 


a 


ffi 


u 




u 


u- 


"Z 


/-^ 




a 


a 




a 




Z 


u 




y, 


a 































« 










>> 


t8 




, 












U 





< 




> 


-0 










J3 











X 

-z 



o 
u 

E 

u- 

a 
u 
a 
u 

a 
u 



o 

a a 
0-z 
i „ 

3-5 

u 

a 
u 



o 
u 

a 
u- 

a 



o 

a* a 

u-z 

a 
u 
o 

a 



0- 



o 
u 
a 
u- 

a 
u- 

a 



H 



a 

d 
u 
a 
o- 

a 
u 

N 

a 
z 



o 

a a 
o-z 

a 
o 

a a 
u-o 

a 
u 

a 
z 



a 



p3 

nv 1 


o' 

3 


o 


-t 






"i 


hi 


fft 




O 


D 




~t 


3/ 


er 


w 


O 


ll) 


<TQ 


3 




3" 


O 














3 


O 


s; 










1= 


n 




rr 












o 





s s. >- 



g* <? 



£L c 






M <-'■ .T* 



abbreviation 



formula 



[M] D of L isomer, 
H 2 0, 25°" 



arginine 
aspartic acid 
asparagine 
glutamic acid 
glutamine 
cysteine 



Arg 
Asp 
Asn 
GIu 
Gin 
CySH 



HN. 



H 2 N 



"C-NH(CH 2 ) 3 CHC0 2 H 



NH, 



H0 2 CCH 2 CHC0 2 H 

I 
NH 2 

NH 2 COCH 2 CHC0 2 H 
I 

NH 2 

H0 2 C(CH 2 ) 2 CHCO 2 H 
I 
NH 2 

NH 2 CO(CH 2 ) 2 CHC0 2 H 
I 

NH 2 

HSCH 2 CHC0 2 H 
I 
NH 2 



[H] 



[O] 



c= 1-2 g 
HCl solut 
e Section 


cystine" 1 


CyS 
CyS 


S-CH 2 CHC0 2 H 
I 1 
1 NH 2 
S-CH 2 CHC0 2 H 


per 100 
ions. 
14-3A). 


methionine' 


Met 


1 

NH 2 

CH 3 S(CH 2 ) 2 CHC0 2 H 


3 






NH 2 



+21.8° 



-6.7° 



-7.4° 



+ 17.7° 



+ 9.2° 



-20.0° 



^509° 
(1 iVHCl) 



-14.9° 



isoelectric 
point, 
pH units 



water solubility at 
isoelectric point,* 
g/100 g, 20° 



11.2 



2.8 



5.4 



3.2 



5.7 



5.1 



5.0 



5.7 



very sol. 



0.4 



2.4 



0.7 



3.6 1 



very sol. 



0.009 



3.0 



sec 17.1 amino acids 461 



Table 17-1 (continued) 



+J 














(8-°. 














>>"£ 














-^ .3 














3 














3 °-o 














3 U© 

"3 '51 


s 




>n 


i/-> 


-. 


o 






•sr 


Tf 




rr 


" o oB 


o 




v^ 


en 






i. IB _ 














U-O 














•S uo 














« — 




























U 




























h <*> 




























« *• 5 

us 






CO 

Mi 




ON 


u"N 


S 5 * 














.a a a 














c 














t> 














s 





























m 


o 


00 


C-) 


NO 


00 


00 


, o 


OO 


en 


On 


ON 


oo 


On 


fa- "> 


7 


1 


on 

1 


On 
1 


NO 

1 


vn 

1 
















JSO 














Si 


a: 


X 

IN 

o 

X 

u 






X 




.3 


o 


i^S 






O 




3 


1 5 








X X 


X 

o 

X X 




u— Z 

x 
u 

1 


o 

1 


X 
O 


X 

o 
u 


u-z 

X 

o 




r^ 


rS 


v \ 

I zx 


Xjj 




ZX 


U-Z 

fN 

X X 




v 


AA- 


zx 
6 r 


, 








T 

o 


T 

o 




x^o 

x x 


( > 


II 5 


a 


X 


X 


X 


\=/ 


X z 































.2 


>> 

H 


>> 

J3 

H 


o 


0, 

X 


H 


X 


jo 














■o 














« 








a 






a> 

E 








"o 
>> 


C3 

J3 


<u 


rt 


0) 




Stf 


x 


& 


.s 


S 


,G 


X 

o 


# o 


o 


o 


.*§ 




o 


>> 

J3 


"o 

P. 


•6 
>> 

J3 


>> 


to 

"-S 

















: Must be included in diet of normal adult humans. 

1 Formed by conversion of the other amino acids after the protein has been synthesized. 

' In ethanolic sodium hydroxide solution. 



chap 1 7 amino acids, proteins, and nucleic acids 462 

glycine in water to contain four species in rapid equilibrium. 
NH 2 CH 2 CO,H 



-H© 



t-H© 



NH 3 CH 2 C0 2 

dipolar ion 
(zwitterion) 



-H© 



+H© 



NH 2 CH 2 C0 2 

conjugate base 
of glycine 



NH 3 CH 2 C0 2 H 

conjugate acid 
of glycine 

The proportions of these species are expected to change with pH, the con- 
jugate acid being the predominant form at low pH values and the conjugate 
base being favored at high pH values. Since the establishment of equilibrium 
between the uncharged molecule and the dipolar ion (often called a " zwit- 
terion") involves no net change in hydrogen or hydroxide ion concentrations, 
the ratio of these two substances is independent of pH. The position of equi- 
librium, however, strongly favors the zwitterion. 



NH,CH 2 CO,H 



NH,CH 2 C0 2 



A titration curve starting with glycine hydrochloride is shown in Figure 
17T. Two equivalents of base are required to convert ffi NH 3 CH 2 C0 2 H to 
NH 2 CH 2 C0 2 e . The pH of half-neutralization during addition of the first 
equivalent of base corresponds to proton loss from the carboxyl group of the 
conjugate acid of glycine, K BK ® = 5 x 10~ 3 , whereas the pH of half-neutral- 
ization during the addition of the second equivalent of base corresponds to 
proton loss from the ammonium group of the zwitterion, K HA = 2 x 10~ 10 . 
There will be a pH on the titration curve where the concentration of zwit- 

terion is at a maximum and where the concentration of NH 3 CH 2 C0 2 H will 
be just equal to the concentration of NH 2 CH 2 C0 2 e . If these two ions con- 
duct electric current equally well, then, in an electrolytic experiment, there 



Figure 17-1 Titration curve of the conjugate acid of glycine, NH 3 CH 2 C0 2 H, 
with base. 



PH 




moles of OH 



sec 17.1 amino acids 463 

will be no net migration of the ions at this particular pH, which is called the 
isoelectric point. Generally, the isoelectric point corresponds also to the pH 
at which the amino acid's water solubility is least. Isoelectric points are 
listed for most of the amino acids shown in Table 17-1. 

Because the basic dissociation constant, K B , of most aliphatic amines and 
the acid dissociation constant, K UA , of most aliphatic acids are comparable, we 
expect the isoelectric point for simple amino acids like glycine to be not far 
from neutrality. The value of 6.0 shown in Table 17-1 indicates that the car- 
boxyl group in glycine is actually very slightly stronger than the amino group. 

A basic amino acid such as lysine exists as a neutral molecule (actually an 
equilibrium mixture containing one of the zwitterions as the principal 
component) at a somewhat higher pH. The isoelectric point of this compound 
is 9.7. 



NH 2 CH 2 CH 2 CH 2 CH 2 CHC0 2 e ^zzr^ NH 3 CH 2 CH 2 CH 2 CH 2 CHC0 2 

I I 

NH 3 NH, 



An acidic amino acid such as glutamic acid, by contrast, is only electrically 
neutral in the presence of enough excess acid to prevent the second carboxyl 
group from ionizing. Its isoelectric point is 3.2, which means that the electri- 
cally neutral forms shown here are at their highest concentrations at this pH. 
(The free amino or carboxyl groups in protein chains provide acidic or basic 
centers in these molecules in addition to serving as points of attachment for 
other groups.) 

H0 2 CCH 2 CH 2 CHC0 2 e -^=-* ®0 2 CCH 2 CH 2 CHC0 2 H 
I I 

NH, NH, 



C. ANALYSIS OF AMINO ACIDS 

Nitrous Acid Reactions. The action of nitrous acid on amino acids pro- 
ceeds in a manner similar to that discussed earlier for ordinary amines (Sec- 
tion 16-1F3). Primary amino groups are lost as nitrogen; secondary amino 
groups are nitrosated, whereas tertiary amino functions react to give complex 
products without evolution of nitrogen. Measurement of the nitrogen evolved 
on treatment of amino acids or their derivatives with nitrous acid provides a 
useful analysis for free — NH 2 groups in such materials (Van Slyke amino- 
nitrogen determination). With amino acids, as with amines, the nitrous acid 
reaction is not to be regarded as a generally satisfactory preparative method 
for conversion of RNH 2 to ROH. 

The Ninhydrin Test. In many kinds of research it is important to have 
simple means of detecting compounds in minute amounts. Detection of 
a-amino acids is readily achieved by the " ninhydrin color test." An alcoholic 



chap 17 amino acids, proteins, and nucleic acids 464 

solution of the triketone hydrate called " ninhydrin," heated with a solution 
containing an amino acid, produces a blue-violet color. The sensitivity and 
reliability of this test is such that 0.01 micromole of amino acid gives a 
colored solution whose absorbance is reproducible to a few percent, provided 
that oxidation of the colored ion by dissolved oxygen is prevented by the 
addition of a reducing agent such as stannous chloride. 

o o 

II II 

c=o ;=± \ T C 

II II 

o o 

indane-l,2,3-trione "ninhydrin" 

The color-forming reaction is interesting because all amino acids except 



Figure 17-2 The color-forming reaction of ninhydrin (in the trione form) 
with an amino acid. 




O 




O 



■H 2 



^V C \ i 



C=0 + RCH-C0 2 H ^^ C=N-CH-CO,H 

9 NH ^^C 7 

II NH 2 || 

o o 




o o 

blue-violet (/l max 5700 A) 



sec 17.1 amino acids 465 




Figure 17-3 Diagram of apparatus used to develop a paper chromatogram. 
Paper is suspended from its top edge within an airtight container (here a 
glass box closed with a glass plate) having an atmosphere saturated with sol- 
vent vapor; the lower edge of paper dips into a trough containing the liquid 
solvent. 



proline and hydroxyproline give the same color (x max = 5700 A). The sequence 
of steps that leads to the color is shown in Figure 17-2. The secondary amino 
acids proline and hydroxyproline give a yellow color (A max = 4400 A) with 
ninhydrin. 



Chromatography. The analysis of mixtures of amino acids (or other 
compounds) can be conveniently carried out by any of three chromatographic 
techniques — paper chromatography, column chromatography, or thin-layer 
chromatography (tic). In Chapter 7 we described the principles of chroma- 
tography on which each of these methods is based. 

In paper chromatography, amino acids are separated as the consequence of 
differences in their partition coefficients between water and an organic solvent. 
The aqueous phase is held stationary in the microporous structure of the 
paper. The differences in partition coefficients show up as differences in rates 
of migration on the surface of moist (but not wet) filter paper over which is 
passed a slow flow of water-saturated organic solvent. 

We shall discuss one of several useful modes of operation. In this, a drop of 
the solution to be analyzed is placed on the corner of a sheet of moist filter 
paper, which is then placed in an apparatus such as shown in Figure 17-3, 
arranged so that the organic solvent can migrate upward by capillarity across 
the paper, carrying the amino acids with it along one edge. The acids that have 
the greatest solubility in the organic solvent move most rapidly and, before 
the fastest moving acid reaches the top of the paper, the paper is removed, 
dried, then turned sideways and a different solvent allowed to diffuse across 
the width. This double migration process gives a better separation of the 
amino acids than a single migration and results in concentration of the 
different amino acids in rather well-defined zones or spots. These spots can 



methionine 
® 






tryptophan . 
«k alanine 

® ® 

arginine 

® , ■ 

lysine 

® 


glycine 
' ® 


aspartic acid 
® 






1 1 


1 


1 



chap 17 amino acids, proteins, and nucleic acids 466 



0.4 s 



0.6 0.4 

phenol-water, 5 : 1 



0.2 



0.2 



- start 



Figure 17-4 Idealized two-dimensional paper chromatogram of a mixture 
of amino acids. The horizontal and vertical scales represent the distance of 
travel of a component of the mixture in a given solvent in a given time relative 
to that of the solvent itself. This is known as the Rf value and is fairly constant 
for a particular compound in a given solvent. A rough identification of the 
amino acids present in the mixture may therefore be made on the basis of 
their Rf values 



solvent 
front 



amino 
acid 



$ Rf=- 



be made visible by first drying and then spraying the paper with ninhydrin 
solution. The final result is as shown in Figure 17-4 and is usually quite 
reproducible under a given set of conditions. The identities of the amino acids 
that produce the various spots are established by comparison with the beha- 
vior of known mixtures. 

Paper chromatography has proved very useful in following the mechanisms 
of biological processes using radioactive tracers. For example, in the study of 
the fixation of carbon dioxide in photosynthesis, it was found possible to 
determine the rate of incorporation of radioactive carbon from carbon dioxide 
into various sugars, amino acids, and the like, by separating the products of 
photosynthesis at a succession of time intervals on paper chromatograms, 
then analyzing their radioactivities by scanning the paper with a Geiger 
counter or by simply measuring the degree of fogging of an X-ray film laid 
over the chromatograms. 

A quantitative method of analysis of amino acids can be achieved using 
column chromatography (Section 7-1). The solution to be analyzed is passed 
through columns packed with an ion-exchange resin and this separates the 
amino acids according to their ability to be complexed with the highly polar 
sites of the resin. The effluent from the column is mixed with ninhydrin solu- 
tion and the intensity of the blue color developed is measured with a photo- 
electric colorimeter and plotted as a function of time at constant flow rates. A 
typical analysis of a mixture of amino acids by a machine constructed to carry 
out the procedure automatically is shown in Figure 17-5. 



sec 17.2 lactams 467 

Thin-layer chromatography (tic) is similar to paper chromatography 
except that the adsorbent (usually silica gel) is applied in a thin layer to a 
glass plate. Silica gel is mixed with plaster of Paris (CaS0 4 ) and a small 
amount of water and the resulting slurry is coated on a plate which may be as 
small as a microscope slide or as large as a medium-sized window pane. The 
samples are then spotted on the adsorbent as in paper chromatography and 
the Rf values determined the same way. 

Thin-layer chromatography is now very widely used because it combines 
the best features of paper and column chromatography — that is, the ease of 
locating compounds by spraying with various reagents as in paper chroma- 
tography and the wide range of adsorbents available in column chromato- 
graphy. It is rapid, and more drastic reagents can be used for locating spots 
than is possible with paper chromatographs. 



17-2 lactams 

The cyclization of hydroxy acids through lactone formation has been dis- 
cussed in Chapter 13. The corresponding cyclization of amino acids leads to 
lactams. 



O 

// 
H,C-C 
2 I \ 
H 2 C V O 

CH 2 



O 

// 
H 2 C-C 
I \ 
H 2 C N ^NH 
CH 2 



y-butyrolactone 
(lactone of 4-hydroxybutanoic acid) 



y-butyrolactam 
(lactam of 4-aminobutanoic acid) 



Figure 17-5 Section of amino acid chromatogram obtained by the method of 
automatic amino acid analysis from a hydrolyzed sample of the enzyme 
ribonuclease. The amino acids are identified by their position of elution and 
are quantified by integration of the area under the curve. 



i 


i 




f Thr S 


llllliSttilllliSlil 




















§&!t^SSK^M 






Ala 




« 






1 




I 










> 










1 












c 'S 




1 










c;i\ 






o o 




















■£ T3 




| 1 




I 












&.S 




j 




1 












V3 'O 








J 












£> >, 




















<" J3 




J 1 


1 


1 












C 




I j 




1 












'2 




' u 


u 


j Pro 

U uL_ 






JL 




100 


1 


i 

40 


180 
volume (ml) of effluent 


i 

220 


*- 


260 



chap 17 amino acids, proteins, and nucleic acids 468 

Formation of a- and /^-lactams is expected to generate considerable ring 
strain, and other more favorable reactions usually intervene. Thus, while 
y-butyrolactam can be prepared by heating ethyl 4-aminobutanoate, the 
corresponding a-lactam is not formed from ethyl 2-aminobutanoate but, 
instead, the dimeric diethyldiketopiperazine [1] with a six-membered ring 
results. 

- CH2 "'° „c /° 



HP or h - c * OH H 2 C I 

H2C -NH° C2H5 hV NH 



CH 



2 I -2 C 2 H 5 OH 

NH, 



CH n 



O" CH 

I 



diethyldiketopiperazine 
[1] 

^-Lactams have been rather intensively investigated following the discovery 
that the important antibiotic penicillin G [2], produced by fermentation with 
Penicillium notatum, possesses a /Mactam ring. 

O H 

/~\ 'I Us ch 

\=/ I I A 

^ C - N -CH X CH 3 

COjH 
penicillin G (benzylpenicillin) 
[2] 

The problem of determining the correct structure of penicillin was an 
extraordinarily difficult one because the molecule is very labile and undergoes 
extensive rearrangements to biologically inactive products even under very 
mild conditions. The /Wactam structure was finally established by X-ray 
diffraction analysis. Penicillin G and many closely related compounds with 
different groups in place of the benzyl group have been synthesized. 



17 '3 peptides 



We saw earlier that the reaction of a carboxyl function with an amine pro- 
duces an amide. 

o o 

R-C + NH 2 R' ► R-C + HZ 

Z NHR' 



sec 17.3 peptides 469 

If both the reactants are bifunctional, each containing an a-amino group 
and a carboxyl function, the reaction can proceed further to form a poly- 
amide. 

o o o 

II II II 

R-C-NH-CH-C-NH-CH-C-NH-CH 

I I I 

R R R 

Such molecules are called peptides (or polypeptides) and, of course, may be 
so constituted that all of the R groups are not the same. They are classified 
according to the number of amino acid groups in the chain and are named as 
derivatives of the amino acid with the free carboxyl group, the amide group 
being called the peptide linkage. 

SH 
I 
O CH 3 O CH 2 O CH 2 OH 

e II I e II I II I 

H 3 N-CH 2 -C-NH-CH-C0 2 e H 3 N-CH-C-NH-CH-C-NH-CH-C0 2 e 

CH 3 

glycy lalanine (H • Gly • Ala • OH) alanylcysteinylserine (H • Ala • CySH • Ser • OH) 

a dipeptide a tripeptide 

Peptides, being polyamides, can be hydrolyzed under the influence of acids 
or bases (or enzymes) to give their constituent amino acids. 

The distinction between a protein and a peptide is not completely clear. 
One arbitrary choice is to call proteins only those substances with molecular 
weights greater than 10,000. The distinction might also be made in terms of 
differences in physical properties, particularly hydration and conformation. 
The naturally occurring peptides have relatively short flexible chains and are 
hydrated reversibly in aqueous solution ; proteins, by contrast, have very long 
chains which appear to be coiled and folded in rather particular ways, with 
water molecules helping to fill the interstices. Under the influence of heat, 
organic solvents, salts, and so on, protein molecules undergo more or less 
irreversible changes, called denaturation, in which both the conformations of 
the chains and the degree of hydration are altered. The result is usually a 
decrease in solubility and loss of ability to crystallize. Proteins thus have a 
very high degree of conformational specificity. Recent success in synthesizing 
polypeptides with enzymic activity (Section 17-3C) suggests that the folding 
of the peptide chain to give the conformationally correct enzyme is a conse- 
quence of the way the amino acids are arranged in sequence. 

Proteins, particularly those with hormonal or enzymic functions, often 
have other groups attached to the polypeptide chain. These groups are called 
prosthetic groups. 



A. PEPTIDE ANALYSIS 

A wide variety of peptides occur naturally. However, of those whose struc- 
tures have been determined, a considerable proportion contain one or more 
amino acids which are not found as constituents of proteins. Indeed, many 



chap 17 amino acids, proteins, and nucleic acids 470 

peptides are cyclic, and in some even D-amino acids occur. Most of the well- 
characterized peptides contain 3 to 10 amino acid units. 

The properties of peptides and of proteins are a critical function of not only 
the number and kind of their constituent amino acids but also the sequence in 
which the amino acids are linked together. Analyses for amino acid content 
can be made by complete hydrolysis and ion exchange separations, as des- 
cribed in Section 17-1C. Establishment of the sequence of amino acids is 
much more difficult but has been carried through to completion on peptide 
chains having more than 100 amino acid units. 

The general procedure for determining amino acid sequences is to estab- 
lish the nature of the end groups and then, by a variety of hydrolytic or 
oxidative methods, break up the chain into peptides having two to five 
amino acid units. The idea in using a variety of ways of cutting the chains is 
to obtain fragments with common units which can be matched up to one 
another to obtain the overall sequence. 

Determination of the amino acid that supplies the terminal amino group in 
a peptide chain (the N-terminal acid) is best made by treatment of the peptide 
with 2,4-dinitrofluorobenzene, a substance which is very reactive in nucleo- 
philic displacements with amines but not with amides. The product is a 
N-2,4-dinitrophenyl derivative of the peptide which, after hydrolysis of the 
amide linkages, yields a N-2,4-dinitrophenylamino acid. See Figure 17-6. 

This substance can be separated from the ordinary amino acids resulting 
from hydrolysis of the peptide, owing to the low basicity of the 2,4-dinitro- 
phenyl-substituted nitrogen, which greatly reduces the solubility of the com- 
pound in acid solution and alters its chromatographic behavior. 



B. PEPTIDE SYNTHESIS 

The problem of synthesizing peptides is of great importance and has received 
considerable attention. The major difficulty in putting together a chain of, say 



Figure 17-6 Determination of the terminal amino acid possessing the free 
amino group in a peptide chain. 



N0 2 NO, 

O / 2 o 



2 N^ V-F + NH 2 CH,C-Z ^* 2 N-f VnH-CH,C- 

\=s peptide 

2,4-dinitrofluorobenzene ^/H 2 



Z 



N0 2 

2 N-f V-N~CH 2 -C0 2 H + amino acids 

N-(2,4-dinitrophenyl) glycine 
(low solubility) 



sec 17.3 peptides 471 

100, amino acids in a particular order is one of overall yield. At least 100 
separate synthetic steps would be required and, if the yields in each step are all 
equal to n x 100%, the overall yield is (« 100 x 100%). Thus, if each yield is 
90%, the overall yield is only 0.003%. Obviously, a laboratory synthesis of a 
peptide chain comparable in size to those which occur in proteins must be a 
highly efficient process. The extraordinary ability of living cells to achieve 
syntheses of this nature, not of just one but of a wide variety of such substances, 
is truly impressive. 

Several methods for the formation of amide bonds have been discussed 
in Chapters 13 and 16. The most generally useful reactions are of the type 
where X is halogen, alkoxyl, or acyloxy, corresponding to acyl halides, 
esters, or acid anhydrides. 

o o 

/ / 

R-C + NH 2 -R' ► R-C + HX 

\ \ 

X NHR' 

When these are applied to join up two different amino acids, difficulty is to be 
expected, because the same reactions can link together two amino acids of the 
same kind. 

o o 

II PCb II 

NH 2 CH 2 C-OH ► NH 2 CH 2 C-C1 

O O O 

II -HCl II II 

2NH 2 CH 2 C-C1 ► NH 2 CH 2 CNHCH 2 C-C1, etc. 

To avoid such reactions, a " protecting group " is substituted on the amino 

function of the acid that is to act as the acylating agent. A good protecting 

group must be easily attached to the amino group without causing racemiza- 

tion of an optically active acid. Moreover, it must be easily removed without 

affecting the peptide linkages or sensitive functional groups on the amino 

acid side chains, such as the sulfur-containing groups of cysteine, cystine, and 

methionine. O 

II 
The benzyloxycarbonyl group, C 6 H 5 CH 2 OC— , is a useful protecting 

group that can be removed at the appropriate time by catalytic hydro- 

genolysis. 

o o 

II II 

C 6 H 5 CH 2 OC-NHCHC-NHCHC0 2 H 
R R 



H 2 



o 

© II 

C 6 H 5 CH 3 + C0 2 + NH 3 CHC-NHCHC0 2 e 
R R 

Another protecting group that has been found to be particularly useful in 
solid-phase peptide synthesis, described in the next section, is the ?-butoxy- 



chap 17 amino acids, proteins, and nucleic acids 472 

o 

ii 

carbonyl group, (CH 3 ) 3 COC— . Treatment of an amino acid such as glycine 
with ?-butoxycarbonyl azide [3] in basic solution produces the protected 
amino acid [4] after acidification. 

O O 

II l.base II 

(CH 3 ) 3 COCN 3 +NH 2 CH 2 C0 2 e >• (CH 3 ) 3 COC-NHCH 2 C0 2 H 

2, H© 

[3] [4] 

The product dissolved in methylene chloride can then be coupled with a 
second amino acid, such as alanine, by treatment with dicyclohexylcarbodi- 
imide, C 6 H 11 N=C=NC 6 H U , an efficient, nonacidic dehydrating agent. 1 

o 

II e 

(CH 3 ) 3 COC-NHCH 2 C0 2 H + NH 3 CHC0 2 e + C 6 H ll N=C=NC 6 H 1 i 

I 
CH 3 

O O O 



>■ (CH 3 ) 3 COC-NHCH 2 C-NHCHC0 2 H + C 6 H! jNH-C-NHCeHj t 

CH 3 

The chain can then be extended by treating this dipeptide with a third amino 
acid, and so on. Cleavage of the protecting group, which occurs readily on 
treatment with hydrochloric acid dissolved in acetic acid, gives the polypep- 
tide. 



O o w 

(CH 3 ) 3 COC-NHCH,C — (CH 3 ) 2 C = CH 2 + C0 2 + NH 3 CH 2 C- 



C. SOLID-PHASE PEPTIDE SYNTHESIS 

The difficulty with the method of peptide synthesis described in the previous 
section is the number of steps involved in order to make a peptide of even 
modest size. We have shown earlier how poor the overall yield in a multi- 
stage process can be if each stage is not highly efficient. 

In recent years, R. B. Merrifield has developed an approach to peptide 
synthesis which permits preparation of peptides of some size (and even 
enzymically active substances) in acceptable yields. The advantage of the 
method is that it avoids the intermediate isolation steps which characterize 
the usual multistage synthesis. At the start, the amino acid that will ulti- 
mately be at the free-carboxyl end of the peptide is bonded to active groups 
on a solid polymer through its carboxyl group. The polymer is usually an 
insoluble polystyrene resin containing chloromethyl groups and is pre- 
pared as small spherical beads. The binding reaction is brought about by 
nucleophilic displacement between the chloride and the salt of the amino 
acid to give an ester. 

1 Great care should be exercised when working with this compound — it is a potent 
contact allergen. 



sec 17.3 peptides 473 

NH 2 CHC0 2 e +C1CH 2 - [POLYMER] 
I 
Rx 

O 
II 
► NH 2 CHC-0-CH 2 -[POLYMER]+Cl e 

Ri 

The resulting solid, insoluble, amino acid-polymer complex is quite 
porous and, when treated with a methylene chloride solution of a second 
amino acid (suitably protected at the amino group by a ?-butoxycarbonyl 
group) and dicyclohexylcarbodiimide, yields a dipeptide, attached at one end 
to the solid resin. 

o o 

II II 

(CH 3 ) 3 COC-NHCHC0 2 H + NH 2 CHC-0-CH 2 - [POLYMER] 
R 2 Ri 



C 6 HiiN=C=NC 6 Hi 
CH2CI2 



OOO 

II II II 

(CH 3 ) 3 COC-NHCHC-NHCHC-0-CH 2 -[POLYMER] 
Ra Ri 



The protecting group can be removed, as described in the previous section, 
and the process continued. The advantage of the procedure is that the prod- 
ucts at each stage need not be isolated and purified as would normally be 
required for reactions in solution. Instead, the solid resin with its attached 
peptide chain can be washed free of impurities and excess reagents at each 
stage with virtually no loss of peptide. The method lends itself beautifully to 
automatic control, and machines suitably programmed to add reagents and 
wash the product at appropriate times have been developed. At present, the 
chain can be extended by six or so amino acid units per day. 

When the synthesis of the peptide chain has been completed, it is removed 
from the resin surface with hydrogen bromide in anhydrous trifluoroacetic 
acid. This treatment removes the final N-protecting group at the same time. 

00 o 

II II II 

(CH 3 ) 3 COC-NHCHC NHCHC-0-CH 2 -[POLYMER] 

R* Ri 



HBr 



O 



CF 3 C0 2 H 



NH 3 CHC NHCHCO, + (CH 3 ) 2 C=CH 2 + C0 2 + BrCH 2 -[POLYMER] 

K R, 



chap 17 amino acids, proteins, and nucleic acids 474 



• 


h: 


• 


,N« 


• C 


r\ , 


• // 


H C. 


H P ,V 

A; J~\ 




." 


ii cr 


• 


9 7 1 


H 


: :c 

: *<5 


' V 


: h /\ 

: >?^/ 
h *<£ : w 


V W H 


• 


C • ^ 

ii ' 


e 


9 • 


\ N 


: H » r— 




\\: 




o *. 




^NH---0 hydrogen 
bonds 



Figure 17-7 Peptide chain of a protein coiled to form an a helix. Configura- 
tion of the helix is maintained by hydrogen bonds, shown as vertical dotted 
(or solid) lines. The helix on the left shows the detailed atom structure of the 
peptide chain (the side chain groups are not shown). The helix on the right 
is a schematic representation without structural detail. 



The Merrifield method has been used to synthesize a polypeptide with 124 
amino acids arranged in the sequence present in the enzyme ribonuclease. 
After removal of the peptide from the resin, it was purified and then exposed 
to air to oxidize the SH bonds in eight cysteine units to form the four S— S 
bonds that are largely responsible for stabilizing the folded pattern of the 
peptide chains in the enzymically active state (see the next section). The prod- 
uct showed enzymic activity although it was less efficient than the natural 
enzyme, probably because minute amounts of impurities were introduced 
during the individual reactions. The synthesis involved 369 reactions and 
almost 12,000 individual operations of the automated peptide-syn thesis 
machine without isolation of any intermediates. 



17-4 protein structures 



Considerable attention has been given to the possible ways in which peptide 
chains can be arranged to give stable conformations. An especially favorable 
arrangement that is found to occur in many peptides and proteins is the 



sec 17.4 protein structures 475 

a helix. The principal feature of the a helix is the coiling of peptide chains in 
such a way as to form hydrogen bonds between the amide hydrogens and 
carbonyl groups that are four peptide bonds apart. The hydrogen bonds are 
nearly parallel to the long axis of the coil and the spacing between the turns is 
about 5.4 A (see Figure 17-7). The amino acid side chains lie outside the coil 
of the a helix. However, proteins are not perfect a helices, because steric 
hindrance between certain of the amino acid side chains or the lack of hydro- 
gen bonding is sometimes sufficient to reduce the stability of the normal 
a helix and allow the chain to fold. 

Various levels of complexity therefore exist in any one protein structure. 
The primary structure is the specific sequence of amino acids in the polypep- 
tide chain; the Secondary structure is the way the chain is coiled, often to 
form an a helix; and the tertiary structure is the way in which the coiled 
chain(s) is folded and hydrated in the natural state. Some proteins, in fact, 
have a quaternary structure as a result of the grouping together of two or 
more of these folded units. Hemoglobin, for example, is a tetramer and 
insulin a hexamer. Ionic forces, van der Waals forces, and hydrogen bonds 
(but not covalent bonds) are responsible for quaternary structure. 

The primary structure of the antidiabetic hormone insulin was elucidated 
in 1950 by Sanger, work for which he received the Nobel Prize. The sequence 
of the 51 amino acids in the peptide chain of beef insulin is shown in Figure 
17-8. Sheep, horse, and hog insulin all have a slightly different arrangement 



Figure 17-8 Amino acid sequence in beef insulin. In agreement with con- 
vention the amino acids on the left-hand side of the chain have free amino 
groups and the other two terminal amino acids have free carboxyl groups. 




chap 17 amino acids, proteins, and nucleic acids 476 

of amino acid residues at one location in the molecule, but this does not affect 
the hormone's physiological function, probably because substitution at this 
location does not greatly affect the higher structures of the protein. Insulin 
from man is identical with pig insulin except that threonine replaces alanine 
at the very end of one of the chains. 

The solution to the problem of the quaternary structure of crystalline insu- 
lin has only recently been, announced, almost 50 years after Banting and Best 
first isolated insulin and 20 years after Sanger established its primary structure. 
The English chemist Dorothy Crowfoot Hodgkin (already a Nobel Prize 
winner for earlier work) has shown by X-ray diffraction studies of crystalline 
insulin that the hormone is a roughly triangular ring made up of six polypep- 
tide chains (such as that shown in Figure 17-8), two to each side of the triangle. 
Two zinc atoms are complexed to the inner side of the ring. 

The bewildering and almost random-appearing sequences of amino acids in 
the peptide chains of proteins such as insulin probably have more than one 
role in influencing the higher structures of proteins and hence their properties. 
In the first place, the electrical behavior of proteins and their isoelectric points 
are determined by the number and location of acidic and basic amino acids. 
Second, the steric effects of substituent groups determine the stabilities and 
positions of folds in the peptide helices. The nature of the amino acid se- 
quences may also influence the degree of intermolecular interactions and pro- 
tein solubilities. Peptides of only one kind of amino acid are often highly 
insoluble, partly because of strong intermolecular forces. If the regularity of 
the chain is broken by having different amino acids in it, the intermolecular 
forces should diminish. 

Proteins are found to occur in a very wide variety of sizes and shapes. 
Determination of the molecular weights and dimensions of proteins has been 
made with the aid of an impressive array of physical techniques. Molecular 
weights can be obtained by analysis for- particular constituents (see Exercise 
17T3), determination of diffusion rates, sedimentation velocities in the ultra- 
centrifuge, light scattering, and even measurements of the sizes of individual, 
very large protein molecules by electron microscopy. The shapes are deduced 
from measurements of rates of molecular relaxation after electric polarization, 
changes in optical properties (double refraction) resulting from streaming in 
liquid flow, directly by electron microscopy, and perhaps most importantly 
by the intensities of light or X-ray scattering as a function of scattering angle. 
The application of all these methods is often rendered difficult by the high 
degree of hydration of proteins and by the fact that many proteins undergo 
reversible association reactions to give dimers, trimers, and so on. The mole- 
cular weights, molecular dimensions, and isoelectric points of a few important 
proteins are compared in Table 17-2. 

It is found that many proteins contain metals such as iron, zinc, and copper, 
and these metal atoms turn out to be intimately involved in the biological 
functions of the molecules to which they are bound. The well-known oxygen- 
carrying property of hemoglobin and the hemocyanins is a case in point. 
These molecules have metal-containing heme rings as the prosthetic groups 
attached to the polypeptide chain. The cytochrome enzymes that are in- 



sec 17.6 the structure of DNA 477 

volved in biological oxidation (Chapter 18) are constructed similarly. Like 
many other substances we call proteins they are more than simple polymers 
of amino acids. 

The biological functions of proteins are extremely diverse. Some act as 
hormones regulating various metabolic processes (e.g., insulin is responsible 
for the control of blood sugar levels); some act as catalysts for biological 
reactions (enzymes), and others as biological structural materials (e.g., col- 
lagen in connective tissue and keratin in hair). The oxygen-carrying properties 
of hemoglobin in mammals (and the copper-containing proteins called hemo- 
cyanins, which function similarly for shellfish) have been mentioned already. 
Some blood proteins function to form antibodies, which provide resistance 
to disease, while the so-called nucleoproteins are important constituents of 
the genes that supply and transmit genetic information in cell division. The 
viruses, such as tobacco mosaic virus, are made up of nucleoproteins, nucleic 
acids encased in a protein "coat." The structures of many viruses are so 
regular that they can be obtained in nicely crystalline form. Viruses function 
by invading the cells of the host and by supplying a genetic pattern that de- 
stroys the normal functions of the cells and sets the cellular enzymes to work 
synthesizing more virus particles. 



11 -S biosynthesis of proteins 



One of the most interesting and basic problems connected with the synthesis 
of proteins in living cells is how the component amino acids are induced to 
link together in the sequences which are specific for each type of protein. 
There is also the related problem of how the information as to the amino acid 
sequences is perpetuated in each new generation of cells. We now know that 
the substances responsible for genetic control in plants and animals are pres- 
ent in and originate from the chromosomes of cell nuclei. Chemical analysis 
of the chromosomes has revealed them to be composed of giant molecules of 
deoxyribonucleoproteins, which are deoxyribonucleic acids (DNA) bonded 
to proteins. Since it is known that DNA rather than the protein component 
of a nucleoprotein contains the genetic information for the biosynthesis of 
enzymes and other proteins, we shall be interested mainly in DNA and will 
first discuss its structure. Note that part or perhaps all of a particular DNA 
is the chemical equivalent of the Mendelian gene — the unit of inheritance. 



17-6 the structure of DNA 



The role of DNA in living cells is analogous to that of a punched tape used for 
controlling the operation of an automatic turret lathe. DNA supplies the in- 
formation for the development of the cells, including synthesis of the necessary 
enzymes and such replicas of itself as are required for reproduction by cell 
division. Despite the enormous differences in gross structure of the many 



name 


mol. wt. 


shape 


isoelectric 
point 


occurrence 


function 


insulin"' 6 


5800 c 




5.4 


pancreas 


regulation of blood 
sugar levels 


ribonuclease"' 6 


13,000 




7.8 


pancreas 


hydrolysis of ribonucleic 
acids 


myoglobin (horse)"' 6 


17,500 


platelets 




horse heart 


respiratory protein 


lysozyme 0,6 


14,600 


prolate 
ellipsoid 


10.7 


egg white 


breaks down the cell 
walls of bacteria 
by hydrolysis of 
/3(1 -» 4) glucose 
linkages 


a-chymotrypsin"' 6 


25,000 




8.4 


pancreas 


hydrolysis of ester 
and peptide linkages 


papain" 


21,000 




8.8 


latex of papaya 
melons 


hydrolysis of peptide 
linkages 


hemoglobin 0,6 


64,450 


nearly 
spherical 


6.7 


red-blood 
corpuscles 


respiratory protein 


catalase 


250,000 


blocks" 




liver and kidney 


destroys H 2 2 


fibrogenin 


330,000 


elongated 


6.8 


blood plasma 


blood clotting 


tobacco mosaic virus 
(protein part)" 


41,000,000 


hexagonal 
prisms or 
rods' 1 


4.1 


infected tobacco 
or tomato plants 


plant virus 



H 

wa 
O 

3 
a 

rt- 

n* 

~a 

n 
O 

2 

5' 



" Complete sequence of amino acids has been established. 

6 Structure investigated by X-ray diffraction methods. 

c Molecular weight of monomeric insulin ; crystalline insulin consists of six such units complexed with two zinc atoms. 

d From electron-microscope photographs. 



sec 17.6 the structure of DNA 479 

varieties of plants and animals the basic structural features of DNA from all 
sources are surprisingly similar. We shall be mainly concerned with these 
basic features in the following discussion. 

In the first place, DNA molecules are quite large — sufficiently so that they 
can be seen individually in photographs taken with electron microscopes. The 
molecular weights vary considerably, but values of 1,000,000 to 4,000,000,000 
are typical. X-ray diffraction indicates that DNA is made up of two long- 
chain molecules twisted around each other to form a double-stranded helix 
about 20 A in diameter. 



.M 



w 



„bS# 



-20A 



#*^ 






As we shall see, the components of the chains are such that the strands can 
be held together efficiently by hydrogen bonds. In agreement with the pro- 
posed structure, it has been found that, when DNA is heated to about 80° 
under proper conditions, the strands of the helix untwist and dissociate to two 
randomly coiled fragments. Furthermore, when the dissociated material is 
allowed to cool slowly (again, under proper conditions), the fragments recom- 
bine and regenerate the helical structure. 

Chemical studies show that the strands of DNA have the structure of a 
long-chain polymer made of alternating phosphate and nitrogen base-sub- 
stituted sugar residues [5]. The sugar is D-2-deoxyribofuranose [6], and each 



— phosphate 



sugar 



phosphate — | sugar — ] phosphate | — sugar 



base 



base 



[5] 



sugar residue is bonded to two phosphate groups by ester links involving 
the 3- and 5-hydroxyl groups. 
-Ox 



HOHX 




OH 



HO 

2-deoxyribofuranose 
[6] 



chap 17 amino acids, proteins, and nucleic acids 480 

The backbone of DNA is thus a polyphosphate ester of a 1,3-glycol. 



OH OH OH 

I III I III I III 

-O-P-O-C-C-C-O-P-O-C-C-C— O-P— o-c— c-c-o— 

II III II III II III 

O o o 



With the details of the sugar residue included, the structure of DNA be- 
comes as shown in Figure 17-9. 

Each of the sugar residues of DNA is bonded at the 1 position to one of 
four bases, adenine [7], guanine [8], cytosine [9], and thymine [10], through 
an N-glycosidic linkage. The four bases are derivatives of either 2-hydroxy- 
pyrimidine or purine, both of which are heterocyclic nitrogen bases. 



/N 



OCX 

H 



purine 



N 
2-hydroxypyrimidine 




adenine 
[V] 



NH 




cytosine 
[9] 



guanine 



OH 
H 3C-^W N 

^N^OH 

thymine 
[10] 



For the sake of simplicity in illustrating N-glycoside formation in DNA, we 
shall show the type of bonding involved for the sugar and base components 
only (i.e., the nucleoside structure). Attachment of 2-deoxyribose to the 
purines, adenine and guanine, is easily envisioned as involving the NH group 
of the five-membered ring and the C-l of the deoxyribofuranose ring, the 
union always being /?. 



OH 

2-deoxyribofuranose 



NH, 




-H 2 



HOH,C 



NH 2 




N 
J 



HO 



adenine deoxyriboside 

(a nucleoside) 

03) 



However, an analogous process with the pyrimidines, cytosine and thymine, 
has to involve tautomerization of the base to an amide-type structure. 



sec 17.6 the structure of DNA 481 



O 
I 
HO-P-0-CH 2/ .0^ R 

O 




O 
I 
HO — P-O— CH 2 (X r 



O 




R = nitrogen base ; 
adenine, guanine, 
cytosine, or thymine 



O 



HO-P-0-CH 2 /0^ R 
O 




O 



Figure 17-9 Structure of the strands of deoxyribonucleic acid (DNA). 



NH 2 

Cx 



N^^OH 



cytosine 



NH 2 

N 2-deoxyribose 

H 



NH 2 



-H 2 



HO-H,C 




i 



HO 



cytosine deoxyriboside 
(a nucleoside) 

00 



Esterification of the 5-hydroxyl group of deoxyribose nucleosides, such as 
cytosine deoxyriboside, with phosphoric acid gives the corresponding 
nucleotides. 



O 

II 

HO — P-O— H,C 
OH 



.O- 




NH 2 

-4 N 



Ni O 



OH 



cytosine deoxyribonucleotide 
(a nucleotide) 

Thus DNA may be considered to be built up from nucleotide monomers 
by esterification of the 3-hydroxyl group of one nucleotide with the phosphate 



chap 17 amino acids, proteins, and nucleic acids 482 



H 


CH 3 


,HN^^ 


„ /°v\ 


o vl T 1 


HNH 1 1 


11 »N N^ 


1 V HN N 


.N^v^wu' II deoxyribose 
\ 1 1 ..-° 


F^Ti^N 1 '' If deoxyribose 
\ J. J ° 


deoxyribose 


deoxyribose 


guanine-cytosine 


adenine-thymine 



Figure 17-10 Hydrogen bonding in the base pairs guanine-cytosine and 
adenine-thymine, which leads to the two strands of DNA being linked. In 
each case the distance between C-l of the two deoxyribose units is 11 A and 
the favored geometry has the rings coplanar. 



group of another. Enzymes are available which hydrolyze DNA to cleave the 

linkage at C-3 and give the 5-phosphorylated nucleotides. There are other 

enzymes which cleave DNA at C-5 to give the 3-phosphorylated nucleotides. 

The number of nucleotide units in a DNA chain varies from a few thousand 



Figure 17-11 Schematic representation of configuration of DNA, showing 
the relation between the axes of hydrogen-bonded purine and pyrimidine 
bases and the deoxyribose-phosphate strands. There are 10 pairs of bases per 
complete 360° twist of the chain. The spacing between the strands is such that 
there is a wide and a narrow helical " groove " which goes around the mole- 
cule. Apparently in the combination of DNA with protein, the protein is 
wound around the helix filling one or the other of the grooves. 



-hydrogen 
bond 



axis of adenine-thymine 
or guanine-cytosine pair 



- deoxyribose-phosphate 
chain 



sec 17.7 genetic control and the replication of DNA 483 

to well over a million. A single DNA molecule can be isolated from the 
bacterium Escherichia coli which has a molecular weight of 2 x 10 9 and whose 
extended length is almost a millimeter. 

Although the sequences of the purine and pyrimidine bases in the chains 
are not known, there is a striking equivalence between certain of the bases 
regardless of the origin of DNA. Thus the number of adenine groups equals 
the number of thymine groups, and the number of guanine groups equals the 
number of cytosine groups (i.e., A = T and G = C). Also, the overall percent- 
age composition of the bases is constant in a given species but varies widely 
from one species to another. 

The equivalence between the purine and pyrimidine bases in DNA was 
accounted for by Watson and Crick (1953). They were the first to realize that 
if two strands are twisted together to form a double helix, hydrogen bonds 
can form between adenine in one chain and thymine in the other or cytosine 
in one chain and guanine in the other. Thus, each adenine occurs paired with 
a thymine and each cytosine with a guanine and the strands are said to have 
complementary structures. The hydrogen bonds that form the base pairs are 
shown in Figure 17-10, and the relation of the bases to the strands in Figure 
17-11. 



17 -7 genetic control and the replication of DNA 

It is now clear that DNA provides the genetic recipe that permits cell division 
to produce identical cells. In reproducing itself, it perpetuates the information 
necessary to regulate the synthesis of specific enzymes and other proteins of 
the cell structure. The genetic information inherent in DNA depends on the 
arrangement of the bases (symbolized as A, T, G, and C) along the phosphate- 
carbohydrate backbone — that is, on the arrangement of the four nucleotides 
specific to DNA. 

There are 20 amino acids that become joined together to form proteins but 
only four nucleotide bases. A single base, therefore, cannot correspond to a 
single amino acid and even a pair of adjacent bases gives only 4 2 or 16 com- 
binations. A triplet of bases, however, provides 4 3 or 64 combinations, more 
than enough to store the information required for the sequential addition of 
20 different amino acids to form a protein of specific structure. Each of the 
64 three-letter code words, called codons (for example, AGC, adenine-guan- 
ine-cytosine), corresponds to an amino acid and it is apparent that an indi- 
vidual amino acid may have several codons that direct its addition to the 
growing polypeptide chain. 

The base sequence in DNA can be modified chemically by treatment of 
DNA in vitro (outside the cell) or in vivo (inside the cell) with nitrous acid, 
which can convert the primary amino groups of adenine, cytosine, and 
guanine to OH groups. This clearly changes the genetic message, since DNA 
modified this way leads to mutations in the organisms from which it was 
originally obtained. A drastic change may occur when the DNA of a bac- 
teriophage (which is no more than a bundle of DNA enclosed in a protein 



chap 17 amino acids, proteins, and nucleic acids 484- 

coat) is introduced into a bacterium. The bacteriophage DNA acts as a primer 
for the synthesis of DNA and proteins of its own kind, finally causing dis- 
solution of the host cell and liberation of new bacteriophage particles. 

Other important experiments which indicate that a given organism manu- 
factures DNA of its own kind are based on the synthesis of DNA in vitro. 
A mixture of the four DNA nucleotides, A, G, C, and T, with triphosphate 
groups in the 5 position, can be polymerized to DNA in the presence of the 
enzyme DNA-polymerase, magnesium ions, and a DNA primer. The last can 
come from a variety of sources, but the synthetic DNA has the composition 
of the primer DNA, even if the relative amounts of the nucleotides that are 
supplied are varied. The role of magnesium is not clear, but it behaves as a 
type of inorganic coenzyme, since the enzyme apparently does not function 
in its absence. The triphosphate grouping on the nucleotides is necessary as 
a source of energy to drive the reaction forward; some 7 kcal/mole is 
liberated in the cleavage of the triphosphoryl group to pyrophosphate (see 
Section 15-5). 



OH OH 

1 1 


OH 






n-HO— P— O — P— O— P-O— H 2 C 

ii ii ii k 

O O O ' 




> R 






T 
OH 




DNA- 


Mg 2 ®, 






poly- 


"DNA 






merase 


primer" 






~ H ^ ^ 




O 
II 
-P— O- 

1 


O 


\_yOH 


+ n-HO- 


II 
-P-OH 

OH 


O P 

II 
O 


n 


OH 



The mechanism of replication of DNA that takes place when the cell divides 
probably involves the unwinding of the DNA double helix into two comple- 
mentary chains. During the unwinding process, each chain serves as a template 
on which is built the complement of itself. 

The genetic information of cells is stored in DNA but protein synthesis 
takes place on small subcellular particles called ribosomes. How does the in- 
formation stored in DNA become translated into a protein molecule in the 
ribosome ? This is accomplished through the intermediacy of ribonucleic acid 
(RNA), a substance similar in structure to DNA 2 but containing ribose in- 



2 Although DNA was first located in cell nuclei of plants and animals (hence the name 
nucleic acid), it is now known to be present outside the nucleus, as well, that is, in the 
cytoplasm. Indeed, bacterial cells, which contain no nuclei, use the same system of infor- 
mation transfer (DNA->RNA-> protein) as do nucleated cells. 



sec 17.7 genetic control and the replication of DNA 48S 

stead of deoxyribose (see Section 15-5). RNA contains the base sequence 
transcribed from DNA, but with the base uracil [11] replacing thymine [10]. 
In contrast to DNA most RNA molecules are single stranded. 



H,C. 



OH 

I 


OH 

i 


I k 
V N-^OH 


ll N 
^•N^OH 


thymine 


uracil 


[10] 


[11] 



There is a striking variation in size of RNA molecules and this seems to 
depend on the particular role they play. One class of RNA molecules is called 
transfer RNA (tRNA). The various molecules contain 75 or so nucleotide 
units and have molecular weights of about 25,000. Specific tRNA molecules 
are esterified enzymically with the corresponding amino acid and the amino 
acids then are transferred to the growing protein chain on the ribosome in a 
complex series of enzyme-catalyzed reactions. Because tRNA is the smallest 
and most soluble form of RNA, it is sometimes called soluble RNA (sRNA). 

Another kind of RNA with a somewhat bigger molecular weight is called 
messenger RNA (mRNA). It carries the message from the DNA in the nucleus 
to the tRNA at the ribosome. Its base sequence is complementary to a portion 
of one strand of DNA and consequently it contains the sequence of codons 
defining the sequence of amino acids in the protein. 

The codon(s) for a particular amino acid can be determined by experiment. 
For example, a synthetic RNA containing only one type of base, uracil, acts 
as the mRNA for a cell-free preparation of the protein-synthesizing system of 
the bacterium E. coli. The product is a polypeptide containing only phenylala- 
nine. This suggests that the codon for phenylalanine is UUU. The messages 
delivered by all 64 codons are now known, at least for E. coli, and it is 
probable that the code is universal. 

In some cases, the codon was determined by synthesis of the simple tri- 
nucleotide molecule made up of the appropriate triplet. This is often sufficient 
to cause amino acyl-tRNA to be bound to the ribosome. For example, the 
trinucleotide UUU causes only phenylalanyl tRNA, and the trinucleotide 
UCU only seryl tRNA, to be bound. Other trinucleotides are much less 
efficient in promoting binding, however, making it difficult to determine the 
amino acid to which they correspond. In these cases, the complex polynucleo- 
tide having the correct base sequence had to be synthesized and its effect on 
amino acid incorporation in polypeptides observed. 

It turns out that only 61 of the 64 codons correspond to amino acids. The 
other three, UAA, UAG, and UGA, are so-called "nonsense" codons. 
Despite this name their message is clear — they stop the growth of the poly- 
peptide chain. 

A third class of RNA is called ribosomal RNA (rRNA). These molecules 
are components of ribosomes but their roles are not as well understood as 
those of tRNA and mRNA. 



chap 17 amino acids, proteins, and nucleic acids 486 

Phenomenal progress has been made in the past 20 years in understanding 
the chemical basis of heredity and we can expect in the future to learn that 
manipulation of human genetic material can be done on a clinical basis. It 
remains to be seen whether or not we possess the wisdom to use such knowl- 
edge in a wholly beneficial way. 



17-8 chemical evolution 

The term " chemical evolution " is used to refer to those events occurring on 
the primitive Earth that led to the appearance of the first living cell. There is 
general agreement that the Earth was formed about A\ billion years ago by 
condensation from a dust cloud. Fossil evidence shows that unicellular orga- 
nisms (protozoa) existed on Earth at least 3 billion years ago, which leaves 
possibly a billion years for chemical evolution to produce the complex organic 
molecules that are the components of a living cell. 

There is good reason to believe that the atmosphere of the prebiotic Earth 
was hydrogen dominated, rather than oxygen dominated as at present, and 
the principal atmospheric constituents were probably methane, ammonia, and 
water vapor. The change from a reducing to an oxidizing atmosphere may 
have been caused by radiolysis of water vapor by ultraviolet radiation fol- 
lowed by escape of hydrogen from the planet's atmosphere. This allowed the 
ozone shield to develop in the upper atmosphere — the protection required by 
cells from the lethal effects of high-energy ultraviolet light. 

Experiments have been conducted in the laboratory which simulate 
"natural" but prebiotic organic syntheses. Application of ultraviolet radia- 
tion and electric discharges (the analogs of unshielded sunlight and lightning 
storms, respectively) to mixtures of CH 4 , NH 3 , H 2 0, N 2 , and H 2 produces 
both amino acids and nucleic acid bases. Other kinds of compounds, too, are 
formed but it is interesting that the naturally occurring amino acid 2-amino- 
propanoic acid (alanine) is always formed in larger amounts than its isomer 
3-aminopropanoic acid. 

Adenine, one of the nucleic acid bases, has the formula C 5 H 5 N 5 , or 
(HCN) 5 , and it has been shown that adenine is one of the compounds formed 
when aqueous solutions of ammonia and hydrogen cyanide (products of 
CH 4 -N 2 irradiation) are heated at 90° for several days. One of several 
possible routes for this reaction begins with the stepwise trimerization of 
hydrogen cyanide to give aminomalononitrile: 

HCN 

2 HCN >• [HN=CHCN] > H 2 NCH(CN) 2 

dimer aminomalononitrile 

This compound, which has been detected in the reaction mixture, is known to 
be able to condense with one mole of formamidine (formed by addition of 
ammonia to hydrogen cyanide) to give the imidazole [12]. The imidazole [12] 



summary 487 



NH 

// 

HCN + NH, ► H— C 

\ 
NH, 



formamidine 



NQ ^NH, 



rH HN NC XT 



1 + C-H > II > + NH 3 

H,N 

N 



>-. T^ V^ II I 

Jl H 2 N H 2 N X N' 



[12] 



can react with a second mole of formamidine to give adenine. All of the 



NH 2 

NH 2 NC V- N \ N^V N 



J N > — L,l 



W % NH H 2 N^N N ^ 

[12] adenine 



NH 3 



above compounds except the HCN dimer have been identified in the aqueous 
solution. There is little doubt that many of the units of biopolymers were 
present under prebiotic conditions and it is not difficult to visualize the for- 
mation of proteins and nucleic acids taking place also. There is still, of course, 
an enormous gap in our understanding of how these substances became 
organized into the first living cell able to replicate itself. 



summary 

Of the 24 a-amino acids that are important constituents of proteins, all except 
glycine, NH 2 CH 2 C0 2 H, are optically active. Eight of them are essential to 
human nutrition in the sense that the body cannot synthesize them. Four of 
the remainder are formed by conversion of other amino acids only after the 
protein has been synthesized. 

Amino acids can be prepared from a-halo acids or by the Strecker synthesis 
(RCHO -► RCHNH 2 CN -+ RCHNH 2 C0 2 H). 

Amino acids with one carboxyl and one amino group (neutral amino acids) 
exist in the dipolar (zwitterionic) form, RCHC0 2 e , which is converted by 

NH 3 ® 

acids to a cation RCHCO,H, and by bases to an anion RCHCO, 6 . 
I I 

NHj® NH 2 

The isoelectric point is the pH at which the concentration of the zwitterion 
is at a maximum and is near pH 6 for many simple amino acids. Basic amino 
acids such as lysine (NH 2 CH 2 CH 2 CH 2 CH 2 CHNH 2 C0 2 H), which have an 
excess of amino over carboxyl groups, have isoelectric points near pH 9, 



chap 17 amino acids, proteins, and nucleic acids 488 

while acidic amino acids such as glutamic acid (HOjCCHjCHjCHNHjCOjH) 
with an excess of carboxyl over amino groups have isoelectric points near 
pH3. 

Analysis and detection of amino acids can be carried out with nitrous acid 
(Van Slyke method), by the ninhydrin color test, or by chromatography. 
Paper chromatography involves the partitioning of compounds caused by 
differences in their rates of migration on the surface of moist filter paper. 
These are expressed as Rf values, the distances that compounds migrate rela- 
tive to that of the solvent front. Thin-layer chromatography (tic) is similar 
except that the filter paper is replaced by a thin layer of absorbent, such as 
silica gel, on glass plates. 

Cyclization of esters of y- and (5-amino acids (amino at C-4 or C-5) produces 
lactams (cyclic amides) whereas a-amino esters form six-membered rings by 
dimeric cyclization. 

Peptides (polypeptides) are poly amides formed by the condensation of 
a-amino acids. Synthesis of peptides in the laboratory normally requires the 

O 

II 
use of a protective group, such as ROC — , on the amino group of the amino 

acid that is being linked to the chain to prevent it from reacting with its own 

carboxyl function. 

o o o o o 

II II II II II 

ROC-NHCHC0 2 H + NH 2 CHCNH~~~ > ROC-NHCHC-NHCHC-NH- 

II II 

Rj R-2 Ri R^ 

Formation of the peptide linkage can be brought about by mild dehydrating 
agents, such as RN=C=NR, or via the acid chloride. The protective group 
can be removed and the process continued. Solid-phase peptide synthesis can 
be achieved by attaching the first amino acid in the chain to a polystyrene 
resin and then using this granular material in all subsequent chemical and 
washing operations. In the final step, the peptide is liberated from the resin 
by a suitable reaction. 

Proteins are peptides with special structural features beyond the primary 
structure, which refers to the sequence of amino acids in the chain. Proteins 
have a secondary structure, which refers to the way the chain is coiled; a 
tertiary structure, which refers to the manner of folding of the coiled chain; 
and a quaternary structure, which refers to the association of the folded units 
by noncovalent links. 

Many proteins, such as hemoglobin, contain non-amino acid functions 
called prosthetic groups. Viruses consist of protein-nucleic acid combinations. 

Deoxyribonucleic acid (DNA) is a polymer of 2-deoxyribose and phos- 
phoric acid, with nitrogen bases (adenine, guanine, cytosine, and thymine) 
attached to the sugar units. The combination of an individual sugar and a 
base is called a nucleoside; the combination of a sugar, a base, and phosphoric 
acid is a nucleotide. Coded genetic information is contained in the sequence 
of bases along the polymer chains, which are coiled around one another in 
the form of a double-stranded helix. 



exercises 489 

Ribonucleic acid (RNA) differs from DNA in the following respects: it 
contains ribose rather than deoxyribose; it contains uracil rather than thy- 
mine ; and it has a much greater variation in molecular weight (corresponding 
to the three different functional types, tRNA, mRNA, and rRNA). The 
genetic information coded in the base triplets of DNA becomes manifest in 
protein synthesis through the action of RNA. 

The appearance of life on Earth was presumably preceded by the synthesis 
of rather complex organic molecules. The effects of radiation on the highly 
reduced atmosphere believed to exist in prebiotic times can be duplicated in 
the laboratory and shown to give rise to compounds such as amino acids and 
nucleic acid bases. 



exercises 

17-1 Pick out the amino acids in Table 17-1 which have more than one asymmetric 
center and draw projection formulas for all the possible stereoisomers of 
each which possess the l configuration for the a carbon. 

17-2 Which of the amino acids in Table 17-1 are "acidic" amino acids and which 
"basic" amino acids? Which of the structures shown would have the most 
basic nitrogen? The least basic amino nitrogen? The most acidic and least 
acidic carboxyl group ? Give the reasons for your choices. 

17-3 Show the sequence of steps involved in the Strecker synthesis of a-amino 
acids. 

17-4 Show how the following amino acids can be prepared from the indicated 
starting materials by the methods described above or earlier. 

a. leucine from 2-methyl-l-propanol 

b. lysine from 1 ,4-dibromobutane 

c. proline from hexanedioic acid (adipic acid) 

d. glutamic acid from a-ketoglutaric acid 

17-5 Devise physical or chemical ways to determine directly or calculate the 
equilibrium constant between neutral glycine and its dipolar ion. Arguing 
from substituent effects on the ionization of carboxylic acids and amines, 
would you expect the equilibrium constant to be closer to 0.1, 1.0, or 10? 
Explain. 

17-6 Suppose one were to titrate an equimolal mixture of ammonium chloride 
and acetic acid with two equivalents of sodium hydroxide. How would the 
titration curve be expected to differ from that of glycine hydrochloride 
shown in Figure 17-1? Take K HA for acetic acid equal to 2x 10"" 5 and 
Kbu® for ammonium ion to be 5 x 10~ 10 . 

17-7 Write mechanisms based insofar as possible on analogy for each of the steps 
involved in the ninhydrin test using glycine as an example. Would you 
expect ammonia or methylamine to give the blue color ? 



chap 17 amino acids, proteins, and nucleic acids 490 

17-8 Arrange the following amino acids in the order in which you would expect 
each to move in a paper chromatogram with the weak organic base 2,4,6-col- 
lidine as the organic phase: glycine, phenylalanine, arginine, and glutamic 
acid. Show your reasoning. 

CH, 

2,4,6-collidine 




17-9 Draw out the complete structure (using projection formulas) of the impor- 
tant hormonal peptide oxytocin. 



H • CyS • Tyr • He- Gin • Asn • CyS • Pro • Leu • Gly • NH 2 

17-10 On what theoretical grounds can we expect the C— O bonds of benzyloxy 
groups to undergo hydrogenolysis more readily than ethoxy groups ? 

17-11 Show how each of the following substances can by synthesized starting with 
the individual amino acids. 

a. glycylalanylcysteine 

b. H0 2 C(CH 2 ) 2 CH(NH 2 )CONHCH 2 C0 2 H 

c. glutamine from glutamic acid 

17-12 Most nonfatty tissue is about 80% water and 15% protein by weight with 
the remainder being made up of carbohydrates, lipids, nucleic acids, 
inorganic salts, and so on. Assuming an. average molecular weight of protein 
of 10 5 calculate the molar ratio of water to protein in such tissue. 

17-13 Hemoglobin, the protein responsible for carrying oxygen from the lungs 
to the body tissues, contains 0.355% of iron. Hydrolysis of 100 g of hemo- 
globin gives 1.48 g of tryptophan; calculate the minimum molecular weight 
of hemoglobin which is consistent with these results. 

17-14 Write equations for the steps involved in hydrolysis of adenine deoxyri- 
bonucleoside to deoxyribose and adenine. Would you expect the reaction 
to occur more readily in acidic or basic solution ? 

17-15 Escherichia coli bacteria grown in a medium containing 15 N-labeled ammo- 
nium chloride produce 15 N-containing DNA. This can be distinguished 
from ordinary 14 N-DNA by ultracentrifugation in concentrated cesium 
chloride solution — at equilibrium 14 N-DNA and 15 N-DNA form separate 
bands of differing density. When the bacteria grown in an 15 N medium 
are transferred to a 14 N medium, DNA replication continues but, after 
one generation, all the DNA present appears to be of one kind, containing 
equal amounts of 15 N and 14 N; after two generations, the DNA is now of 
two kinds present in equal amounts — all 14 N-DNA and 14 N, 15 N-DNA. 
What do these results tell about the replication of DNA and its stability in 
the cell? 



exercises 491 

17-16 Draw structures for the tautomeric forms of uracil (formula [11], Section 
17-7) and show how a nucleotide of uracil can form hydrogen bonds with 
adenine. 

17-17 Propynonitrile (cyanoacetylene), which is a product of CH 4 -N 2 irradiation, 
has been suggested as a precursor for the amino acid aspartic acid (2- 
aminobutanedioic acid) under "primitive Earth" conditions. Write a 
reaction path for this conversion making use only of NH 3 , HCN, and H 2 
as reactants. 

17-18 Write reasonable mechanisms for the cyclization steps shown in Section 
17-8 for the conversion of HCN to its pentamer, adenine. (Two of the 
reactions involved bear resemblance to reactions met earlier, the conden- 
sation of aldehydes with amines, Section 11-4D, and the addition of nucleo- 
philes to nitriles. Section 16-2B.) 










i chapter 18 
and net* 



chap 18 enzymic processes and metabolism 495 

The time scale of man's awareness ranges from about 10 10 seconds to 
possibly 10" 2 second. The former is a lifetime; the latter is the relaxation 
period of the eye. A person's reaction time — the time between observation 
and action — is measured in tenths of a second and we might well wonder 
how the complex physiological processes that result in a particular response — 
the batter's swing, the fencer's parry — can possibly occur in such brief periods 
of time. How fast can chemical reactions occur? Is a few tenths of a second 
long enough for nerve impulses to be translated into muscle action, a process 
essentially chemical in nature? 

The fastest chemical reactions that can occur in solution are those that are 
diffusion controlled. This means that, if the rate of the reaction of A and B 
depends only on their encounter rate, then every encounter produces product. 
For solvents of normal viscosity at room temperature, the rate of diffusion, 
expressed as a bimolecular rate constant, is 10 10 liters mole -1 sec -1 . This 
means that if the reactants A and B are each 1 M, the initial rate of 
production of product will be 10 10 moles liter" 1 sec -1 , a staggeringly high 
rate. Put another way, the reaction between A and B would be 99 % complete 
in 1CT 8 second. 

A number of diffusion-controlled reactions are known; one simple example 
is H® + OH e -> H 2 0. Such reactions require no activation of the reactants 
for the reaction to occur. At the other extreme are countless reactions with 
heats of activation greater than 50 kcal, which means that their half-lives (the 
time for half of the reactants to be consumed) at room temperature would be 
measured in thousands or millions of years. 

If the fastest known chemical reactions can be substantially complete in 
10~ 8 second or so at room temperature, a 10" 2 -second period does not appear 
to be quite so short, even though the chemistry of physiological processes is 
clearly much more complicated than the H® + OH® reaction. 

The study of enzymes and the way in which they accelerate and control 
biological processes is a fascinating area of study that combines the disciplines 
of organic chemistry, physical chemistry, biochemistry, and physiology. 
Because enzymes are essentially catalysts which operate by lowering the 
activation energy of what are ordinarily slow processes, we shall begin with 
a discussion of catalysis in simple organic systems. 



18-1 catalysis in organic systems 

A catalyst is a substance that will increase the rate of a reaction without itself 
being consumed. We saw in Chapter 8 that the rate of a reaction is governed 
by the energy difference between reactants and transition state, the latter 
being the highest point on the lowest energy path from reactants to products. 
Catalysis simply provides another path from reactants to products with a 
lower-energy transition state. A simple example should suffice to illustrate 
the point. 

The reaction between chloromethane and hydroxide ion to give methanol 



chap 18 enzymic processes and metabolism 496 

and chloride ion is accelerated by the addition of small amounts of iodide ion. 

ie 
CH 3 C1 + OH e ► CH3OH + Cl e 



The uncatalyzed reaction is a one-step S N 2 displacement and its rate is 
determined by the free-energy difference between the reactants and the tran- 
sition state having a configuration roughly like the following : 

H 
HO— C — CI 

(The free energy of activation, AG*, is made up of the heat of activation, AH*, 
and the entropy of activation, AS 1 *, Section 8-9.) 

The addition of iodide ion to the system provides another pathway — a two- 
stage route — to products. Iodide ion is a large polarizable ion and is both a 
good nucleophile and a good leaving group in displacement reactions. Ac- 
cordingly, it reacts rapidly with chloromethane to give iodomethane but this, 
in turn, suffers rapid displacement by hydroxide ion. The iodide ion is thus 
regenerated and is a true catalyst. 



CH3CI + I e 
CH 3 I + OH e 



CH3I + Cl e 

CH30H + I s 



The energetics of this process are shown in Figure 18-1. (The analysis of 
the activation process in terms of AH X and AS 1 * need not concern us at this 
point.) 

Figure 18-1 Energy profiles for the reaction of chloromethane with hy- 
droxide ion in the presence (solid line) and absence (dashed line) of iodide 













H n 








ge | ge 








^ — 


HO— C— CI 








-- "C 


A 








'' X 


H H 








/ \ 












H 












- ge | ge 
I— C— CI 

A 

H H 




>> 










f* 
























H 




<0 










ge | ge 
HO — C— I 

A 

H H 




CH 3 C1 + OH e 






c 


H3OH + CI e 


reaction coordinate 



sec 18.1 catalysis in organic systems 497 

The two-step reaction has two transition states that correspond to energy 
maxima, but only the higher of the two determines the rate of the overall 
reaction. If the second transition state is the higher one, the overall rate of the 
reaction will be proportional to the concentrations of the catalyst (iodide ion) 
and hydroxide ion. If the first transition state is the higher one, the rate will 
be proportional to the concentration of iodide ion and not the concentration 
of hydroxide. Thus, in either case, iodide ion is a catalyst for the overall re- 
action. Catalysts have no influence on the position of equilibrium of a system 1 
and it follows that iodide ion must catalyze the reverse reaction also. The rate 
of interconversion of chloromethane and methanol will be increased by the 
presence of iodide even though their equilibrium concentrations are unchanged. 

There are two main categories of catalysis : homogeneous catalysis, in which 
all the components are in a single phase as in the above example, and hetero- 
geneous catalysis, in which interactions occur at a solid surface. The hydro- 
genation of alkenes catalyzed by finely divided metal surfaces (Section 2-6A) 
is an example of heterogeneous catalysis. Most enzymic processes can be con- 
sidered to be homogeneous although the way in which some high-molecular- 
weight enzymes orient small reactant molecules on their surfaces bears a 
strong resemblance to the action of heterogeneous catalysts. 

The iodide ion in the example above is a nucleophilic catalyst and we shall 
see that some enzymes also operate this way. Many other organic processes 
are catalyzed by acids and bases. Since all bases are also nucleophiles it is 
sometimes difficult to tell if catalytic action is due to a molecule or ion acting 
as a base (removing a proton) or as a nucleophile (supplying the electrons to 
form a bond to carbon). Furthermore, acid catalysis and nucleophilic catalysis 
often go together, as we shall see in the following sections. 



A. ACID CATALYSIS 

A number of acid-promoted reactions have been met previously, such as the 
hydration of alkenes and dehydration of alcohols (Sections 4-4 and 10-5B), 
the formation and hydrolysis of hemiacetals (Section 11-4B), the formation 
and hydrolysis of esters (Section 10-4C), and the acid hydrolysis of amides. 
The course of these reactions is shown in Equations 18-1 to 18-4. 

ffi OH 2 

RCH=CH 2 + H® < ' RCH-CH 3 , + 2 -^ RCH-CH, 5= 

-H 2 3 



(18-1) 
O OH OH OH 

/ ~ „ // +EtOH I ffi I 

-C + H e < R-C , R-C-OEt « R-C-OEt + H® 

\ \ -EtOH I I I 

H H H H H (18-2) 



1 The only exception would be where the catalyst forms a complex with one or more of 
the reactants or products and is present in such high concentration as to change appreciably 
the concentrations from the values they would otherwise have. An effect of this kind will 
change the position of equilibrium but not the equilibrium constant. 




chap 18 enzymic processes and metabolism 498 



OH 



R — C-OEt 

EtOH | | 

OH H 



OH OH O 

■ I -HjO /; // 

R- C-OEt ^ R-C© j ► R-C + H® 

I +H2O % \ 

®OH 2 OEt OEt 

(18-3) 



,0 


OH 


+ H 2 0. 


OH OH 

1 1 e 








\ 

NH 2 


NH 2 


'-H 2 


s>OH 2 y> OH 
OH 




R— C® + NH 3 ,_ " RC0 2 H + NH 4 
OH (18-4) 









The dual role of the proton in each of the four reactions should be clear. 
At one stage, a carbon atom is activated toward an attacking nucleophile by 
protonation of the adjacent atom; at another stage, protonation increases the 
leaving ability of a substituent group. Proton transfers between oxygen atoms 
have low activation energies, and most of the protolytic equilibria shown in 
Equations 18-1 to 18-4 are fast. Despite the mechanistic similarity between 
these four processes, only the first three are examples of simple catalysis. In 
the fourth, the acid-promoted hydrolysis of amides, acid is consumed instead 
of regenerated as the reaction proceeds. Thus, although a catalytic quantity 
of acid would initiate the hydrolysis of an amide, it would not be sufficient to 
carry the reaction to completion. 

If we examine the esterification reaction as a function of the concentration 
of reactants we find that the reaction has the following kinetic form (rate = 
A;[RC0 2 H][EtOH[[H ffi ]. The rate is proportional to the first power of the 
concentration of each of the two reactants and the catalyst. If the concentra- 
tion of catalyst is doubled the rate doubles. Can the rate be indefinitely in- 
creased by addition of more and more catalyst ? Only up to a point. When 
most of the carboxylic acid that is present in the system is already in the pro- 
tonated form, further increases in acidity can have only a marginal effect on 
the concentration of protonated acid. Furthermore, protonation of the alcohol 
gives a non-nucleophilic species and therefore decreases the alcohol concentra- 
tion. Carboxylic acids and alcohols are rather weak bases and it requires a 
considerable amount of acid to reach this point. Amides, however, are less 
weakly basic, and the rate of acid-induced hydrolysis of amides reaches a 
plateau much earlier as the acid concentration is increased than for esterifi- 
cation. 



sec 18.1 catalysis in organic systems 499 



B. BASE CATALYSIS 



Base-promoted reactions already met include the aldol addition (Section 
12-2A), hemiacetal (but not acetal) formation and hydrolysis (Section 1 1-4B), 
and alkylation of ketones (Section 12-2B). The course of these reactions is 
shown in Equations 18-5 to 18-7. 



HO fc 



p 



CH, 



-+ H,0 +°CH 2 -C, 



CH 3 CHO 



I 
CH 3 -CH-CH 2 -C 



\ 



H 2 



OH 

I 

CH,— CH-CH, 



P 

H 



OH e 



(18-5) 



RO e + CH,CH,OH 5= 



ROH + CH,CH,O a 



RCH0 



O u 
I 
R— CH— OCH 2 CH 3 



/ROH 



o 

e II 

NH, + R-C-CH3 



OH 



R-CH-OCH 2 CH 3 + RO to 
O 



-» NH, 



R-C-CH 




(18-6) 



(18-7) 



For these reactions, the added base (hydroxide ion, alkoxide ion, and 
amide ion) generates an organic anion which undergoes further reaction to 
give the product. In the last reaction (18-7), however, the base is consumed as 
the reaction proceeds, so this reaction does not qualify as an example of 
base catalysis. In the other two reactions, base is regenerated in the final steps, 
so only catalytic amounts of base are required. 



C. NUCLEOPHILIC CATALYSIS 



The chloromethane-hydroxide ion reaction catalyzed by iodide ion (Section 
18T) is an example of nucleophilic catalysis. 



chap 18 enzymic processes and metabolism 500 

ie 
CH 3 Cl + OH e ► CH 3 OH + Cl e 

The catalyst forms a covalent bond to the substrate during the course of 
the reaction (CH 3 I is a discrete intermediate). This type of reaction, particu- 
larly when it involves an enzyme, is sometimes called " covalent catalysis." 

Bases are also nucleophiles and it is reasonable to ask how we distinguish 
base catalysis from nucleophilic catalysis. A reagent is a base when it re- 
moves a proton, and a nucleophile when it bonds to any atom other than H. 
We pointed out earlier that nucleophilicity does not always parallel basicity. 
A number of polarizable ions such as I e , N 3 e , HS e , and HOO e have much 
greater nucleophilicities than their basicities would indicate. 

The molecules of the heterocyclic compound imidazole (Section 25-3) are 
neutral and possess high nucleophilicity. 



HN N 
imidazole 

Although the basicity of imidazole is low (pK B = 7), it is effective in cata- 
lyzing the hydrolysis of certain esters in neutral solution. These reactions 
follow the course shown in Equation 18-8. 

— /--"^x ° — ° e — ° 

HN N: + CH 3 — C ►HNoN-C— OR ► N N-C— CH 3 + ROH 

\^ \ ^s i •%/ ... 

OR CH 3 ['] 

.0 

*3 — *" 



H2 ° CH 3 -C + N NH 
\ ^V 



R= -C„H t NO,-p OH 



The intermediate N-acetylimidazole [1] can be isolated and there is no 
doubt that the imidazole is acting here as a nucleophile rather than as a base. 

The imidazole ring is present in the amino acid histidine, which is a key 
constituent of enzymes catalyzing hydrolytic reactions. This, of course, has 
led to speculation that the enzyme operates in the same general manner as 
shown in Equation 18-8. It is doubtful, however, that the catalysis is this 
simple. Thus, in enzyme-catalyzed hydrolyses of acid derivatives, the acyl 
group often becomes attached to the alcohol hydroxyl group of the amino 
acid serine in the enzyme. Nonetheless, the imidazole ring in the enzyme plays 
an important role in the hydrolysis and we shall examine this reaction further 
in Section 18-3. 



D. GENERAL ACID AND BASE CATALYSIS 

It might be expected that the rate of an acid-catalyzed reaction in buffered 
aqueous solution would depend only on the pH of the solution and that the 
concentration of buffer would be unimportant, except possibly for a small 
salt effect. Some reactions are indeed like this. For example, the rate of the 



(18-8) 



sec 18.1 catalysis in organic systems SOI 

acid-catalyzed hydrolysis of acetals (or its reverse) depends only on the pH 
of the solution. This is called specific-acid catalysis. 

OH 

CH 3 CH(OEt) 2 +H 2 « " CH 3 CHOEt + EtO pH dependent 

(specific acid catalysis) 

Other acid-catalyzed reactions depend on the nature and the concentration 
of the buffer solution as well as on the pH. For example, the rate of the acid- 
catalyzed bromination of ketones, which takes place through the enol form, 
varies with the identity and concentration of the buffers used to control the pH. 

O OH 

II H e I 

R-C-CH 3 » R-C=CH 2 

O 

Br? , p _r_rH Rr , H Rr P H and buffer de P endent 
R C CH 2 Br + HBr ( genera l acid catalysis) 

It is difficult to see the cause of the difference without a complete kinetic 
analysis of the mechanisms in the two cases. Suffice it to say here that specific- 
acid catalysed reactions are those in which the substrate reacts in an equilib- 
rium step with a proton and then suffers a rate-controlling unimolecular 
decomposition. 

Z + H® < > ZH® 

slow 

ZH® » products 

In the case of acid-catalyzed acetal hydrolysis, the protonated acetal de- 

®,OEt 
composes slowly to the cation R — C , which then undergoes a rapid 

H 

reaction with water to give the hemiacetal. The rate now depends simply and 
directly on the pH because the concentration of ZH® depends on the pH. 
General acid catalysis results when the intermediate ion, ZH®, can only 
decompose with the help of a base or a nucleophile. 

Z +H® , ZH® 

slow 

ZH® + :B ► products 

®OH 

II 
For acid-catalyzed enolization, the protonated ketone, R— C— CH 3 , 

requires a base to remove a proton from the methyl group. The base might be 

a water molecule or a hydroxide ion (if present in significant concentrations), 

or it might be the anion A® of the acid used as buffer. The intervention of the 

base A® alters the kinetic form of the equation from one containing [H ffi ] to 

one containing [H ffi ][A®]. Because [H®][A®] is kinetically equivalent to 

[HA], 1 we will have contributions to the rate which appear to depend on all 

iH®irA®i 

1 K «* = J A1 : therefore [H®][A®] = K HA [HA] and [H®][AS] oc [HA]. 
[HA] 



chap 18 enzymfc processes and metabolism 502 

of the undissociated acids (HA, HA', etc.) which are present, provided 
that A e , A' e , and so on are effective bases in removing a proton from the 
carbon of the protonated ketone. 

A number of enzymic processes are known to be subject to general acid 
catalysis, but it is not always easy to distinguish between the action of a base 
and a nucleophile in the second step. In an enzyme, the proton source is often 
a protonated free amino group in the protein chain. The base or nucleophile 
is often an imidazole ring on a histidine unit, also in the protein chain. 

The situation with respect to base catalysis is similar and examples of both 
specific and general base catalysis are known. 



E. INTRAMOLECULAR CATALYSIS 

One of the great advantages that enzymes have in accelerating reactions is 
that their precisely coiled and folded protein chains place two or more groups 
in a position where they can simultaneously interact with the substrate(s). In 
some simple molecules, a neighboring group to the reaction site can behave 
similarly and cause rate enhancements. 

The conversion of 2-bromopropanoic acid to lactic acid can be accomp- 
lished in strong base in the normal way, with the product having the inverted 
configuration expected for an S N 2 reaction. 

CH 3 CHBrC0 2 e + OH e > CH 3 CHOHC0 2 e + Br e 

2-bromopropanoate ion lactate ion 

d series l series 

However, if the concentration of hydroxide ion is lowered by the use of 
appropriate buffers, this reaction becomes slower and is eventually superseded 
by a process that has a moderate rate, independent of pH. The resulting pro- 
duct is lactate ion as before, but it now has the same configuration as the start- 
ing material. 

The explanation is that the bromide is displaced by the neighboring car- 
boxylate group in an internal nucleophilic substitution reaction (called S N i). 
The intermediate then suffers a second displacement by a water molecule, 
which restores the original configuration. 

H 2 
Br ) OH 



CH 3 -CH-C=0 v.w 3 - 



D series L series D series 

The intervention of a neighboring group this way is called anchimeric 
assistance. Because the reaction illustrated here is accelerated by the neigh- 
boring carboxylate group, and because this group emerges from the reaction 
intact, it is, in effect, an intramolecular catalyst. 

Many other examples of intramolecular catalysis are known. The two phos- 
phonate esters [2] and [3] differ enormously in their hydrolysis rates. The 
difference is clearly due to intramolecular catalysis by the neighboring car- 
boxyl group in [2]. 




sec 18.2 enzymes and coenzymes 503 



° 9 

n ii 

P(OEt) 2 ^\.P(OEt) 2 

half-time for hydrolysis in 22 

30% aq. dimethyl sulfoxide 15 minutes > 10 years 

at 36° [2] [3] 

In nonhydroxylic solvents, molecules that can change from one tautomeric 
form to another by donating a proton at one site and accepting a proton at 
another are effective catalysts for enolization and other processes. a-Pyridone 
[4] is such a molecule because it exists in equilibrium with its tautomer [5]. 



"N-^o ^isr X)H 

H 

W [5] 

The mutarotation of tetramethylglucose in chloroform solution and a 
number of other reactions in which hydrogen shifts are important are 
catalyzed by this reagent (Equation 18-9). 



in^o 



tr T>H 



H A ► « (18-9) 






A carboxylic acid can also act as a catalyst of this type; it exists in two 

tautomeric forms that are exactly equivalent, R — C and R — C 

X OH N 

Even though Equation 18-9 shows proton shifts, these are concerted so as 
not to give free ions at any stage of the reaction. Tautomeric catalysis of the 
type shown above appears to be important only in nonhydroxylic media. In 
hydroxylic solvents, it is likely that separate protonation and deprotonation 
steps occur. 



18-2 enzymes and coenzymes 



Enzymes are invariably proteins. Some enzymes consist only of peptide chains 
and others require the presence of non-amino acid groups or molecules as 
well. If these other groups are attached directly to the peptide chain they are 
often called prosthetic groups (Section 17-3). If they are complexed to the 
enzyme in a looser fashion, they are called coenzymes. In some ways the co- 
enzyme resembles a reagent which undergoes a chemical change under the 



chap 18 enzymic processes and metabolism 504 

influence of the enzyme, just as does the substrate. The difference is that a 
coenzyme is restored to its original condition in a subsequent step. Although 
all biological oxidation systems make use of coenzymes, many hydrolytic 
systems do not. 

The most remarkable characteristics of enzymes are their catalytic effective- 
ness and their specificity. A striking example of their effectiveness is provided 
by the way in which the enzyme urease catalyzes the hydrolysis of urea, a 
product of protein metabolism. The nonenzymic reaction under neutral con- 

o 

II urease 
NH 2 -C-NH 2 + H 2 > 2NH 3 +C0 2 



ditions in water is so slow that the reaction is virtually undetectable at room 
temperature. The rate can be measured at high temperatures, however, and 
extrapolated to room temperature. This reveals that at low concentrations of 
urea the enzymic hydrolysis is about 10 1A times faster than the uncatalyzed 
reaction. 

There is an important limitation on the catalytic effectiveness of enzymes ; 
it is commonly observed that as the concentration of substrate is increased 
the rate of reaction tends to level off. The explanation for this is that the sub- 
strate and enzyme form a complex which then decomposes to products. De- 
spite the large sizes of enzyme molecules, there are usually only a few sites 
(often only one) at which reaction occurs ; these are called the active sites. The 
function of the rest of the molecule is normally to bring substrate(s) and active 
site together. With a large excess of substrate, the active sites are continually 
being filled and the rate or decomposition to products and their removal 
from the active sites is the rate-limiting factor. Increasing the number of sub- 
strate molecules awaiting reaction increases the rate of product formation only 
up to the point where the enzyme becomes saturated with substrate. This 
phenomenon is called the Michaelis-Menten effect. 

The reasons for the high specificity of enzymes have been the subjects of 
lively debate. According to one view — the " induced-fit " theory — the catalytic 
groups at the active site in an enzyme only take up positions in which they can 
interact with a substrate as a result of a conformational change that forces 
the enzyme into a slightly less energetically favorable, but catalytically active, 
spatial arrangement. This theory accounts for the observation that certain 
compounds, even though they may bind to the active site of an enzyme, do 
not undergo further reaction. They do not have the necessary structural 
features to induce the critical conformation at the active site. 



18 -3 hydrolytic enzymes 



A large number of hydrolytic enzymes that catalyze esters and amide hydrol- 
ysis are known. Undoubtedly the most intensively studied of these is chymo- 
trypsin, an enzyme of the digestive tract. It is a protein molecule with a 
molecular weight of 24,500, consisting of three peptide chains. There are 
two amino acid units in the enzyme that are known to be intimately involved 



sec 18.3 hydrolytic enzymes SOS 

in the bond-breaking steps of ester hydrolysis. These are histidine [6] and 
serine [7]. 

CH 2 CHC0 2 H HOCH,CHC0 2 H 



HN N NH 2 NH 2 

histidine serine 

[6] [7] 

When an ester such as phenyl acetate, CH 3 — C , is hydrolyzed 

OC 6 H 5 

by the action of chymotrypsin, the acetyl group, CH 3 — C , is actually 

transferred to the enzyme. In a subsequent step the acetylated enzyme is 
hydrolyzed to acetic acid and the enzyme is regenerated. The hydroxyl group 

O ° 

/ II 

E — H + CH 3 — C ► E— C — CH 3 + C 6 H 5 OH 

OC 6 H 5 

acetylated 
enzyme enzyme 

O O 

II / 

E — C— CH 3 + H 2 ► EH + CH 3 C 

OH 

in the serine unit in the chain has been identified as the group that becomes 
acetylated. The imidazole ring of the histidine unit is known to aid this trans- 
fer, possibly by acting as a base. In the mechanism shown in Figure 18-2, the 



Figure 18-2 Possible mechanism for the chymotrypsin-catalyzed hydrolysis 
of phenyl acetate; E = enzyme. 



* \ ri )T 

HN \^ N! ^ H ~°~" ' C-CH 3 -HN. ? ,HN O-C-CH, 

C 6 H s O oc H 

[8] 



O 

HN \J* 6— C + C 6 H s OH 

X CH 3 
[9] 

H,0 



+ CH 3 C0 2 H 



NH. N OH 



chap 18 enzymic processes and metabolism 506 

imidazole accepts a proton from the serine hydroxyl as the latter attacks the 
carbonyl carbon of the ester. 

The ester must be held in position at the active site of the enzyme by com- 
plex formation. This presumably involves hydrogen bonds to the carbonyl 
oxygen atom, thus avoiding a full negative charge being generated at this 
site in the intermediate [8]. The cleavage of the phenoxy group in [8] and the 
deacylation of [9] that restores the enzyme to its original state also seem to be 
assisted by the imidazole ring, acting as a base in each case. There are, in 
fact, two histidine units in the peptide chain of chymotrypsin, and it is possible 
that both are involved in one or more of these steps. 

Another important hydrolytic enzyme is acetylcholinesterase. Nerve cells 
contain the molecule acetylcholine [10] in a bound state. Stimulation of the 
cell releases acetylcholine, which stimulates the neighboring cell to release 
acetylcholine and this, in turn, its neighbor, thus transmitting the nerve im- 
pulse. Deactivation of the stimulant must be done very quickly once the im- 
pulse is transmitted. Deactivation is achieved by the enzyme acetylcholines- 
terase, which catalyzes the hydrolysis. 

o 

II ffi acetylcholinesterase ? T ,^r, s 

CH 3 -C-OCH 2 CH 2 N(CH 3 ) 3 + H 2 ► CH 3 C0 2 H + HOCH 2 CH 2 N(CH 3 ) 3 

acetylcholine choline 

[10] 

A nerve poison such as diisopropyl ftuorophosphate, [(CH 3 ) 2 CHO] 2 POF, 
forms a stable ester with the serine hydroxyl at the active site in the enzyme, 
thus preventing the deactivation step from occurring. 



18-4 oxidative enzymes 

The groups ordinarily present in a peptide chain do not undergo facile oxi- 
dation or reduction and, as a consequence, the enzymes involved in biological 
oxidation and reduction processes utilize a coenzyme which serves as the 
actual oxidant or reductant. One of the most important coenzymes in this 
regard is the molecule nicotinamide-adenine dinucleotide (Figure 18-3), ab- 
breviated NAD®. [This substance is often called " diphosphopyridine nucleo- 
tide" (DPN®). The name used here is that approved by the International 
Union of Biochemistry.] The structure of this molecule is not unlike that of 
adenosine triphosphate (ATP) met earlier (Section 15-5). It contains the 
adenosine ring (upper right in Figure 18-3) attached to a ribose molecule 
which in turn has a phosphate link. In NAD®, this is diphosphate, not tri- 
phosphate, and is further linked through another ribose ring to a nicotinamide 
unit (upper left in the formula). The pyridinium ring in the latter unit is the 
active oxidant. Since the molecule contains two nitrogen bases, two sugar 
units, and two phosphates it is a dinucleotide. 

Ethanol is oxidized in the liver by NAD® under the influence of the enzyme 
alcohol : NAD oxidoreductase. (This enzyme is often called alcohol dehydro- 
genase, but such a name implies that the enzyme acts in one direction only. 
Like all catalysts, enzymes increase the rates of both forward and reverse 



sec 18.4 oxidative enzymes 507 



H,N' 



Xa — kc 

L/OHHON 



CH,— O- 



o e 
I 

-p— o— 
II 
o o 



NH 2 



P-0-H 2 ^0^\ N J^ N J 




HO OH 



Figure 18-3 The coenzyme nicotinamide-adenine dinucleotide (NAD®). 
Although it is anionic at neutral pH, because of the ionization of the diphos- 
phate linkage, the reactive part of the coenzyme bears a positive charge, 
hence the abbreviation NAD®. 



reactions). In the oxidation of ethanol to acetaldehyde a hydride ion is trans- 
ferred from C-l to the pyridinium ring in the coenzyme at the same time as 
a proton is lost by the hydroxyl group. 



CH 3 CH 2 OH + 



H,N 



CH,-C + H e 



HH 




1SK 



There is extensive evidence in support of this mechanism which will not 
be covered here. Instead we will consider three important general questions. 
First, what is the function of the remainder of the coenzyme molecule? 
Second, what is the function of the enzyme itself? And third, how is the re- 
duced form of the coenzyme (abbreviated NADH) oxidized back to NAD® 
so that it can be used again ? 

The coenzyme contains a number of hydrogen-bonding groups that must 
function to bind the coenzyme to the enzyme in such a way that the pyri- 
dinium ring is in a favorable position to react. For its part, the enzyme func- 
tions to complex both the coenzyme and the substrate. But it must also be 
involved in removing the hydroxyl hydrogen as a proton because otherwise 
the energetically unfavorable species CH 3 CHOH® (protonated acetaldehyde) 
would be formed. 

Either a carboxylate ion or an amino group in the coiled protein chain of 
the enzyme might be hydrogen bonded to the hydroxyl group of the alcohol 
and be able to accept the proton completely at the critical stage of C— H bond 
rupture. We saw in the previous chapter that a number of amino acids contain 
such additional groups. Because amino groups are extensively protonated at 
physiological pH, we show a carboxylate ion as the proton acceptor in Figure 
18-4, which shows a plausible, but rather simplified, view of the enzyme- 
catalyzed oxidation of ethanol by NAD®. 

The NADH that is formed is reoxidized by another enzyme-coenzyme 
system. The ultimate oxidizing agent for most physiological oxidations is, of 
course, molecular oxygen, but its reaction with NADH is extremely slow. 



chap 18 enzymic processes and metabolism 508 





transition state 




Figure 18-4 Possible mechanism for the NAD© oxidation of ethanol under 
the influence of the enzyme alcohol: NAD oxidoreductase; E= enzyme. 



A whole battery of enzymes is required to catalyze the overall reaction : 



CH 3 CH 2 OH + 3 2 



-* 2C0 2 + 3H 2 



We have seen that NAD®, a hydride acceptor, reacts directly with ethanol. 
The enzyme that reacts with oxygen must have rather different characteristics 
because oxygen is a substance which invariably reacts by one-electron steps. 
(Transfer of hydride ion, H e , amounts to a two-electron or two-equivalent 
group oxidation-reduction step. Transfer of H® is neither oxidation nor re- 
duction, transfer of H- is a one-electron step, and transfer of H : e is a two- 
electron step.) 

The enzyme systems that react directly with oxygen in these reactions are 
called cytochromes. They contain an iron atom complexed in a cyclic system 
that resembles that in hemoglobin (Section 25-4). Whereas the ferrous ion in 
hemoglobin complexes with oxygen, the ferrous ion in the cytochromes is 
oxidized to the ferric state, each iron atom undergoing a one-electron change. 



4 Fejl, + 2 + 4 H s 



-» 4Fe m ,+2H 2 



Between Fe'" t and NADH come a number of other enzyme systems, one 
of which must be capable of reacting both with a one-equivalent couple such 
as Fe n -Fe m and with a two-equivalent couple such as NAD®-NADH. The 
enzyme systems that play this role are known as flavins. 

Batteries of enzymes able to bring about the overall oxidation of ethanol 



sec 18. S the energetics of metabolic processes 509 

to acetaldehyde (and similar processes) are located in the mitochondria. These 
are subcellular, oblong bodies found in the oxygen-consuming cells of plants 
and animals. They are reponsible for oxidizing carbohydrates and fats to 
carbon dioxide and water and supplying the energy for processes such as 
muscle action, nerve impulses, and chemical synthesis. The sequence of re- 
actions is often called the electron-transport chain, although this name ob- 
scures the fact that most of the reactions, including the oxidation steps, are 
not, in fact, simple electron transfer processes. 



18-5 the energetics of metabolic processes 

The mitochondrial enzymes not only speed up the rate of the overall reaction 
of oxygen with say, glucose, by an enormous factor (probably in excess of 
10 20 ), but they also control the energy release so that it can be used for a 
multitude of purposes and not simply appear as heat. This is done by har- 
nessing the many oxidation steps to the synthesis of adenosine triphosphate 
(ATP) from adenosine diphosphate (ADP) and inorganic phosphate (Equa- 
tion 18T0). Hydrolysis of ATP releases this energy in such a way that the 
system can utilize it as work. Neither the mechanisms of the coupling of the 
oxidation steps to ATP synthesis (oxidative phosphorylation) nor the coupling 
of the hydrolysis step to production of work are very well understood at 
present. 



e 

^-I'- 
ll 
o 



I 

O-P — OH,C 

II 

o 




NH 2 



+ HP0 4 + H £ 



OH OH 
ADP 




H,0 



(18-10) 



AG = +7 kcal (at pH 7.0, 25°) 



ATP and ADP are shown in Equation 18T0 in the ionic forms which 
predominate at pH 7. 

The complete oxidation of one mole of glucose releases 673 kcal of heat 
and the standard free energy change for the reaction is almost the same. It is 



chap 18 enzymic processes and metabolism 510 
Table 18*1 Heats of combustion of various substances 



compound 


state 


formula 


Ml, 


kcal 


per mole 


per gram 


ethane 


g 


CH3CH3 


370 


12.3 


octane 


1 


■ CH 3 (CH 2 ) 6 CH 3 


1303 


11.4 


cetane 


s 


CH 3 (CH 2 ) 14 CH 3 


2560 


11.3 


ethanol 


1 


C2H5OH 


328 


7.1 


glyceryl tri- 
stearate 
(a typical fat) 


s 


CH 2 2 CCi7H 35 

1 

CHO2CC17H35 

1 
CH 2 2 CCnH 35 


8290 


9.3 


glucose 


s 


CeHi 2 6 


673 


3.7 


starch 


s 


(CeHioOs)* 


x-678 


4.2 


glycine 
protein 


s 
s 


NH 2 CH 2 C0 2 H 

O 

11 
(-NHCHC-), 
1 
R 


235 
159" 


3.1 

2.1" 

<4" 



" To give carbon dioxide, liquid water, and nitrogen. (Some of these values differ from those 
given in Chapter 2, where all the products were assumed to be gases.) 
* To give carbon dioxide, liquid water, and urea. 



clear that such a process, even if it could occur rapidly in a cell, has to be 
partitioned into small packets of energy for ATP synthesis (AG = +7 kcal) 
to be accomplished. 



C 6 H 12 6 +6 2 



-*• 6 C0 2 + 6 H 2 



AG = -686 kcal 
A#= -673 kcal 



Oxidation (in effect combustion) of fats, carbohydrates, and proteins pro- 
vides the energy for metabolic processes of animals. A comparison of the 
heats of combustion of these classes of compounds reveals that fats release 
far more heat on a weight basis than do either carbohydrates or proteins. 
Table 18T gives the heats of combustion of compounds of each type and in- 
cludes for comparison those of some other fuels. The combustion energy 
clearly decreases as the degree of oxidation of the compound increases. Thus, 
hydrocarbons (AH ~ —11 kcal g _1 ) release the most energy, followed by 
fats (A//~ —9.3 kcal g _1 ) which contain long hydrocarbon-like chains. 
Carbohydrates (AH ~ —4.0 kcal g _1 ) are already in a partly oxidized state 
and so their available combustion energy is less. (This is reflected in the smaller 
amounts of oxygen they consume during combustion.) 

Why are proteins (AH 4 kcal g _1 or less), which contain less oxygen 

than carbohydrates, not better sources of energy? To a great extent because 



summary 5 1 1 

they do not undergo complete combustion. In man and other terrestrial verte- 
brates most of the nitrogen in proteins is converted to urea and excreted as 
such or as its hydrolysis product, ammonia. This is rather wasteful energetic- 
ally because the combustion of urea with formation of nitrogen would release 
an additional 2.5 kcal g -1 . 

O 

NH 2 -C-NH 2 + | 2 ► C0 2 +N 2 +2H 2 AH= -152 kcal 

ur ea (2.5 kcal g" 1 ) 

However, having urea as the end product of protein metabolism provides 
those organisms that are unable to fix N 2 with a source of nitrogen for 
chemical synthesis. 

The fixation of N 2 by certain bacteria is a biological reduction reaction of 
great interest. Very few nonbiochemical systems are known which will react 
with N 2 at room temperature and atmospheric pressure. Lithium metal com- 
bines slowly with nitrogen to form lithium nitride Li 3 N. Another is the 
titanium(II) compound titanocene, whose simplest formula is (C 5 H 5 ) 2 Ti, sug- 
gesting that it might be a sandwich compound like ferrocene (C 5 H 5 ) 2 Fe 
(Section 9-9F). Titanocene appears to be a dimer, however, and its true struc- 
ture has yet to be established. It reacts with nitrogen at room temperature to 
give a compound (C 5 H 5 ) 2 TiN 2 (again apparently dimeric) which can be 
reduced to ammonia. 



summary 

The fastest chemical reactions are those which occur at each encounter be- 
tween reactants. The controlling factors for such reactions are the rates of 
diffusion of the reactant molecules ; in solution at room temperature, these have 
rate constants of about 10 10 liters mole -1 sec -1 . The rates of other chemical 
reactions depend on the energy differences between reactants and transition 
states (the activation energy). A catalyst speeds up a reaction by providing an 
alternate path from reactants to products via a lower-energy transition state. 
Heterogeneous catalysts operate by providing a favorable surface for the re- 
action to occur on, while homogeneous catalysts form reactive combinations 
in solution which subsequently give the reaction products and regenerate the 
catalyst. 

Many acid-catalyzed reactions of carbonyl compounds take place by way 
of this general path: 

O e OH OH OH 

# +H ®. / HY I ® 

R-C . > R-C > R-C-YH 

z z 




chap 18 enzymic processes and metabolism 512 

The rate of such a reaction is proportional to [RCOZ] [HY][H e ], provided 
the concentration of acid catalyst is not so great as to substantially proto- 
nate the carbonyl compound. 

Many base-catalyzed reactions of carbonyl compounds occur by this 
general path: 



ZH + RO e 



=t Z + ROH 



R-C 



O e OH 

R _C-Y -5£»* R-C-Y + RO e 

I I 

z z 



Closely related to base catalysis is nucleophilic catalysis, in which the 
catalyst forms a bond to the substrate rather than removing a proton from it. 
Imidazole is particularly effective in this regard. 

General acid catalysis occurs when the reaction rate depends on the con- 
centration of all the acids present in the system, not simply on the hydrogen 
ion concentration. Most reactions of this type involve attack of a base on a 
protonated intermediate. The combined effect of the proton H e and the base 
A e appears in the kinetic expression as the function [HA]. 

Some reactions are subject to neighboring group (anchimeric) assistance. 
If the neighboring group is regenerated in a subsequent step, this can be 
thought of as intramolecular catalysis. 



■col 



-X s 



o 

O 



Y s 



-co? 



Tautomeric catalysts are those which can donate and accept protons si- 
multaneously, such as a-pyridone, which is especially effective for catalysis 
of mutarotation of sugars. 

Enzymes are highly efficient biological catalysts. They are protein molecules 
that sometimes have smaller units such as coenzymes associated with them. 
The active site is the place in the protein chain where the substrate becomes 
bound and where reaction actually occurs. 

A hydrolytic enzyme such as chymotrypsin does not require a coenzyme. 
Instead, histidine and serine units in the protein chain appear to function 
together to effect the hydrolytic cleavage of esters and amides. Histidine con- 
tains an imidazole ring which probably acts as a base in removing a proton 
from the serine hydroxyl group at the same time that the oxygen of the latter 
bonds to the carbonyl carbon of the ester. The acylated serine that is thus 
formed is subsequently cleaved to give the free acid. 




HN N 



exercises SI 3 

Oxidative enzymes and coenzymes include nicotinamide-adenine dinucleo- 
tide, NAD®; the flavins; and the cytochromes. NAD® is a coenzyme that 
behaves as a two-equivalent oxidant by removing hydride ion from com- 
pounds such as ethanol. The pyridinium ring is thereby reduced and the com- 
pound NADH is formed. The function of the enzyme is to orient the co- 
enzyme and substrate in such a way as to facilitate the reaction. 

o o 

II II H H 



N-" 




N' 



R R 

In those parts of living cells called mitochondria, the biological oxidation 
chain includes NAD® at one end (oxidants for various covalent organic 
compounds) and the cytochromes at the other (systems that are oxidized by 
molecular oxygen). These disparate kinds of redox systems are linked by a 
number of other enzyme-coenzyme systems. 

Much of the energy released by virtue of combustion of organic compounds 
in living cells is stored by synthesis of adenosine triphosphate (ATP) from the 
diphosphate (ADP) and inorganic phosphate, and subsequently used to drive 
a multitude of physiological processes. 

Hydrocarbons, fats, carbohydrates, and proteins release on combustion 
about 1 1 , 9, 4, and less than 4 kcal per gram, respectively. This order reflects 
to a great extent the state of oxidation of the compounds themselves. 



exercises 

18-1 Will the rate of a diffusion-controlled reaction be affected by a change in 
temperature of the system ? 

18-2 The reaction of 1-chlorobutane with sodium hydroxide to give w-butyl 
alcohol is catalyzed by sodium iodide. Work out the stereochemistry to be 
expected for both the catalyzed and the uncatalyzed reactions if right-handed 
1 -chlorobutane- 1 -D 

H 

I 
(CH 3 CH 2 CH 2 — C— CI) were used as the starting material. Show your 
I 
D 

reasoning. 

18-3 Is the hydrolysis of amides brought about by hydroxide ion an example of a 
base-catalyzed reaction ? 

18-4 Why do the laws of conservation of energy require that a catalyst increase 
both the forward and reverse rates of a chemical reaction ? Is there a viola- 
tion of this principle in the system used for the preparation of diacetone 
alcohol? (See Figure 12-1.) 



chap 18 enzymic processes and metabolism 514 

18-5 Draw resonance structures for the cation and for the anion of imidazole 
(formed by protonation and deprotonation, respectively). 

18-6 What might one conclude about the active site of a-chymotrypsin from the 
fact that negatively charged inhibitors are less effective in reducing catalytic 
activity than neutral molecules of the same type of structure ? 

18-7 Is important conjugation possible between the amide group and the ring 
nitrogen in the dihydropyridine ring of NADH? (See Figure 18-4). 

1 8-8 Which of the following reagents are likely to be one-equivalent oxidants and 
which two-equivalent oxidants ? 

O 



Co'", TV", Pd", (CH 3 ) 3 C®, ©OC1, Fe(CN), 



36 



CH 3 ° 



18-9 Assuming the carboxylate group shown in Figure 18-4 belongs to a non- 
terminal amino acid in the enzyme, which amino acids could fill this role ? 

18-10 Arrange the following compounds in the order of their expected heats of 
combustion (on a per gram basis): 1,4-butanediol, 1,3-butadiene, 1,2- 
butadiene, 2-methylpropanoic acid. 

18-11 Let us assume that memory depends on proteins with particular amino acid 
sequences being deposited in the brain. Calculate the approximate number 
of amino acid molecules that would be deposited per second if one gram 
(0.04 ounce) of protein is added to the brain in this way in one year. In your 
opinion is this number sufficiently high that the source of memory could be 
explained on this basis ? 



<m- $?& m. 



mmmMm%MMfM : m 



1 &*gjf?ii^ ! 

phosphorus, silicon, 

' and boron 



tegn ! 



Hi: 







^11 




^ 


■'SV 


IHIjBi 


.;■;*{■■;' 








£.5 


,--> \ 




', * 'i'-- 





af S iSJ 



L "'.-. v "'.'-'- '^' -V='i 






chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 517 

The four elements we shall consider in this chapter occupy positions in the 
periodic table that are adjacent to carbon, nitrogen, and oxygen, the three 
elements whose compounds have been our chief concern up to this point. 
Figure 19-1 shows the locations of these seven elements and the number of 
valence-shell electrons that each possesses. 

A similarity is to be expected between the bonding pattern exhibited by a 
second-row element and by the element immediately above it, because both 
possess the same number of valence electrons. However, a second-row element 
has an additional degree of freedom since it is not necessarily restricted to a 
maximum of eight electrons in its bonding shell. The compounds SF 6 and 
PF 5 are both stable gaseous compounds, and it is clear that sulfur and 
phosphorus in these molecules have a share in 12 and 10 bonding electrons, 
respectively. The electron configurations of atomic oxygen and sulfur are 
shown in Figure 19-2. It appears that sulfur must make use of its 3d orbitals in 
forming SF 6 , as does phosphorus in forming PF 5 . Although silicon can 
also use its d orbitals in bonding (SiF 6 2e is known), it generally tends to be 
tetracovalent like carbon. Boron is different from any of the elements we have 
been discussing and we shall consider its pattern of bonding in Section 19-5. 



19-1 d orbitals and chemical bonds 

The maximum number of s electrons is two, of p electrons six, and of d 
electrons 10 in any atomic shell. Although s orbitals are spherical and p 
orbitals dumbbell shaped, it was shown in Chapter 1 that the shape of the 
bonding orbitals in a molecule may bear little resemblance to these atomic 
precursors. Thus, CH 4 has a tetrahedral configuration for its bonds because 
this is the arrangement in which repulsion between the bonding electrons is at 
a minimum. 

The three p orbitals on an isolated atom run along the three geometric 
axes, x, y, and z, and are all clearly equivalent. The four bonding orbitals in a 
molecule such as CH 4 are tetrahedrally arranged and, again, all four are 
clearly equivalent. When we come to the five d orbitals that assume impor- 
tance in the bonding of second-row elements, we encounter a new situation. 
If five bonds around a central atom are separated to the maximum extent they 
cannot be equivalent. This may not seem obvious at first glance. One might 



Figure 19*1 The first two rows of the periodic table showing the relative 
positions of boron, silicon, phosphorus, and sulfur with respect to carbon, 
nitrogen, and oxygen, and the numbers of valence electrons of each. 



Li 


Be 


•B 


•C- 


:N- 


:6: 


F 


Ne 


Na 


Mg 


Al 


•Si- 


:P- 


:S: 


CI 


Ar 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron S18 



SB 

1- 

c 
o 

.5 


oxygen 

3 d o o o o o 

3PO O O 
3s O 


sulfur 

ooooo 

© © © M 

© 


2p © © © 

2s © 


@ © © 

© 


Is © 


@ K 



Figure 19-2 Electronic configurations of oxygen and sulfur. 

expect that five lines can be drawn radiating from a central point so that all 
are equivalent, just as four lines can. 



Solid geometry shows us that this is not so. There are only five regular 
polyhedra, the tetrahedron (four sides, four apexes), the cube (six sides, 
eight apexes), the octahedron (eight sides, -six apexes), the dodecahedron 
(12 sides, 20 apexes), and the eicosahedron (20 sides, 12 apexes). This means 
that only those three-dimensional spatial arrangements with four, six, eight, 
12, and 20 apexes have their apexes in identical and equivalent positions. 
Since the five atomic d orbitals are, in fact, exactly equivalent in energy, what 
shapes do they assume? If they are all completely filled or half-filled, the 
resulting electron cloud is spherically symmetrical. But this does not help to 
provide us with a picture of the five component orbitals. A mathematical 
solution to the electronic wave equation, however, provides the shapes of the 
five d orbitals as shown in Figure 19-3. Despite the different appearance of the 
d z 2 orbital, it is equivalent in energy to the other four rosette-shaped d orbitals. 

O 



Should we assume that the bonds in molecules such as SF fi or HO— S- 

II 
O 



OH 



point in directions corresponding to the directions of maximum extent of the 
d orbitals? The same problem arises here as in inferring the tetrahedral ar- 
rangement of bonds in methane from the shapes of s or p orbitals. The maxi- 
mum number of electrons that are found in the valence shell of sulfur is 12 
and, if the bonds are all equivalent (as they are in SF 6 ), then the six pairs of 
electrons will on the average be located at the corners of a regular octahedron. 



sec 19.1 d orbitals and chemical bonds 519 




The six bonds are sometimes designated d 2 sp 3 hybrids, or octahedral 
bonds, just as the four single bonds to carbon are designated sp 3 or tetra- 
hedral bonds. 

Double bonds to tetravalent or hexavalent sulfur, as in (CH 3 ) 2 S=0 or 

O 

II 
CH,— S— CH, , differ from double bonds to carbon in that the n bonds are 
II 
O 

formed not by p orbitals on sulfur but by something close to d orbitals. 
Divalent sulfur compounds, on the other hand, such as thioketones, R 2 C=S, 
are formed by interaction of p orbitals on both carbon and sulfur, Such com- 
pounds are relatively uncommon and often are unstable with respect to 
polymerization, which can be ascribed to low effectiveness of 71-type interac- 
action involving 3p orbitals. In this connection we may note that S 2 , unlike 
2 , is highly unstable and that elemental sulfur is most stable in the cyclic S g 
form. The reluctance of sulfur to form double bonds to carbon is also exhib- 
ited by phosphorus and silicon, and no stable compounds are known with 
C=Si or C=P bonds. 

Dimethyl sulfoxide, (CH 3 ) 2 S=0, has an unshared pair of electrons on 
sulfur and, unlike acetone, (CH 3 ) 2 C=0, is nonplanar. 



Figure 19-3 The five atomic d orbitals; all are equivalent in energy. 




A,2__ v 2 





ld 2 2 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 520 



/h 3 c^ „ 

/ x 

/ H 3 C 


V 




o 

\ \® e || 

Some persons prefer to write S=0 bonds as S— O and — S— bonds as 
e / / II 

I ° 

— S— , thus preserving an octet arrangement about sulfur. The difference in 

e 

electronegativity between oxygen and sulfur suggests that such structures 
should make important contributions to the resonance hybrid. Nonetheless, 
we shall use the double-bonded formulas in this book because they emphasize 
the d-orbital participation in the bonding. It is important to recognize that 
drawing S— O does not imply any necessary correspondence to carbon- 
oxygen or carbon-nitrogen double bonds. 



19-2 types and nomenclature of organic 
compounds of sulfur 

A number of typical sulfur compounds with their common and IUPAC 
names are listed in Table 19-1. It can be seen that the divalent sulfur deriva- 
tives are structurally analogous to oxygen compounds of types discussed in 
earlier chapters. These sulfur derivatives are often named by using the prefix 
thio in conjunction with the name of the corresponding oxygen analog. The 
prefix thia is occasionally used when sulfur replaces carbon in an organic 
compound. 



CH 3 CH 2 SH 

ethanethiol 



CH 3 -C x 

S-H 

dithioaeetic acid 
(ethanethionthioic acid) 



Q-SH 

thiophenol 



C„H, 



-CH, 

I " 
-CH, 



thiacyclobutane 

(trimethylene sulfide) 

(thietane) 



C„H 



C„H< 



6"5 "j* V '6' J 5 

C 6 H 5 

1 ,2,4,6-tetraphenylthiabenzene 



1 Terms such as organosulfur, organophosphorus, and so on actually refer to those com- 
pounds containing the bonds C— S, C— P, and so on. Just as sodium acetate is not regarded 
as an organometallic compound, so dimethyl sulfate, (CH 3 0) 2 S0 2 , is not regarded as an 
organosulfur compound; the links between carbon and the other element are through 
oxygen in both cases. 



sec 19.2 types and nomenclature of organic compounds of sulfur 521 
Table 19-1 Typical organic compounds of sulfur 



compound 


oxygen analog 


common name 


IUPAC name 


CH 3 SH 


ROH 




methyl mercaptan 


methanethiol 


C2H5SC2H5 


ROR 




diethyl sulfide 
diethyl thioether 


ethylthioethane 


C 6 HsSSC6H 5 


R-O 


-0-R 


diphenyl disulfide 


phenyldithio- 
benzene 


© e 
(CH 3 ) 3 S Br 


R3O 


e 
X 


trimethylsulfonium 
bromide 




(C 6 H 5 )2C=S 


R 2 C= 


O 


thiobenzophenone 




N0 2 










2N -^-s- 


-CI ROCi 


2,4-dinitrobenzene- 
sulfenyl chloride 


2,4-dinitrobenzene- 
sulfenyl chloride 


O 

II 
C2H5-S-OH 






ethanesulfinic acid 


ethanesulfinic acid 


O 

II 

CH 3 -S— OH 

II 

O 






methanesulfonic 
acid 


methanesulfonic ■ 
acid 


O 

II 
CH 3 -S— CI 
II 
O 






methanesulfonyl 
chloride 


methanesulfonyl 
chloride 


O 

II 
CH3 S — C2H5 



II 

CeHs S CeH 5 

11 







ethyl methyl 
sulfoxide 


methylsulfinyl- 
ethane 






diphenyl sulfone 


phenylsulfonyl- 
benzene 




11 
CH3O-S— 0CH3 

11 







dimethyl sulfate 


dimethyl sulfate 











A. THIOLS 



The thiols (or mercaptans) are derivatives of hydrogen sulfide in the same way 
that alcohols are derivatives of water. The volatile thiols, both aliphatic and 
aromatic, are like hydrogen sulfide in possessing characteristically disagree- 
able odors. 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 522 

A variety of thiols occurs along with other sulfur compounds to the extent 
of several percent in crude petroleum. Besides having objectionable odors, 
these substances cause difficulties in petroleum refining, particularly by 
poisoning metal catalysts. The odors from pulp mills arise from volatile 
sulfur compounds formed during the digestion operation (Section 15-7). 

Thiols also have animal and vegetable origins; notably, butanethiol is a 
component of skunk secretion; propanethiol is evolved from freshly chopped 
onions, and, as we have seen in Chapters 17 and 18, the thiol groups of 
cysteine are important to the chemistry of proteins and enzymes. 

In many respects, the chemistry of thiols is like that of alcohols. Thus 
thiols can be readily prepared by the reaction of sodium hydrosulfide (NaSH) 
with those alkyl halides, sulfates, or sulfonates which undergo S N 2 displace- 
ments. This synthesis parallels the preparation of alcohols from similar 
substances and hydroxide ion (Chapter 8). 

C 2 H 5 Br + SH e ► C 2 H 5 SH + Br e 

Since thiols are acids with strength comparable to hydrogen sulfide 
C^ha — 6 x 10~ 8 ), some thioethers may be produced by the following 
sequence of reactions, unless the sodium hydrosulfide is used in large 
excess : 

C 2 H 5 SH + SH e . C 2 H 5 S e +H 2 S 

C 2 H 5 S e + C 2 H 5 Br ► (C 2 H 5 ) 2 S + Br e 

Thiols can also be prepared by the reaction of Grignard reagents with sulfur 
(Section 9-9C). 

Br MgBr SMgBr SH 

Mg f) S f] H ffi ,H 2 



ether 



cyclohexyl cyclohexane- 

bromide thiol 

Thiols do not form as strong hydrogen bonds as do alcohols, and conse- 
quently the low-molecular-weight thiols have lower boiling points than 
alcohols; thus ethanethiol boils at 35° compared to 78.5° for ethanol. The 
difference in boiling points diminishes with increasing chain length. 

The lack of extensive hydrogen bonding is also evident from the infrared 
spectra of thiols, wherein a weak band characteristic of S— H linkages 
appears in the region 2600 to 2550 cm -1 . In contrast to the O— H absorption 
of alcohols, the frequency of this band does not shift significantly with 
concentration, physical state (gas, solid, or liquid), or the nature of the solvent. 

It is perhaps surprising, in view of the smaller electronegativity of sulfur 
than oxygen, that thiols are considerably stronger acids than the correspond- 
ing alcohols. Thus K UA of ethanethiol is about 10" u , compared to 10~ 17 for 
ethanol. However, this behavior is not unusual, in that H 2 is a weaker acid 
than H 2 S, HF is weaker than HC1, and NH 3 is a weaker acid than PH 3 . 

Thiols form insoluble salts with heavy metals, particularly mercury. This 
behavior is the origin of the common name (now out of favor), for thiols, 
mercaptan. As mentioned above, alkali metal salts of thiols react readily in 



sec 19.2 types and nomenclature of organic compounds of sulfur 523 

S N 2-type displacements to yield thioethers, and this provides a general 
method of synthesis of these substances. The pronounced nucleophilicity of 
sulfur combined with its relatively low basicity makes for rapid reaction with 
little competition from elimination, except for those compounds where 
S N 2-type displacements are quite unfavorable and E2-type elimination is 
favorable. 

Thiols react with carboxylic acids and acid chlorides to yield thioesters and 
with aldehydes and ketones to yield dithioacetals and dithioketals, respec- 
tively. 

o o 

II II 

CH,CH,SH + CH,-C-C1 



ethyl thioacetate 

O H 3 C S— CH 2 

« HO \ / \ 

HSCH 2 CH 2 CH 2 SH + CH3-C-CH3 — ^-* C CH 2 

1 ,3-propanedithiol H 3 C S— CH 

acetone trimethylene 
dithioketal 

An important difference between thiols and alcohols is their behavior 
toward oxidizing agents. In general, oxidation of alcohols occurs with increase 
of the oxidation level of carbon rather than that of oxygen ; carbonyl groups, 
not peroxides, are formed. It takes a powerful oxidizing agent (e.g., Co 111 ) to 
achieve one-electron oxidation of oxygen by removing a hydrogen atom from 
the hydroxyl group of an alcohol. 

R— O— H + -X ► RO • + HX (a generally unfavorable reaction) 

In addition, the hydroxyl oxygen of an alcohol does not accept an oxygen 
atom from reagents like hydrogen peroxide, although these same reagents 
readily donate an oxygen atom to nitrogen of amines to form amine oxides 
(see Section 16-1F5). Why does the oxidation of thiols take a different course? 

:0: 9 
ROH + H 2 2 / > R:5:H ► R-O-O-H 



RN(CH 3 ) 2 + H 2 2 ► RN(CH 3 ) 2 + H 2 

First, because the strength of S— H bonds (83 kcal) is considerably less than 
that of O— H bonds (111 kcal); there is therefore good reason to expect that 
reaction mechanisms that are unfavorable with alcohols might well occur 
with sulfur. Thus we find that oxidation of thiols with a variety of mild 
oxidizing agents, such as atmospheric oxygen, halogens, sulfuric acid, and so 
on, produces disulfides, probably by way of thiyl radicals. The coupling of 
the amino acid cysteine to give cystine is an important example of this 
reaction (Section 17-1). 

R-S-H + [O] »■ R-S- + H [O] 

2RS- * RS-SR 

disulfide 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 524 

The second reason for the difference between alcohol and thiol oxidations 
is that compounds in which sulfur is in a higher oxidation state are frequently 
stable. Thus, vigorous oxidation of thiols with nitric acid, permanganate, or 
hydrogen peroxide gives sulfonic acids, possibly by way of the disulfides, or 
else through intermediate formation of the sulfenic and sulfinic acids, which 
are themselves too readily oxidized to be isolated under these conditions. 



R — S— S— R 



disulfide 



O 

II 
-♦ R-S — S-R - 

O 

thiosulfonate ester 



O O 

II II 
->• R— S — S— R 



o o 

disulfone 



R — SH 



sulfonic acid 



R-S-OH 

sulfenic acid 



O 

II 
R— S — OH 

sulfinic acid 



B. ALKYL SULFIDES 



Organic sulfides or thioethers, R— S— R', are readily obtained by displace- 
ment reactions between alkyl compounds and salts of thiols (Section 19-2A). 



HOCH 2 CH 2 SH + (CH 3 ) 2 S0 4 
ethan-l-ol-2-thiol 



aq. 25% NaOH 

1 

60°-70° 



HOCH 2 CH 2 SCH 3 

2-(methylthio)ethanol 



Sulfides undergo two important reactions involving the electron pairs on 
sulfur. They are rather easily oxidized to sulfoxides and sulfones (see next 
section), and they act as nucleophilic agents toward substances that undergo 
nucleophilic displacement readily to give sulfonium salts. The formation of 
sulfonium salts from alkyl halides is reversible, and heating of the salt 



CH 3 

:sfTcH 3 
/ 
CH 3 



n 



CH 3 

\ e 

®S-CH 3 I 
/ 
CH 3 



trimethylsulfonium iodide 



causes dissociation into its components. Sulfonium salts are analogous in 
structure and properties to quaternary ammonium salts ; sulfonium hydrox- 
ides, R 3 S e OH e , like quaternary ammonium hydroxides, R 4 N e 0H e 
(Section 16-IE1), are strong bases. 

A noteworthy feature of sulfonium ions is that, when substituted with three 
different groups, they can usually be separated into optical enantiomers. Thus 
the reaction of methyl ethyl sulfide with bromoacetic acid gives a sulfonium 
ion that is separable into dextro- and levorotatory forms by crystalliza- 
tion as the salt of an optically active amine. The asymmetry of these ions 



sec 19.2 types and nomenclature of organic compounds of sulfur S25 

stems from the nonplanar configuration of the bonds formed by sulfonium 

's'f .s® 

H 3 C V"CH 2 CH 3 CH 3 CH 2 y CH 3 

CH 2 C0 2 H H0 2 CH 2 C 

enantiomers of methylethyl(carboxymethyl)— sulfonium ion 

sulfur. The optically active forms of unsymmetrically substituted sulfonium 
ions are quite stable — surprisingly so, in view of the very low configurational 
stability of analogously constituted amines. Apparently, nonplanar com- 
pounds of the type R 3 Y:, where Y is an element in the second row of the 
periodic table, undergo inversion much less readily than similar compounds 
for which Y is a first-row element. Thus phosphorus compounds resemble 
sulfur compounds in this respect, and several asymmetric phosphines 
(R 1 R 2 R 3 P:) have been successfully resolved into enantiomeric forms. 



C. SULFOXIDES AND SULFONES 

Oxidation of sulfides, preferably with hydrogen peroxide in acetic acid, yields 
sulfoxides and sulfones. The degree of oxidation is determined by the ratio of 
the reagents, and either the sulfoxide or the sulfone can be obtained in good 
yield. 



O NH 2 

30% H 2 2 , 1.5 moles „„ | „„„.!, 



NH 2 

CI 

methionine 



CH 3 C0 2 H 



30% H Z Q 2 , 3.2 moles 
CH 3 CQ 2 H 



-+. CH,S-CH 2 CH 2 CHC0 2 H 



methionine sulfoxide 

O NH 2 

II I 2 

- CH 3 S-CH 2 CH 2 CHC0 2 H 

II 2 2 2 

O 

methionine sulfone 



Dimethyl sulfoxide (DMSO) is a particularly useful substance in the 
laboratory, both as a solvent and as a reagent. It is a polar substance with a 

o 

II 

CH 3 -S-CH 3 

dimethyl sulfoxide 
mp 18°, bp 189° 

fairly high dielectric constant (s = 48) and hence it dissolves polar and ionic 
substances quite well. It lacks hydrogen-bonding protons, however, and the 
activity of anions, particularly those whose charge is not dispersed, is high in 
DMSO solution. We have seen how this property greatly enhances S N 2 
reactivity (Section 8-1 ID). It is also responsible for the powerfully basic 
character of solutions of hydroxide or alkoxide ions in DMSO. Judging by 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 526 

their ability to remove protons from feeble acids such as aromatic amines 
(Table 22-1, footnote) and hydrocarbons these solutions are about 10 14 times 
more basic than the corresponding aqueous solutions. DMSO itself is very 
weakly acidic (pK HA = 33). Its anion, CH 3 SOCHf , called the dimsyl ion, has 
found wide use as a powerful base in elimination and other reactions. 

One of DMSO's most unusual solvent properties is its extraordinary ability 
to penetrate through cell membranes. When this property was discovered 
it was thought that DMSO might provide a vehicle for introducing medicinals 
directly into cells through the skin. However, large doses of DMSO have 
been found to cause retinal damage and this plan has been discarded. 

DMSO can also function as a mild but effective oxidant for alcohols. 
Treatment of an alcohol with DMSO and dicyclohexylcarbodiimide (a mild 
nonacidic dehydrating agent) produces high yields of aldehydes or ketones. 

o 

II 

RCH 2 OH + CH 3 SCH 3 + C 6 H! iN=C=NC 6 H! i 

O 
II 
* RCHO + CH3SCH3 + C 6 H! iNHCNHC 6 H! t 



This reaction is particularly useful if the alcohol is sensitive to acid rearrange- 
ment (Section 10-5B) or if, as in the case of the secondary hydroxyl groups in 
many carbohydrates, the alcohol resists oxidation by conventional means. 

D. SULFENIC, SULFINIC, AND SULFONIC ACIDS 

The sulfenic acids, RSOH, are unstable with respect to self-oxidation and 
reduction and generally cannot be isolated. However, certain derivatives of 
sulfenic acids are relatively stable, notably the acid halides RSC1. 

Sulfinic acids, RS0 2 H, are more stable than sulfenic acids but are none- 
theless easily oxidized to sulfonic acids, RS0 3 H. They are moderately strong 
acids with K HA values comparable to the first ionization of sulfurous acid 
(K HA ~10- 2 ). 

Many sulfonic acids, RS0 3 H, have considerable commercial importance as 
detergents in the form of their sodium salts, RS0 3 Na. Many commercial 
detergents are sodium alkylarylsulfonates of types which are readily syn- 
thesized from petroleum by reactions discussed in the next chapter. 



C9.! S H I9 . 3 ,-f VsOjNa 



The resistance of highly branched alkyl chains of arylalkylsulfonates to 
biochemical degradation and the water pollution that results led to their 
elimination as detergents for domestic purposes. Sodium arylalkylsulfonates 
with nonbranched side chains or sodium alkylsulfonates derived from long- 
chain alcohols are more easily degraded by bacteria. The principal advantage 
that sodium sulfonates have as detergents over the sodium salts of fatty acids 
(Section 13T) used in ordinary soaps is that the corresponding calcium and 
magnesium salts are much more soluble, and hence the sulfonates do not pro- 
duce scum (bathtub ring) when used in hard water. 



sec 19.2 types and nomenclature of organic compounds of sulfur 527 

Sulfonic acid groups are often introduced into organic molecules to increase 
water solubility. This is particularly important in the dye industry, where it is 
desired to solubilize colored organic molecules for use in aqueous dye baths 
(see Section 28-7A). 

Aliphatic sulfonic acids can be prepared by the oxidation of thiols (Sec- 
tion 19-2A). 

CH 3 CH 3 

I CH3CO2H I 

CiCH 2 CH 2 C-SH + 3 H 2 2 > C1CH 2 CH 2 C-S0 3 H + 3 H 2 

I I 

CH 3 CH3 

4-chloro-2-methyI- 4-chloro-2-methyI- 

2-butanethiol butane-2-sulfonic acid 

92% 

a-Hydroxysulfonate salts result from the addition of sodium bisulfite to 
aldehydes (Section 11-41). The free sulfonic acid RCHOHS0 3 H is unstable 
since addition of acid to the salt drives the reaction back to aldehyde by 
converting HS0 3 e to S0 2 . 

p OH 

R— C + NaHS0 3 y > R-CH-SOf Na® 

H 

Arylsulfonic acids are almost always prepared by sulfonation of the cor- 
responding hydrocarbon (see next chapter). They are strong acids, comparable 
in strength to sulfuric acid. Furthermore, the sulfonate group is an excellent 
leaving group from carbon in nucleophilic displacement reactions and, in 
consequence, conversion of an alcohol to a sulfonate ester is a means of 
activating alcohols for replacement of the hydroxyl group by a variety of 
nucleophilic reagents. A sulfonate ester is best prepared from a sulfonyl 
chloride and an alcohol, and many sulfonyl chlorides that can be used for 
this purpose are available commercially. The use of ^-toluenesulfonyl chloride 
(often called tosyl chloride) is illustrated in Equation 19-1. 

CH 3 -<f \-S0 2 Ci - 



(CH 3 CH 2 ) 2 CHOH *=* > (CH 3 CH 2 ) 2 CHO-S-/ V 



O 

C 2 H 5 S e 



jr~\ (i9-i) 

(CH 3 CH 2 ) 2 CH-S-C 2 H 5 + CH 3 -f V-SOf 



A number of important antibiotic drugs, the so-called sulfa drugs, are 
sulfonamide derivatives. 



NH 

// 



H 2 N-^f Vs0 2 NH-f \ H 2 N-f \-S0 2 NH-c' 

sulfadiazine sulfaguanidine 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 528 
E. SULFATE ESTERS 

Sulfate esters such as dimethyl sulfate lack a sulfur-carbon bond and as 
mentioned earlier are not classified as organosulfur compounds. Dimethyl 

o 
II 
CH3-0-S-0-CH3 
II 
o 

dimethyl sulfate, bp 188° 

sulfate is a good methylating agent; the methyl carbon easily undergoes 
attack by the substance being methylated because the group being displaced is 
a good leaving group. The leaving group in this reaction is the anion of a 
strong acid CH 3 — O— S0 3 H, methylsulfuric acid or methyl hydrogen sul- 
fate. (It should not be confused with the compound methanesulfonic acid, 
CH 3 — SO3H, also a strong acid.) 

O 

Nu + CH3-O-S-O-CH3 ► Nu— CH 3 + CHj-O-SOf 

O 

Dimethyl sulfate on hydrolysis produces sulfuric acid. 
(CH 3 0) 2 S0 2 + 2 H 2 >■ H 2 S0 4 + 2CH 3 OH 

Like most volatile esters of inorganic acids, dimethyl sulfate is toxic and 
should be handled with care. Contact of the vapors with the eye can cause 
permanent corneal damage. 



19-3 phosphorus compounds 



The two important groups of organic compounds of phosphorus are the 
phosphate esters, which contain oxygen-phosphorus bonds, and the organo- 
phosphorus compounds, which contain carbon-phosphorus bonds. 



A. PHOSPHATE ESTERS 

Phosphoric acid, H 3 P0 4 , has a tendency (absent in nitric acid) to exist in 
polymeric forms, such as diphosphoric acid, H 4 P 2 7 , and triphosphoric 
acid, H 5 P 3 O 10 . 

OO OOO 
II II II II II II 

HO-P-OH HO-P-O-P-OH HO-P-O-P-O-P-OH 

1 II III 
OH OH OH OH OH OH 

phosphoric acid diphosphoric acid triphosphoric acid 

(mono) 

Polyphosphate salts such as the sodium salts of triphosphoric acid are 
used in large amounts (up to 40 %) in detergents to bring clay and similar 
particles into suspension. Because of their high phosphorus content, deter- 
gents are believed to be a major contributor to eutrophication of lakes — 



sec 19.3 phosphorus compounds 529 

overfertilization caused by nutrient abundance. Runoff from fertilized 
agricultural land is also an important factor. Phosphorus being a nutrient 
for plant life stimulates the growth of algae to such an extent that other forms 
of marine life may be extinguished. Unlike nitrogen, which also contributes 
to eutrophication, phosphorus does not enter into biochemical reactions that 
allow it to escape from water as a gas. 

We have already met extremely important derivatives of each of the three 
acids, monophosphoric acid, diphosphoric acid, and triphosphoric acid. 
Nucleic acids are derivatives of monophosphoric acid (Figure 17-9) while 
adenosine diphosphate, ADP, and adenosine triphosphate, ATP (Section 
15-5), are derivatives of diphosphoric acid and triphosphoric acid, respec- 
tively. 



NH, 



-»© r>e 



H 2 C /°\\-S^ 




o° cr o" 

I I I 

R— O— P— O— P— O— P— OH 

II II II 

o o o 

OH OH 

adenosine triphosphate (ATP) adenosine group 

At pH 7, only one of the four phosphate hydroxyl groups in ATP remains 
un-ionized in the cell. The resulting triple negative charge on the triphosphate 
group is considerably reduced, however, by complex formation with mag- 
nesium ions. 

The energies of the ADP-ATP system have been extensively studied be- 
cause it is the link between high-energy phosphate donors, formed during 
the oxidation of foods (Section 18-5), and low-energy phosphate acceptors. 
The latter are activated by phosphorylation and can then perform cellular 
work: muscle contraction, biological transport, biosynthesis, and so on. 

The hydrolysis of ATP has a negative free energy ; that is, hydrolysis is 
favored at equilibrium. The free energy is more negative still in living cells, 
because of magnesium complexing in the cell. 

ATP + H 2 > ADP + HP0 4 2e AG standard = -7 kcal 

AGcellular ~ — 12 kcal 

Thus ATP is a high-energy phosphate compound. The frequently used phrase 
"high-energy bond" in connection with the ATP-phosphate link is a mis- 
nomer. We saw in Chapter 2 that the higher the energy of a bond, the more 
stable it is. The sense of this is opposite to the hydrolysis of ATP, which is 
energetically favorable. 

Phosphorous acid, H 3 P0 3 , has the structure [1], not [2], although organic 

OH 
II I 

H-P-OH :P-OH 

1 I 
OH OH 

[1] [2] 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 530 

derivatives of both of these structures are known. Ironically, it is the esters of 
[2] rather than the esters of [1] that are known as phosphites. The latter are 
called phosphonates. 

O OC 2 H 5 

II I 

C 6 H 5 -P-OC 2 H s :P-OC 2 H 5 

0-C 2 H 5 OC 2 H 5 

diethyl phenylphosphonate triethyl phosphite 



B. ORGANOPHOSPHORUS COMPOUNDS 

These compounds contain carbon-phosphorus bonds and resemble to some 
extent their nitrogen analogs. Thus trimethylphosphine is a weak base that 
can form phosphonium compounds analogous to ammonium compounds. 

(CH 3 ) 3 P:+HC1 >■ (CH 3 ) 3 PHC1 

trimethylphosphonium chloride 

(CH 3 ) 3 P: +CH 3 C1 ► (CH 3 ) 4 P ffi Cl e 

tetramethylphosphonium chloride 

Despite being weaker bases than the corresponding amines the phosphines 
are actually more nucleophilic, probably because phosphorus is a larger, more 
electropositive atom than nitrogen, and its outer-shell electrons are conse- 
quently less firmly held and more polarizable. 



C. REACTIONS OF QUATERNARY PHOSPHONIUM COMPOUNDS 

There are interesting differences in behavior of quaternary ammonium and 
quaternary phosphonium salts toward basic reagents. Whereas tetraalkyl- 
ammonium salts with hydroxide or alkoxide ions generally form alkenes and 
trialkylamines by E2-type elimination (Section 8-12), corresponding reactions 
of tetraalkyl- or arylphosphonium salts lead to phosphine oxides and hydro- 
carbons. 

(CH 3 ) 3 NCH 2 CH 3 + e OH >■ (CH 3 ) 3 N + CH 2 =CH 2 +H 2 

O 

© II 

(C 6 H 5 ) 3 PCH 2 CH 3 + e OH > (C 6 H 5 ) 2 PCH 2 CH 3 + C 6 H 6 

There is an alternative course of reaction of quaternary phosphonium salts 
with basic reagents. It involves attack of a base, usually phenyllithium, at a 
hydrogen a to phosphorus. The product [3] is called an alkylidenephos- 
phorane. 

a e e 

R 3 P-CH 2 R' + C 6 H 5 Li ► R 3 P=CHR' + C 6 H 6 + Li X 

X e [3] 

Compounds of this type are frequently written as dipolar structures such as 



sec 19.4 organosilicon compounds 531 

[4a]. However, they are better considered as hybrids of the contributing struc- 
tures [4a] and [4b], the latter involving p n -d K bonding. 

© e 
R 3 P-CHR < — > R 3 P=CHR 
[4a] [4b] 

Compounds in which two adjacent atoms bear opposite formal charges 2 
are called ylids (pronounced " illids. "). Although [4a] is only a contributor 
to the hybrid, alkylidenephosphoranes are often regarded as ylids. 

Alkylidenephosphoranes are reactive, often highly colored substances, 
that rapidly react with oxygen, water, acids, alcohols, and carbonyl com- 
pounds — in fact, with most oxygen-containing compounds. Two of these 
reactions are illustrated here for methylenetriphenylphosphorane and again 
the driving force is formation of a phosphorus-oxygen bond at the expense of 
a phosphorus-carbon bond. 



(C 6 H 5 ) 3 PCH 3 Bre -%Mi>. (C 6 H 5 ) 3 P=CH 2 



methyltriphenylphosphonium 
bromide 



methylenetriphenyl- 
phosphorane 



(C 6 H 5 ) 3 P=CH 2 + H 2 



♦ (C 6 H 5 ) 2 P=0 + C 6 H 6 
I 
CH 3 

methyldiphenylphosphine 
oxide 



(C 6 H 5 ) 3 P=CH 2 + (~/=0 



(C 6 H 5 ) 3 P=0 + <^)=CH 2 

triphenylphosphine methylene- 
oxide cyclohexane 



The preparation of alkenes from the reactions of alkylidenetriphenylphos- 
phoranes with aldehydes and ketones is known as the Wittig reaction. The 
example given here probably proceeds by way of the following intermediate. 



19-4 organosilicon compounds 



Silicon, like carbon, normally has a valence of four and forms reasonably 
stable bonds to other silicon atoms, carbon, hydrogen, oxygen, and nitrogen. 
Compounds of some of these types are listed in Table 19-2, together with the 

2 Betaine is a term given to compounds in which two nonadjacent atoms bear opposite 
charges. An amino acid zwitterion (Section 17-1B) is an example of a betaine. 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron S32 
Table 19-2 Principal types of silicon compounds and their carbon analogs 



silicon compound 


carbon compoun 


d 


silanes and organosilanes 




alkanes 




H 3 Si-SiH 3 


disilane 


CH 3 -CH 3 


ethane 


CH 3 -SiH 3 


methylsilane 


CH 3 -CH 3 


ethane 


(CH 3 ) 4 Si 


tetramethylsilane 


(CH 3 )*C 


neopentane 


organosilyl halides (halosilanes) 


alkyl halides (haloalkanes) 


(CH 3 ) 3 SiCl 


trimethylsilyl chloride (CH 3 ) 3 CC1 


r-butyl chloride 


H 2 SiCl 2 


dichlorosilane 


CH 2 C1 2 


dichloromethane 


silanols 




alcohols 




H 3 SiOH 


silanol 


H 3 COH 


methanol (carbinol) 


(CH 3 ) 3 SiOH 


trimethylsilanol 


(CH 3 ) 3 COH 


trimethylcarbinol (t- 
butyl alcohol) 


(CH 3 ) 2 Si(OH) 2 


dimethylsilanediol 


(CH 3 ) 2 C(OH) 2 


acetone hydrate 
(unstable) 


CH 3 Si(OH) 3 


methylsilanetriol 


CH 3 C(OH) 3 


orthoacetic acid 
(unstable) 


siloxanes and alkoxysilanes 


ethers 




(CH 3 ) 3 SiOSi(CH 3 ) 3 


hexamethyldisiloxane (CH 3 ) 3 COC(CH 3 ) 3 


di-r-butyl ether 


(CH 3 ) 3 SiOCH 3 


trimethylmethoxy- 
silane 


(CH 3 ) 3 COCH 3 


methyl Nbutyl ether 


(CH 3 ) 2 Si(OCH 3 ) 2 


dimethyldimethoxy- 
silane 


(CH 3 ) 2 C(OCH 3 ) 2 


acetone dimethyl ketal 


CH 3 Si(OCH 3 ) 2 


methyltrimethoxy- 
silane 


CH 3 C(OCH 3 ) 3 


methyl orthoacetate 



corresponding carbon compounds. Some idea of the strength of bonds to 
silicon relative to analogous bonds to carbon may be obtained from the 
average bond energies shown in Table 19-3. Significantly, the Si — Si bond is 
weaker than the C— C bond by some 30 kcal/mole, whereas the Si— O bond 



Table 19-3 Average bond energies 



bond 


bond energy, 
kcal/mole 


bond 


bond energy, 
kcal/mole 


Si— Si 


53 


C— C 


83 


Si— C 


76 


C— Si 


76 


Si— H 


76 


C— H 


99 


Si— O 


108 


C— O 


86 



sec 19.4 organosilicon compounds 533 

is stronger than the C— O bond by some 22 kcal/mole. These bond energies 
account for several differences in the chemistry of the two elements. Thus, 
while carbon forms a great many compounds having linear and branched 
chains of C— C bonds, silicon is less versatile; the silanes of formula Si„H 2 „ +2 
analogous to the alkanes of formula C„H 2n + 2 are relatively unstable and 
react avidly with oxygen. On the other hand, the silicone polymers have 
chains of Si — O— Si bonds and have a high thermal stability as corresponds to 
the considerable strength of the Si— O bond. 

No compounds containing silicon double bonds of the type 

\ / \ / \ \ 

Si = Si , Si = C , Si=0, or Si=N— 

/ \ / \ / / 

have been prepared to date. Thus, there are no organosilicon compounds that 
are structurally analogous to alkenes, alkynes, arenes, aldehydes, ketones, 
carboxylic acids, esters, or imines. One clear illustration is the formation of 
silanediols of the type R 2 Si(OH) 2 . The silanediols do not lose water to form 
"silicones" of structure R 2 Si=0 in the way the alkanediols, R 2 C(OH) 2 , 
which are normally unstable, lose water to form the corresponding ketones, 
R 2 C=0. Loss of water from silanediols results in formation of Si— O— Si 
bonds, and this is the basic reaction by which silicon polymers are formed. 

R R R 

I -H 2 I I 

2 HO-Si-OH * HO-Si-O-Si-OH 

I I I 

R R R 

In its inability to form the p n -p % type of double bond, silicon resembles 
other second-row elements of the periodic table, such as sulfur and phos- 
phorus. 



A. BONDING INVOLVING d ORBITALS IN ORGANOSILICON 
COMPOUNDS 

Silicon is normally tetracovalent in organosilicon compounds and, by analogy 
with carbon, we may reasonably suppose the bonds involved to be of the sp 3 
type and the substituent groups to be tetrahedrally disposed in space. Evi- 
dence that this is so comes from the successful resolution of several silicon 
compounds having a center of asymmetry at the silicon atom; for example, 
both enantiomers of l-naphthylphenylmethylsilane [5] have been isolated. 





[5] 

M D =+32° 

X-Ray and electron-diffraction studies of silicon tetraiodide, silicon tetra- 
chloride, and tetramethylsilane also indicate tetrahedral structures. Sub- 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 534 

stances with hexacovalent silicon such as hexafluosilicate ion, SiF 6 2e , are 
known, however, and this shows that silicon can expand its valence shell to 
accommodate 10 electrons by utilizing the 3d orbitals. The silicon 3d orbitals 

\ 

may also be involved in the bonds of compounds of the type —Si — X, Where 

X is an atom or group having electrons in a p orbital so situated as to be able 
to overlap with an empty. 3d orbital of silicon. The result would be a Si— X 
bond with partial double-bond character of the d„-p n type, in which the 
silicon has an expanded valence shell. The bonding can be symbolized by 
these resonance structures (examples of X include oxygen, nitrogen, and the 
halogens, as well as unsaturated groups such as the vinyl and phenyl groups) : 

\ •• \ e B 

-Si— X < ► — Si=X 

/ / 



B. PREPARATION AND PROPERTIES OF ORGANOSIL1CON 
COMPOUNDS 

Organosilicon compounds are prepared from elementary silicon or the silicon 
halides. A particularly valuable synthesis of organochlorosilanes involves 
heating an alkyl chloride or even an aryl chloride with elementary silicon in 
the presence of a copper catalyst. A mixture of products usually results; 
nonetheless, the reaction is employed commercially for the synthesis of 
organochlorosilanes, particularly the methylchlorosilanes. 

CH 3 C1 + Si C " ( ^ %) -^ SiCl 4 + CH 3 SiHCl 2 + CH 3 SiCl 3 + (CH 3 ) 2 SiCl 2 

9% 12% 37% 42% 



/ V C1 + Sl samu ( Vsicu + (f Vl-s lC , 



The physical properties of some organosilanes may be seen from Table 
19-4 to be roughly similar to those of the analogously constituted carbon 
compounds. 



C. SILANOLS, SILOXANES, AND POLYSILOXANES 

The silanols are generally prepared by hydrolysis of silyl halides and some- 
times by hydrolysis of hydrides and alkoxides. 

R 3 SiCl + H 2 ► R 3 SiOH + HCl 

R 2 SiCl 2 + 2H 2 >• R 2 Si(OH) 2 + 2 HC1 

e 

OH 

R 3 SiH + H a O > R 3 SiOH + H 2 

The reaction conditions have to be controlled to avoid condensation of the 



sec 19.4 organosilicon compounds 535 

Table 19-4 Physical properties of some representative silicon compounds 
and their carbon analogs 



silicon 


bp, 


mp, 


carbon 


bp, 


mp, 


compound 


°C 


°C 


compound 


°C 


°C 


SiH 4 ° 


-112 


-156.8 


CH 4 


-162 


-183 


SiH 3 SiH 3 " 


-14.5 


-133 


CH 3 CH 3 


-88.6 


-172 


CH 3 SiH 3 


-57.5 


-156.8 


CH 3 CH 3 


-88.6 


-172 


(CH 3 ) 4 Si 


27 




(CH 3 ) 4 C 


9.5 


-20 


(CH 3 ) 3 SiC 6 H 5 


172 




(CH 3 ) 3 CC 6 H 5 


169 


-58 


(C 6 H 5 ) 4 Si 


430 


237 


(C 6 H 5 ) 4 C 


431 


285 


SiCl 4 


57.6 


-70 


CC1 4 


76.8 


-22.8 


SiHCl 3 


33 


-134 


CHC1 3 


61 


-63.5 


SiH 2 Cl 2 


8.3 


-122 


CH 2 C1 2 


40 


-96.7 


CH 3 SiCl 3 


65.7 




CH 3 CC1 3 


74 




C 6 H 5 SiCl 3 


201 




C 6 H 5 CC1 3 


214 


-5 


(CH 3 ) 3 SiOH 


98.6 




(CH 3 ) 3 COH 


82.8 


25.5 


(CH 3 ) 2 Si(OH) 2 




100 









' Spontaneously flammable in air. 



silanols to siloxanes, especially when working with silanediols. This may 
necessitate working in neutral solution at high dilution. 



R 2 Si(OH) 2 + R 2 Si(OH) 2 



-H 2 



R R 

I I 

HO— Si-O— Si-OH 



R 



R 



The silanols are less volatile than the halides and siloxanes because of 
intermolecular association through hydrogen bonding, and the diols, 
R 2 Si(OH) 2 , are more soluble in water than the silanols, R 3 SiOH. Compared 
to alcohols, silanols are more acidic and form stronger hydrogen bonds. This 
can be ascribed to d n -p n bonding of the Si— O bond. 



e © 
R 3 Si=0- 



■H 



R 3 Si-0-H «- 

In the presence of either acids or bases, most silanols are unstable and 
condense to form siloxanes. The ease with which these reactions occur 
compared to corresponding reactions of alcohols is likely to be associated 
with the ease of formation of pentacoordinate silicon intermediates. 



Nu + X— Si- 



Nu \le 
Si- 



z» Nu — Si- + X" 



Acid-catalyzed condensation: 
R 3 SiOH + H® .. R 3 SiOH 2 



R 3 SiO + R 3 Si — OH 2 
H 



R 3 SiOSiR 3 + H 3 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 536 

Base-catalyzed condensation: 




R 3 SiOH + OH , R 3 SiO + H 2 

e e 

R 3 SiO + R 3 Si-OH ► R 3 SiOSiR 3 + OH 

The same type of condensation reaction, when carried out with the silane- 
diols, leads to linear chains and cyclic structures with Si— O and Si— C bonds 
which are called polysiloxanes. 

(CH 3 ) 2 
/ si \ 

H® nr ©OH ^ ^ 

(CH 3 ) 2 Si(OH) 2 " ° r > I I 

(CH 3 ) 2 Si^ /Si(CH 3 ) 2 



+ HO 



The higher-molecular-weight products are the "silicone polymers." 
The linear silicone polymers are liquids of varying viscosity depending on 
the chain length. They remain fluid to low temperatures and are very stable 
thermally, which makes them useful as hydraulic fluids and lubricants. 

Cross-linking results in hard and sometimes brittle resins, depending upon 
the ratio of methyl groups to silicon atoms in the polymer. 



19-5 organoboron compounds 

Boron has three valence shell electrons available for bonding and it utilizes 
these to form trigonal (sp 2 ) bonds in compounds of the type BX 3 . Typical 
examples are boron trifluoride, BF 3 ; trimethylborane, B(CH 3 ) 3 ; and boric 
acid, B(OH) 3 . In such compounds the boron normally has only six electrons 
in three of the four available bonding orbitals and therefore is said to be 
"electron deficient." There is a considerable tendency for boron to acquire an 
additional electron pair to fill the fourth orbital and so attain an octet of 
electrons. The boron halides and the organoboranes (BR 3 ) are Lewis acids 
and may accept an electron pair from a base to form tetracovalent boron com- 
pounds in which the boron atom has a share of eight electrons. 



b1 CH3 3b. 

|Vi20° + :NH 3 ► H 3 C<4"'CH 3 

CH 3 109° CH 3 

trimethylborane ammonia trimethylborane 

(planar) (tetrahedral) 

A change in configuration at boron occurs in these reactions because 



sec 19.S organoboron compounds 537 

tetracovalent boron is tetrahedral (sp 3 hybrid orbitals), whereas tricovalent 
boron is trigonal and planar (sp 2 ). 

A particularly interesting class of compounds is the borazines. These 
compounds are six-membered heterocycles with alternating boron and nitro- 
gen atoms. They are formally analogous to benzene in that there are six 
electrons — one pair at each nitrogen — which could be delocalized over six 
orbitals, one from each boron and nitrogen in the ring. 

H H e H e 

HN: :NH HI^T NH HN ^NH 



I 



HB. ..^,BH HB^ .BH HB. ^BH 

H H 95 H e 

[6a] [6b] [6c] 

The similarity between benzene and borazine is obvious from the Kekule 
structures [6b] and [6c]. The degree to which these structures can be regarded 
as contributing to the actual structure of the borazine molecule, however, has 
been a topic of controversy. The borazine molecule has a planar ring with 
120° bond angles and six equivalent B — N bonds of length 1.44 A, which is 
shorter than the expected value of 1 .54 A for a B— N single bond and longer 
than the calculated value of 1.36 A for a B=N double bond. 



A. MULTICENTER BONDING AND BORON HYDRIDES 

The simple hydride of boron, BH 3 , is not stable, and the simplest known 
hydride is diborane, B 2 H 6 . Higher hydrides exist, the best known of which are 
tetraborane, B 4 Hi ; pentaborane, B 5 H 9 ; dihydropentaborane, B 5 H U ; 
hexaborane, B 6 H 10 ; and decaborane, B 10 H 14 . These compounds are especial- 
ly interesting with regard to their structures and bonding. They are referred to 
as "electron deficient" because there are insufficient electrons with which to 
form all normal electron-pair bonds. This will become clear from the following 
description of the structure of diborane. 

The configuration and molecular dimensions of diborane resemble ethene 
in that the central B— B bond and four of the B— H bonds form a planar 
framework. However, the remaining two hydrogens are centered above and 
below this framework and form bridges across the B— B bond, as shown in 
Figure 19-4. The presence of two kinds of hydrogens in diborane is also con- 
sistent with its infrared and nmr spectra. 

If we try to write a conventional electron-pair structure for diborane, we 
see at once that there are not enough valence electron pairs for six normal 
B— H bonds and one B— B bond. Seven normal covalent bonds require 14 
bonding electrons, but diborane has only 12 bonding electrons. The way the 
atoms of diborane are held together would therefore appear to be different 
from any we have thus far encountered, except perhaps in some carbonium 
ions (Section 4-4B). Nonetheless, it is possible to describe the bonds in 
diborane in terms of electron pairs if we adopt the concept of having three 
(or more) atomic centers bonded by an electron pair in contrast to the usual 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 538 





1.33 k/ \ 




121. 5°?^ 


rflft ' ' "" A 4 




'^ HS ^ v y , ioo J 






ljT^^jjIlfci 



Figure 19-4 Configuration of diborane. 

bonding of two atomic centers by an electron pair. We can formulate diborane 
as having two three-center bonds, each involving an electron pair, the two 
boron atoms, and a bridge hydrogen. 

The three-center bonds can be represented in different ways — one possible 
way being with dotted lines, as in [7]. 



U X V H 



[7] 



The structures of many of the higher boron hydrides, such as B 4 H 10 , B 5 H 9 , 
B 5 H n , B 6 H 10 , and B 10 H 14 , have been determined by electron and(or) 
X-ray diffraction. These substances resemble diborane in having an overall 
deficiency of electrons for the total number of bonds formed unless some are 
formulated as multicenter bonds. 

The structure of pentaborane is shown in Figure 19-5. This molecule has 24 



Figure 19-5 Structure of pentaborane. 



H 




B v 




















i ! \ s 










-H 


.*-!/ \ 1 \ 




H'' // \ / \ 




\ // \ / -H 




\ i' \ i , - 




n -B' 








\T 





sec 19.5 organoboron compounds 539 

valence electrons; of these, 10 can be regarded as utilized in forming five 
two-center B— H bonds (solid lines), and eight in forming four three-center 
BHB bonds (dashed lines). The remaining six electrons can be taken to con- 
tribute to multicenter binding of the boron framework (dashed lines). 

A number of alkylated diboranes are known, and their structures are similar 
to diborane. 

H 3 C x ,h., CH 3 

H 3 C / ~'H''' X CH 3 
.ym-tetramethyldiborane 

However, when a boron atom carries three alkyl groups, three-center 
bonding does not occur, and the compound is most stable in the monomeric 
form. Apparently, alkyl groups are unable to form very strong three-center 
bonds with boron. 



B. NOMENCLATURE OF ORGANOBORON COMPOUNDS 

The literature on organoboron compounds is inconsistent as to nomenclature, 
and many compounds are called by two or more names. For example, the 
simple substance of formula B(CH 3 ) 3 is called variously trimethylborane, 
trimethylborine, or trimethylboron. Insofar as possible, we will name organo- 
boron compounds as derivatives of borane, BH 3 . Thus, substitution of all the 
B — H hydrogens by methyl groups gives B(CH 3 ) 3 , trimethylborane. Several 
more examples follow: 



B(C 6 H 5 ) 3 


B 2 H 6 


C6H5BG2 


triphenylborane 


diborane 


phenyldichloroborane 


(CH 3 ) 2 BCi 


© e 
(CH 3 ) 3 N-BH 3 


(CH 3 ) 2 N-BH 2 


dimethylchloroborane 


trimethylamine borane 


dimethylaminoborane 



C. HYDROBORATION 

Trialkylboranes can be prepared by addition of boron hydrides to multiple 

bonds. In general, these reactions of boron hydrides conform to a pattern of 

e e 
cleavage of B— H bonds in the direction B: H, with boron acting as an electro- 

phile. 

\0\ / \ f> I I 

C = C + B — H _ - - —C — C — 

/ \ / II 

Addition of a B— H linkage across the double bond of an alkene is known 
as hydroboration and has been discussed earlier (Section 4-4F). Cleavage of 
the B— C bond with a carboxylic acid gives an alkane, with alkaline hydrogen 
peroxide, an alcohol. (Note that the alcohol that is formed is the anti- 
Markownikoff product.) 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 540 
CH 3 CH=CH 2 + BH 3 ► CH 3 CH 2 CH 2 BH 2 CH3CH=CH * . (CH 3 CH 2 CH 2 ) 2 BH 



(CH 3 CH 2 CH 2 ) 3 B 
RCO^H^/ H0 e\H 2 2 

CH 3 CH 2 CH 3 CH 3 CH 2 CH 2 OH 



summary 

Sulfur, phosphorus, and silicon, being second-row elements in the periodic 
table, have low-lying d orbitals available and are not restricted to an octet of 
bonding electrons as are the elements in the first row. Sulfur can have as many 
as 12 bonding electrons (six covalent bonds), and phosphorus 10 bonding 
electrons (five covalent bonds). Silicon is usually tetravalent although even 
here there is often some d orbital participation in bond formation. Boron is 
unique in having only three electrons in its valence shell. Neutral boron can 
form either three covalent bonds and be electron deficient or four covalent 
bonds and be anionic, as in BH 4 e . Many compounds are known with three- 
center bonds between two boron atoms and a hydrogen atom. 

Sulfur in its divalent state is analogous to oxygen. Thiols, also called 
mercaptans (RSH), correspond to alcohols (ROH); sulfides, also called 
thioethers (RSR), correspond to ethers (ROR); and disulfides (RSSR) cor- 
respond to peroxides (ROOR). There are no oxygen analogs for the sulfur 
compounds with four and six bonds, the most important of which are sul- 

o o o 

II II II 

foxides, R— S— R, sulfones, R— S— R, and sulfonic acids, R— S— OH. 

O O 

Thiols are more acidic than alcohols and also more nucleophilic. Oxida- 
tion of thiols produces disulfides. 

RSH [ ° ] > RS-SR (disulfide) 

Alkylation of thiols, RSH, gives sulfides, RSR, which can be converted to 
sulfonium salts, R 3 S®X e . If the R groups are all different the salts can be 
resolved into optical enantiomers. 

Oxidation of sulfides gives sulfoxides and sulfones. 





O 

jj 




^. R-S-R (sulfoxide) 


R-S-R 


O 

^» R— S— R (sulfone) 




1! 

o 



Dimethyl sulfoxide (DMSO) is an example of a polar aprotic solvent. It 



summary 541 

enhances the reactivity of small anions and is often used as a solvent for S N 2 
reactions. It is also an excellent mild oxidant for alcohols. 

There are three kinds of organosulfur acids: sulfenic acids (RSOH), 
sulfinic acids (RS0 2 H), and sulfonic acids (RS0 3 H). The last are strong 
acids, and are the most important of the series. They will be met again in the 
next chapter. 

O 

II 
Sulfate esters, RO— S— OR, lack a carbon-sulfur bond and hence are not 
II 
O 

classed as organosulfur compounds. They are useful alkylating agents. 

Organic compounds of phosphorus include phosphate esters, R— O— 
PO(OH) 2 , phosphonic acids, R— PO(OH) 2 , and organophosphines, R 3 P. 
Phosphate esters may exist as polyphosphates, an important example of 
which is adenosine triphosphate, ATP. 

Phosphines can form phosphonium-salts, R 4 P®X e , analogous to ammo- 
nium salts. When treated with phenyllithium, a methylphosphonium ion loses 
a proton to give a methylenephosphorane. 

© -H® 

R 3 P— CH 3 ► R 3 P=CH 2 (methylenephosphorane) 

These compounds can convert ketones to ethene derivatives as a result of 
the exchange of a carbonyl and a methylene group (the Wittig reaction). 

R 3 P=CH 2 + R' 2 C=0 > R' 2 C=CH 2 + R 3 P=0 

Silicon, like carbon, has four valence electrons but, unlike carbon, it forms 
only single bonds. Compounds containing only silicon and hydrogen (silane 
SiH 4 , disilane Si 2 H 6 , etc.) ignite spontaneously in air. Alkylsilanes, such as 
R 4 Si, are more stable. Silanediols, R 2 Si(OH) 2 , can be prepared whereas their 
carbon analogs can not. They are made from alkyl chlorides via the chloro- 
silanes. 

RCl + Si — ^—> R 2 SiCl 2 (+ other chlorosilanes) 

H 2 

► R 2 Si(OH) 2 



Silanediols tend to polymerize to form siloxanes, compounds of extremely 
high thermal stability. 

R R R 

I I I 

R 2 Si(OH) 2 ► -O-Si-O-Si-O— Si— O- 

I I I 

R R R 

Boron possesses only three valence electrons and as a result its tricovalent 
compounds, such as BF 3 and B(CH 3 ) 3 , are electron deficient and highly 
reactive as electron-pair acceptors — that is, Lewis acids. When an electron 
pair is accepted an octet arrangement is achieved, for example, BF 4 e . In the 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 542 

borazines n bonding tends to complete the octet of boron. 
H H 

I | < > e l He (borazine) 

,B. .. .B. / B: =%x^/ B \ 

H H 

Diborane and higher boranes are constructed with three-center bonds, each 
consisting of an electron pair, two boron atoms, and a hydrogen bridge. 

Hydroboration is the addition of boron hydrides, particularly BH 3 , to 
multiple bonds. Treatment of an alkene with diborane results in the addition 
of BH 3 to the double bond. The reaction continues and finally gives a tri- 
alkylborane. These compounds can then be converted to alkanes or alcohols. 



rco^h^, RCH 2 CH 3 
RCH=CH 2 — 1 -^ (RCH 2 CH 2 ) 3 B 



H 2 2 , OH e 



exercises 

19-1 Write structural formulas for the following substances: 

a. di-s-butyl thioketone e. tris-(methylsulfonyl)-methane 

b. ethyl sulfide /. trimethylene disulfide 

c. methyl thioacetate g. 5-thia-l,3-cyclopentadiene 

d. /3,/3'-dichloroethyl sulfide (thiophene) 
(mustard gas) 

19-2 Formulate each of the following substances in terms of electronic structure, 
types of bonds (i.e., a and n) and probable molecular geometry (i.e., linear, 
angular, planar, pyramidal, etc.) with rough estimates of bond angles. 

/. S 6 (six-membered ring) 

g. CS 2 

h. SOCl 2 

/. S s (eight-membered ring) 



19-3 Thiols are unlike alcohols in that they do not react readily with hydrogen 
bromide to yield bromides. Explain how a difference in behavior in this 
respect might be expected. 

19-4 Write mechanisms for the conversion of a thiol to a disulfide by oxidation 
with air or iodine which are in accord with the observation that the reaction 
with either oxidizing agent is accelerated by alkali. 

19-5 Dimethyl sulfide reacts with bromine in the absence of water to produce a 



a. 


H 2 S 


b. 


so 2 


c. 


SF 4 


d. 


H 2 S0 4 


e. 


Na 2 SO : 



exercises 543 

crystalline addition compound which reacts with water to produce dimethyl 
sulfoxide. What is the likely structure of the addition compound and the 
mechanism of its formation and reaction with water ? 

19-6 How many and what kinds of stereoisomers would you expect for each of the 
following compounds? 

a. methylethyl-s-butylsulfonium bromide 

b. [CH3(C 2 H 5 )SCH 2 CH 2 S(CH3)C 2 H 5 ] 2 ® 2Br e 

H 2 C— CH 2 

© / \ ® - 

c. CH 3 S SCH 3 2Br e 

3 \ / 3 
H 2 C-CH 2 

19-7 Unsymmetrically substituted sulfoxides, but not the corresponding sulfones, 
exhibit optical isomerism. Write structures for the stereoisomers you 
would expect for 

a. methyl ethyl sulfoxide 

b. the disulfoxide of 1,3-dithiacyclohexane 

19-8 Write equations for a practical synthesis of each of the following substances 
based on the specified starting materials. Give reagents and approximate 
reaction conditions. 

O 

II 

a. CH 3 -C-S-CH 2 CH 2 CH 3 from «-propyl alcohol 

b. optically active methylethyl-«-butylsulfonium bromide from «-butyl 
alcohol 

c. neopentanethiol from neopentyl chloride 

19-9 Suppose it were desired to study the addition of bromine to an alkene double 
bond in water in the absence of organic solvents. A possible substrate for 
this purpose would be CH 2 =CHCH 2 CH 2 CH 2 CH 2 -X, where -X is a 
solubilizing group. What would be the merits of having —X = — S0 3 e Na® 

over X= -OH, -NH 2 , -C0 2 e Na®, -N(CH 3 )3Br e ? 

19-10 Give for each of the following pairs of compounds a chemical test, prefer- 
ably a test tube reaction, which would serve to distinguish one from the 
other. 

a. CH 3 CH 2 SH and CH 3 SCH 3 

b. CH 3 S(0)OCH 3 and CH 3 CH 2 S0 3 H 

c. CH 3 S(0)OCH 3 and CH 3 S(0) 2 CH 3 

d. CH 3 SCH 2 CH 2 OH and CH 3 OCH 2 CH 2 SH 

19-11 Name the following compounds: 

a. CH 3 PHCH 2 CH 3 

b. CH 3 P(C 5 H 5 ) 2 

II 
O 



chap 19 organic compounds of sulfur, phosphorus, silicon, and boron 544 

c. P(OC 2 H 5 ) 3 

d. 0=P(OC 2 H 5 ) 3 

e. / Vp(CH 3 ) 3 B°r 



<2 



/• (CH 3 ) 3 P 



19-12 Write structures for the following compounds : 

a. methyl diphosphate c. diisopropyl methylphosphonate 

b. tricyclohexyl phosphate d. tetraphenylphosphonium iodide 

19-13 Using the Wittig reaction (Section 19-3C) in one of the steps, show how you 
could convert 1,1-dicyclopropylpropene to 1,1-dicyclopropylethene. 

19-14 Name the following compounds according to the nomenclature used in 
Table 19-2. 

a. (C 6 H 5 ) 4 Si c. (C 6 H 5 ) 2 Si(OH) 2 

b. (C 2 H 5 ) 2 SiBr 2 d. (CH 3 ) 3 SiOSiH 3 

19-15 Write resonance structures for trimethylsilanol and vinylsilane involving 
silicon d orbitals, that is, with five bonds to silicon. 

19-16 Explain why trisilylamine, (SiH 3 ) 3 N, is a weaker base than trimethylamine 
and why trimethylsilanol, (CH 3 ) 3 SiOH, is a stronger acid than ?-butyl 
alcohol. 

19-17 With reference to the discussion of reactivity of silicon compounds in Sec- 
tion 19-4C, explain how it is possible for the following reaction to occur 

e 
readily by the rate law, v = fc[R 3 SiCl][OH]. 




o + * 



-Si 
CI OH 

Refer also to Exercise 8-15. 

19-18 Which compound in each of the following pairs would you expect to be 
more stable? Give your reasons. 

a. B,B,B-trimethylborazine or N,N,N-trimethylborazine 

b. borazine or B,B,B-trichloroborazine 

c. B-methoxyborazine or B-(trifluoromethyl)-borazine 

d. ammonia complex of (CH 3 ) 2 B— N(CH 3 ) 2 or the ammonia complex of 
(CH 3 ) 2 B-P(CH 3 ) 2 



exercises 545 

19-19 Write names for the following compounds: 

a. CH 2 =CH-CH 2 -B(CH 3 ) 2 

b. (C 2 H 5 ) 2 BHBH 3 

Oe e 
^-BH 3 

d. (CH 3 ) 2 N-B(CH 3 ) 2 

CI 

I 

/ B \ 

e. HN NH 

I I 

.B- ^B. 

H 

C 6 H 5 

^ B \ 

f. o o 

I I 

/B. .B. 
QH^ ^O^ ^C 6 H 5 

19-20 Write structures for the following: 

a. di-«-butyl-(p-dimethylaminopheny])-borane 

b. di-p-tolylchloroborane 

c. dichloromethoxyborane 

d. tri-(dimethylphosphino)-borane 

19-21 Would you expect boron-phosphorus analogs of borazines to have aromatic 
character ? 

19-22 Which would you expect to form a more stable addition compound with 
ammonia, trivinylborane or triethylborane ? What all-carbon system has an 
electronic structure analogous to dimethylvinylborane ? 



mx- «"•*'«£>■: &-i 



L^^&Si- 







arenes. electr#pjulic aromatic 



chap 20 arenes. electrophilic aromatic substitution 549 

Benzene, C 6 H 6 , and the other aromatic hydrocarbons usually have such 
strikingly different properties from typical open-chain conjugated polyenes, 
such as 1,3,5-hexatriene, that it is convenient to consider them as a separate 
class of compounds called arenes. In this chapter we shall outline their salient 
features, and in subsequent chapters we shall discuss the chemistry of their 
halogen, oxygen, and nitrogen derivatives. 

Some of the important properties of benzene were discussed in Chapter 6 
in connection with the resonance method. Most noteworthy is the fact that 
benzene has a planar hexagonal structure in which all six carbon-carbon 
bonds are of equal length (1.397 A), and each carbon is bonded to one hydro- 
gen. If each carbon is considered to form sp 2 -<7 bonds to its hydrogen and 
neighboring carbons, there remain six electrons, one for each carbon atom, 
which are termed % electrons. These electrons are not to be taken as localized 
in pairs between alternate carbon nuclei to form three conventional con- 
jugated bonds. Rather, they should be regarded as delocalized symmetrically 
through the p z orbitals of all six carbons (Section 6-1). The bonds between the 
carbons are therefore neither single nor double bonds. In fact, they are 
intermediate between single and double bonds in length. 

However, there is more to the bonding than just the simple average of 
C— C single and double bonds, because benzene C— C bonds are substan- 
tially stronger than the average of the strengths of single and double bonds. 
This is reflected in the heat of combustion of benzene, which is substantially 
less than expected on the basis of bond energies ; the extra stability of benzene 
by virtue of its stronger bonds is what we have called its stabilization energy. 
A large part of this stabilization energy can be ascribed to derealization or 
resonance energy of the six carbon -n, electrons. 

The choice of a suitable and convenient graphical formula to represent the 
structure of benzene presents a problem, since the best way to indicate de- 
localized bonding electrons is by dotted lines, which are quite time consuming 
to draw. Dotted-line structural formulas are preferred when it is necessary to 
show the fine details of an aromatic structure — as when the degree of bonding 
is not equal between different pairs of carbons, in phenanthrene, for example. 
Shorthand notations are usually desirable, however, and we shall most often 
use the conventional hexagon with alternating single and double bonds (i.e., 
Kekule cyclohexatriene) despite the fact that benzene does not possess 
ordinary double bonds. Another and widely used notation for benzene is a 
hexagon with an inscribed circle to represent a closed shell of n electrons. 
However, as we have mentioned before (Section 6-1), this is fine for benzene 
but can be misleading for polynuclear hydrocarbons (Section 20- IB). 



20-1 nomenclature of arenes 

A. BENZENE DERIVATIVES 

A variety of substituted benzenes are known with one or more of the hydrogen 
atoms of the ring replaced with other atoms or groups. In almost all of these 



chap 20 arenes. electrophilic aromatic substitution 550 

compounds, the special stability associated with the benzene nucleus is 
retained. A few examples of " benzenoid " hydrocarbons follow, and it will be 
noticed that the hydrocarbon substituents include alkyl, alkenyl, alkynyl, and 
aryl groups. 



CH 



CH, 



CHXH, 




CH, 



CH=CH, 



toluene ethylbenzene cumene styrene 

(methylbenzene) (isopropylbenzene) (vinylbenzene) 



C = CH 




5 6 6' 5' 



phenylacetylene biphenyl 

(ethynylbenzene) (phenylbenzene) 



r\ rH fi 



CH 



diphenylmethane 



The naming of these hydrocarbons is fairly straightforward. Each is named 
as an alkyl, alkenyl, or arylbenzene, unless for some reason the compound has 
a trivial name. The hydrocarbon group (C 6 H 5 — ) from benzene itself is called 
a phenyl group and is sometimes abbreviated as the symbol cf> or as Ph. 
Aryl groups in general are often abbreviated as Ar. Other groups that have 
trivial names include 




CYcH. 



benzyl 



,- ^„ o- 

benzal benzo 



x ' • \=/ I \=J 

benzhydryl 



When there are two or more substituents on a benzene ring, position isom- 
erism arises. Thus, there are three possible isomeric disubstituted benzene 
derivatives according to whether the substituents have the 1,2, 1,3, or 1,4 
relationship. The isomers are commonly designated as ortho, meta, and para 
(or o, m, and/?) for the 1,2, 1,3, and 1,4 isomers, respectively. The actual sym- 
metry of the benzene ring is such that only one 1,2-disubstitution product is 
found despite the fact that two would be predicted if benzene had the 1,3,5- 
cyclohexatriene structure. 

x (X-CX 




not 





CH, 



CH, 



ortho-xylene tneta-xylene para-xylene 

(1 ,2-dimethylbenzene) ( 1 ,3-dimethy lbenzene) ( 1 ,4-dimethylbenzene) 



sec 20.1 nomenclature of arenes SSI 



CH 3 



CH 3 




■.-***• 



o-bromotoluene 





B. POLYNUCLEAR AROMATIC HYDROCARBONS 

A wide range of polycyclic aromatic compounds are known that have benzene 
rings with common ortho positions. The parent compounds of this type are 
usually called polynuclear aromatic hydrocarbons. Three important examples 
are naphthalene, anthracene, and phenanthrene. In anthracene, the rings are 
connected linearly, while in phenanthrene they are connected angularly. 




naphthalene 



5 10 * 

anthracene 



&5 43 
phenanthrene 



There are two possible monosubstitution products for naphthalene, three 
for anthracene, and five for phenanthrene. The accepted numbering system 
for these hydrocarbons is as shown in the formulas; however, the 1 and 2 
positions of the naphthalene ring are frequently designated as a and /?. Some 
illustrative substitution products are shown. 



CH, 





CH, 




1 -methylnaphthalene 2-methylnaphthalene 
(a-methylnaphthalene) (/?-methylnaphthalene) 



1 -methylanthracene 



Substances that can be regarded as partial or complete reduction products 
of aromatic compounds are often named as hydro derivatives of the parent 
system — the completely reduced derivatives being known as perhydro com- 
pounds. 







9,10-dihydro- 


1,2,3,4-tetrahydro- 


decahydro- 


perhydro- 


anthracene 


naphthalene 


naphthalene 


phenanthrene 




(tetralin) 


(decalin) 





The names that have been given to the more elaborate types of polynuclear 
aromatic hydrocarbons are for the most part distressingly uninformative in 
relation to their structures. (A thorough summary of names and numbering 



chap 20 arenes. electrophilic aromatic substitution S52 

systems has been published by A. M. Patterson, L. T. Capell, and D. F. 
Walker, " Ring Index," 2d Ed., American Chemical Society, 1960.) Two such 
compounds are shown here, with their systematic names given in parentheses. 





naphthacene 
(benz[b]anthracene) 



pyrene 
(benz[d,e,f]phenanthrene) 



These are named as derivatives of simpler polynuclear hydrocarbons such as 
naphthalene or phenanthrene, to which are fused additional rings. An extra, 
ring, which adds only four carbons at the most, is designated by the prefix 
benzo 1 (or benz if followed by a vowel), and its positions of attachment either 
by numbers or, as shown, by letters. (The sides of the parent compound are 
lettered in sequence beginning with the 1 ,2 side, which is a.) 

One can insert a Kekule structure in any of the rings in the above com- 
pounds and be reasonably confident that alternating double and single bonds 
can be placed in the remaining rings. This is not true for the compound 
phenalenyl, a rather reactive hydrocarbon of formula C 13 H 9 (Figure 20-1). 
Note that the Kekule formula [1] reveals that one of the carbon atoms in the 
molecule has only three bonds to it whereas formula [2] does not. There 
are 10 contributing resonance structures for this molecule, each containing a 
three-bonded carbon atom having one unpaired electron. The molecule is 
thus a radical. The various Kekule structures of other polynuclear hydro- 
carbons are discussed in Section 20-6. 



1 Confusion between the two meanings of benzo (see Section 20-1 A) seldom arises since 

I 
it is usually clear from the context whether it refers to a C 6 H 5 C— group or a fused ring. 



Figure 2(M Two ways [1] and [2] of representing phenalenyl, C 13 H 9 . The 
skeleton showing only the location of the carbon and hydrogen atoms is also 
given. 





H 


1 


C^ 


C 


^ C \ c / C \ c/ H 


C 


! 

H 


- Cv C^H 
H 



sec 20.2 physical properties of arenes 553 

Many polynuclear hydrocarbons, like many aromatic amines (Section 
22-9B), are carcinogens. 



20-2 physical properties of arenes 

The pleasant odors of the derivatives of many arenes are the reason they are 
often called aromatic hydrocarbons. The arenes themselves, however, are 
generally quite toxic and inhalation of their vapors should be avoided. The 
volatile arenes are highly flammable and burn with a luminous, sooty flame, 
in contrast to alkanes and alkenes, which burn with a bluish flame leaving 
little carbon residue. 

A list of common arenes and their physical properties is given in Table 20- 1 . 
They are less dense than water and are highly insoluble. Boiling points are 
found to increase fairly regularly with molecular weight, but there is little 
correlation between melting point and molecular weight. The melting point is 
highly dependent on the symmetry of the compound; benzene thus melts 
100° higher than toluene, and the more symmetrical p- xylene has a higher 
melting point than either the o~ or the w-isomer. 



Table 20-1 Physical properties of arenes 



compound 


mp, 

°C 


bp, 

°C 


density, 

dl° 


benzene 


5.5 


80 


0.8790 


toluene 


-95 


111 


0.866 


ethylbenzene 


-94 


136 


0.8669 


rc-propylbenzene 


-99 


159 


0.8617 


isopropylbenzene 


-96 


152 


0.8620 


(cumene) 








/-butylbenzene 


-58 


168 


0.8658 


o-xylene 


-25 


144 


0.8968 


m-xylene 


-47 


139 


0.8811 


p-xylene 


13 


138 


0.8541 


mesitylene 


-50 


165 


0.8634 


(1 ,3,5-trimethylbenzene) 








durene 


80 


191 




(1,2,4,5-tetramethylbenzene) 








naphthalene 


80 


218 




anthracene 


216 


340 




phenanthrene 


101 


340 





chap 20 arenes. electrophilic aromatic substitution 554 

20-3 spectroscopic properties of arenes 



A. INFRARED SPECTRA 

The presence of a phenyl group in a compound can be ascertained with a 
fair degree of certainty from its infrared spectrum. Furthermore, the number 
and positions of substituent groups on the ring can also be determined from 
the spectrum. For example, in Figure 20-2, we see the individual infrared 
spectra of four compounds: toluene and o-, m-, and /^-xylene. That each 
spectrum is of a benzene derivative is apparent from certain common features, 
notably the two bands near 1600 cm -1 and 1500 cm - - 1 which, although of 
variable intensity, have been correlated with the stretching vibrations of the 
carbon-carbon bonds of the aromatic ring. In some compounds, there is an 
additional band around 1580 cm -1 . The sharp bands near 3030 cm -1 are 
characteristic of aromatic C— H bonds. Other bands in the spectra, especially 
those between 1650 and 2000 cm -1 , between 1225 and 950 cm -1 , and below 
900 cm -1 , have been correlated with the number and positions of ring sub- 
stituents. Although we shall not document all these various bands in detail, 
each of the spectra in Figure 20-2 is marked to show some of the types of 
correlations that have been made. 



B. ELECTRONIC ABSORPTION SPECTRA 

Compared to straight-chain conjugated polyenes, aromatic compounds have 
relatively complex absorption spectra with several bands in the ultraviolet 
region. Benzene and the alkylbenzenes possess two bands in which we shall be 
primarily interested, one lying near 2000 A and the other near 2600 A. 

The 2000 A band is of fairly high intensity and corresponds to excitation of 
a % electron of the conjugated system to a %* orbital (i.e., a n ~* n* transition). 
The excited state has significant contributions from dipolar structures such as 
[3]. This is analogous to the absorption bands of conjugated dienes (Section 



[3] 



7-5) except that the wavelength of absorption of benzene is shorter. In fact, 
benzene and the alkylbenzenes absorb just beyond the range of most com- 
mercial quartz spectrometers. However, this band (which we say is due to the 
benzene chromophore) is intensified and shifted to longer wavelengths when 
the conjugated system is extended by replacement of the ring hydrogens by 
unsaturated groups (e.g., -HC=CH 2 , -C^CH, -HC=0, and -C=N; 
see Table 20-2). The absorbing chromophore now embraces the electrons of 
the unsaturated substituent as well as those of the ring. In the specific case 



sec 20.3 spectroscopic properties of arenes SSS 



■ 

"■* r~ ^ — - ' 
















A" " " "*~ 




: — ■ : X = " 


o 
o 


















L 




- *§• 












<^_ 






o 


o J- ^-~^r;:r 






o 








o 


1 ■«>«—.. m „..— .yp**^ 


i - 


— 


— 


o-t -CI.—-— -— — - 








*s» - 






o 


.jr 






o 


t_ 






(N 










1 < ^" 
















X__. 








... L « — 




-• — 


o 
o 










a. _____ 




rfr^H^^- 


T 


-' ! ^-, 






o£ 


fi -Ok C 


4r 


~""N / 


o o 
- >> 






O -zj 


o 

c 

8E 


« ==. - 




-' . _, 


CO Cr 


i *"i, ^ 




<~> ts 


— u 




C 




<H-< 


Af" 


j^ 


illilll 


8 

o 

o 
o 








o 


< 






cn 


? 






o 
o 


•»r i 








c 




Ift^fc^ffittWi^ftB 


o 








o 

CO 

o 








o 


"Ii 




^^^^tff^^fi 


8 


1 r 






« 



1 
















1 ^. 


























T 




CO 


1 V 












S 




4 <" 










a 




o 




~~I_^=r=- 


=- 










o 


aX. C= — 


=^_ 
















— 










1 ° 

r 





























Hi o 


i 


*— — ^ 










1 "* 


a. 1 




-=^_ 








J 1 


J3 , j^~ 








T 




BOU 
















C of- .jT~^ 

- r c 


~^=- -« 










pi CJ 


2 L 
























lo3 














II Or? 


& CL 




'w 


s 






1"* 


1 .r"" 


^ 




!>' 






■j 
















"Tj* 




w 


i» 






J* 


t— 








-iwL 




o 


10 


4. 








<u 




1 ^" 


"* 


i_ 






— 


y 


Z3S2 


111 o 

PI CN 

11 o 
J o 
















Is 

1"' 



UOISSIUISUBI} <■ 



UOISSlUlSUBll • 



rv i— 




, — — — — — .^ 






=1 


___-**- 


__=== 


== 


"a 1 *L, 


^T £ 


^. 






o r; 




























">> 


















t* 


', S5 


2 ! '----■= 




w 


15 


■ 8a 


* ^-~ 


5 




O 


l co cr 










I £ 




=~ 5 


W 


ca 


■Xc^~ 








§ 








H ° 


' 1 




H 




[CN 






V 
























- o 




















** 




o 
o 




L_ 


.... 


-*>^, — _ 




CN 

- o 

i 1 r. 

-Is 


m 


1 








r 



"" 


<r - — 


r==- 




i ° 


o 


["' f " ' 










1 C^~ 


= 




o 


CO 












1 ~^-^. 


■ — — 




o 


N. 


— 


v^__" 




lw 


a. 






"T" 


; | 


x 




^- 


op 












bo 


(f~ 




\ 1 




> 


















03 


£ 


c__^ 


-Ss 


\^ ^, 


co Cr 




^ 


S g 


-■1 5 


i} 


>o 


.1^* 


X 


^ „ 


1 

O 

O 


in 


/ 






O 
O 




fc 








n 


-f 




^^^^^B 


8 
•0 



UOISSIUISUBI} % 



UOISSIUISUBI} (■ 



chap 20 arenes. electrophilic aromatic substitution 556 
Table 20-2 Effect of conjugation on the ultraviolet spectrum of the benzene chromophore 



o 

benzene 



/ \-CH=CH 2 
styrene 



/ V-CH— O 
benzaldehyde 



biphenyl 



stilbene 



A m „,A 1980 
£ ma , 8000 



2440 
12,000 



2440 
15,000 



2500 
18,000 



2950 
27,000 



Table 20-3 Effect of substituents on the ultraviolet spectrum of the benzene 
chromophore 



benzene phenol 



phenoxide ion iodobenzene aniline 






1980 
8000 



2100 
6200 



2350 
9400 



2260 
13,000 



2300 
8600 



of styrene, the excited state is a hybrid structure, composite of [4a] and [4b] 
and other related dipolar structures. 




C 




etc. 



[4b] 

Similar effects are observed for benzene derivatives in which the substituent 
has unshared electron pairs in conjugation with the benzene ring (e.g., 
— NH2 > — OH, —CI:). An unshared electron pair is to some extent delocal- 
ized to become a part of the aromatic 7i-electron system in both the ground 
and excited states, but more importantly in the excited state. This may be illus- 
trated for aniline by the following structures, which can be regarded as con- 
tributing to the hybrid structure of aniline. (The data of Table 20-3 show 
the effect on the benzene chromophore of this type of substituent.) 



NH, 





®NH, 





As already mentioned, the benzene chromophore gives rise to a second 
band at longer wavelengths, as shown in Figure 20-3. This band, which is of 
low intensity, is found to be a composite of several equally spaced (1000 



sec 20.3 spectroscopic properties of arenes 557 



Table 20-4 The effect of substituents on absorption corresponding 
to the benzenoid band 



o 

benzene 



o- o 



CH=CH 2 



toluene 



styrene 



O^ 1 

iodobenzene 



Q-NH 2 



aniline 



''max » A 



' v max 
c-max 



2550 
230 



2610 
300 



2820 
450 



2560 
800 



2800 
1430 



Table 20- S Benzenoid band of linear polycyclic aromatics 




benzene naphthalene anthracene 



naphthacene 



A m „,A 2550 



3140 
316 



3800 
7900 



4800 
11,000 



5800 
12,600 



cm" 1 ) narrow peaks. It is remarkably characteristic of aromatic hydro- 
carbons, for no analogous band is found in the spectra of conjugated polyenes. 
For this reason, it is often called the benzenoid band. The position and intensity 
of this band, like the one at shorter wavelengths, are affected by the nature of 
the ring substituents, particularly by those which extend the conjugated 
system, as may be seen from the data in Table 20-4 and Table 20-5. 



Figure 20-3 Ultraviolet absorption spectrum of benzene (in cyclohexane) 
showing the "benzenoid" band. 




2300 



2400 2500 2600 
wavelength, A *- 



chap 20 arenes. electrophilic aromatic substitution 558 

C. NUCLEAR MAGNETIC RESONANCE SPECTRA 

The chemical shifts of aromatic protons (6.5 to 8.0 ppm) are characteristi- 
cally toward lower magnetic fields than those of protons attached to ordinary 
double bonds (4.6 to 6.9 ppm). The difference is usually about 2 ppm and 
has special interest because we have already formulated the hydrogens in both 
types of systems as being bonded to carbon through sp 2 -<7 bonds. 

In general, the spin-spin' splittings observed for phenyl derivatives are 
extremely complex. An example is given by nitrobenzene (Figure 20-4), which 
has different chemical shifts for its ortho, meta, and para hydrogens and six 
different spin-spin interaction constants: / 23 , J 24 , J 2 s, J26' ^34 > ^35 ( tne 
subscripts correspond to position numbers of the protons). 




Such a spectrum is much too complex to be analyzed by any very simple 
procedure. Nonetheless, as will be seen from Exercise 20-7, nuclear magnetic 
resonance can be useful in assigning structures to aromatic derivatives, 
particularly in conjunction with integrated line intensities and approximate 
values of the coupling constants between the ring hydrogens, as shown here. 






Figure 20-4 Nuclear magnetic resonance spectrum of nitrobenzene at 60 
MHz with reference to tetramethylsilane at 0.00 ppm. 




_ p— . . p . , . 

500 450 400 Hz 



ill 



/! '1* It - 



'i'i iii. 



in n ] 

> 









8.0 7.0 ppm 



sec 20.4 reactions of aromatic hydrocarbons 559 

20-4 reactions of aromatic hydrocarbons 



A. ELECTROPHILIC AROMATIC SUBSTITUTION 

In this section we shall be mainly interested in the reactions of arenes that 
involve attack on the aromatic ring. We shall not at this point elaborate on 
the reactions of substituent groups around the ring, although, as we shall see 
later, these and reactions at the ring are not always independent. 

The principal types of reactions involving aromatic rings are substitution, 
addition, and oxidation. Of these, the most common are electrophilic 
substitution reactions. A summary of the more important substitution reac- 
tions of benzene is given in Figure 20-5 and includes halogenation, nitration, 
sulfonation, alkylation, and acylation. 

There are certain similarities between the aromatic substitution reactions 
listed in Figure 20-5 and electrophilic addition reactions of alkenes (Section 
4-4). Indeed, many of the reagents that commonly add to the double bonds of 
alkenes also substitute an aromatic nucleus (e.g., Cl 2 , Br 2 , H 2 S0 4 , HOC1, 
HOBr). Furthermore, both types of reaction are polar, stepwise processes 
involving electrophilic reagents. The key step for either is considered to be the 
attack of an electrophile at carbon to form a cationic intermediate. We may 
represent this step by the following general equations in which the attacking 

reagent is represented either as a formal cation, X®, or as a neutral but 

a® se 
polarized X— Y molecule. 

Electrophilic aromatic substitution (first step) : 



5e Se 
(or X - Y) 



Electrophilic addition to alkenes (first step) : 

Se Se e 

H 2 C=CH 2 + X® (or X - Y) ► H 2 C-CH 2 X 

The intermediate depicted for aromatic substitution no longer has an 
aromatic structure; rather, it is an unstable cation with four n electrons 
delocalized over five carbon nuclei, the sixth carbon being a saturated carbon 
forming sp 3 -hybrid bonds. It may be formulated in terms of the following 
contributing structures, which are assumed here to contribute essentially 
equally. (Note that the partial charges are at three positions, the two ortho 

XH XH XH XH X^ H 



« > 



positions and the para position.) 

Loss of a proton from this intermediate to Y e results in regeneration of an 
aromatic ring, which is now a substitution product of benzene. 



chap 20 arenes. electrophilic aromatic substitution 560 
Figure 20-5 Typical benzene substitution reactions. 





HN0 3 , H 2 SO» 


Qhno 2 








nitration 






nitrobenzene 






Br 2 , FeBr 3 


/ Y-Br + HBr 


bromination 








bromobenzene 






CI 2 , FeCl 3 


/ V-Cl + HC1 


chlorination 








chlorobenzene 




' 1 1 


HOC1, H®, Ag® 

/ ' 


O-c, 

chlorobenzene 


chlorination 





cone. H2SO4 


/ VS0 3 H + H 2 


sulfonation 




., \ \ 




benzenesulfonic acid 






\ CH 3 CH 2 C1 . 


/^-CH 2 CH 3 


alkylation 


AICI3 






ethylbenzene 






\ CH 3 CH=CH 2 
H 3 P0 4 


f>CH /H ' 

isopropylbenzene 
(cumene) 


alkylation 




\ O 

\ " 


O 






\ CH 3 CC1 


(yi- CH , 


acylation 


AICI3 






methyl phenyl ketone 
(acetophenone) 






\ D 2 SQ 4 


<y° 


deuteration 








monodeuteriobenzene 





sec 20.4 reactions of aromatic hydrocarbons 561 

Electrophilic aromatic substitution (second step) : 
X H ? 



The gain in stabilization attendant on regeneration of the aromatic ring is 
sufficiently advantageous that this, rather than combination of the cation 
with Y e , is the actual course of reaction. Here is the difference between aro- 
matic substitution and alkene addition. With alkenes, there is usually no 
substantial resonance energy to be gained by loss of a proton from the inter- 
mediate, which tends instead to react by combination with a nucleophilic 
reagent. 

Electrophilic addition to alkenes (second step) : 
CH 2 -CH 2 X + Y e > YCH 2 -CH 2 X 

B. NATURE OF THE SUBSTITUTING AGENT 

It is important to realize that in aromatic substitution the electrophilic sub- 

Se Se 
stituting agent, X® or X— Y, is not necessarily the reagent that is initially 
added to the reaction mixture. For example, nitration in mixtures of nitric 
and sulfuric acids is not usually brought about by attack of the nitric acid 
molecule on the aromatic compound, but by attack of a more electrophilic 
species, the nitronium ion, N0 2 ®. There is good evidence to show that this ion 
is formed from nitric acid and sulfuric acid according to the following equa- 
tion: 

HN0 3 +2H 2 S0 4 „ N0 2 ® + H 3 O a + 2 HS0 4 e 

The nitronium ion so formed then attacks the aromatic ring to give an 
aromatic nitro compound. 

NO, 

-H® 



NOf 



nitrobenzene 

In general, the function of a catalyst (which is so often necessary to promote 
aromatic substitution) is to generate an electrophilic substituting agent from 
the given reagents. 

C. NITRATION 

We have already mentioned that the nitronium ion, N0 2 ffi , is the active 
nitrating agent in nitric acid-sulfuric acid mixtures. The nitration of toluene 
is a fairly typical example of a nitration that proceeds well using nitric acid 
in a 1 : 2 mixture with sulfuric acid. The nitration product is a mixture of o-, 
m-, and ^-nitrotoluenes. 





chap 20 arenes. electrophilic aromatic substitution 562 

CH 3 CH 3 CH 3 CH 3 

HN0 3 , H 2 S0 4 



62% 5% 

The presence of any appreciable concentration of water in the reaction 
mixture is deleterious since water tends to reverse the reaction by which 
nitronium ion is formed. 

HNO3 + H 2 S0 4 . N0 2 ® + HS0 4 e + H 2 

It follows that the potency of the mixed acids can be increased by using 
fuming nitric and fuming sulfuric acids, which have almost negligible water 
contents. With such mixtures, nitration of relatively unreactive compounds 
can be achieved. For example,/?-nitrotoluene is far less reactive than toluene but 
when heated with an excess of nitric acid in fuming sulfuric acid (H 2 S0 4 + 
S0 3 ), it can be converted successively to 2,4-dinitrotoluene and to 2,4,6- 
trinitrotoluene (TNT). 

CH 3 CH 3 CH 3 

-NOi 0,N JL NO, 




HNO 3> 50° fi T HNO 3 ,120° 




so 3 ,h 2 so 4 K^y so 3 ,h 2 so4 \^ 

I 

N0 2 NO a 

^-nitrotoluene 2,4-dinitrotoluene 2,4,6-trinitrotoluene 

There are several interesting features about the nitration reactions thus far 
discussed. In the first place, the conditions required for nitration of j?-nitro- 
toluene would, in contrast, rapidly oxidize an alkene by cleavage of the double 
bond. 

/ CP V 
HNO, CH 2 "C0 2 H 

" I 

CH 2 C0 2 H 

CH 2 

adipic acid 

We may also note that the nitration of toluene does not lead to equal 
amounts of the three possible mononitro toluenes. The methyl substituent 
apparently orients the entering substituent preferentially to the ortho and para 
positions. This aspect of aromatic substitution will be discussed later in the 
chapter in conjunction with the effect of substituents on the reactivity of 
aromatic compounds. 

D. HALOGENATION 

The mechanism of halogenation is complicated by the fact that molecular 
halogens, Cl 2 , Br 2 , and I 2 , form complexes with aromatic hydrocarbons. 



sec 20.4 reactions of aromatic hydrocarbons 563 

Although complex formation assists substitution by bringing the reactants 
in close proximity, it does not always follow that a substitution reaction will 
occur. A catalyst is usually necessary. The catalysts most frequently used are 
metal halides that can act as Lewis acids (FeBr 3 , A1C1 3 , and ZnCl 2 ). Their 
catalytic activity may be attributed to their ability to polarize the halogen- 
halogen bond: 

S® Se 

Br-Br-FeBr 3 

The positive end of the halogen dipole attacks the aromatic compound 
while the negative end is complexed with the catalyst. We may then represent 
the reaction sequence as in Figure 20-6, with the slow step being formation of 
a a bond between Br® and a carbon of the aromatic ring. 

The order of reactivity of the halogens is F 2 > Cl 2 > Br 2 > I 2 • Fluorine is 
too reactive to be of practical use for the preparation of aromatic fluorine 
compounds and indirect methods are necessary (see Chapter 21). Iodine is 
usually unreactive and, in fact, its reaction with some arenes is energetically 
unfavorable. Use of iodine monochloride instead of iodine usually improves 
both the rate and the equilibrium condition to the point where good yields of 
iodination products are obtained: C 6 H 6 + IC1 ->■ C 6 H 5 I + HCL Alterna- 
tively, molecular iodine can be converted to a more active species (perhaps 
I®) with an oxidizing agent such as nitric acid. With combinations of this 
kind, good yields of iodination products are obtained. 




I 

o-iodotoluene />-iodotoluene 



Figure 20-6 A mechanism for the bromination of benzene in the presence of 
ferric bromide catalyst. 



So 
Br Br-FeBr, 

I , i 

Br dffi Br 



/ Y-Br + FeBr, + HBr 




chap 20 arenes. electrophilic aromatic substitution 564 



E. ALKYLATION 

An important method of synthesizing alkylbenzenes utilizes an alkyl halide 
as the alkylating agent together with a metal halide catalyst, usually aluminum 
chloride. 

CH,CH, 



+ CH,CH,Br 



benzene ethyl bromide 
(large excess) 



AICI3 

80° 



ethylbenzene 
83% 



HBr 



The class of reaction is familiarly known as Friedel-Crafts alkylation. The 
metal-halide catalyst functions much as it does in halogenation reactions; 
that is, it provides a source (real or potential) of a positive substituting agent, 
which in this case is a carbonium ion. 



CH, 



CH, 



\ 
CH— CI + A1C1, 

/ - 



CH, 



CH 3 

■^i \e> e 

+ CH— C1-A1C1, 

/ 3 

CH 3 



CH — CI— A1C1, 

/ 3 



CH 




+ A1C1 3 + 



HC1 



cumene 
(isopropylbenzene) 

Alkylation is not restricted to alkyl halides ; any combination of reagents 
giving carbonium ions will serve. Frequently used combinations are alcohols 
and alkenes in the presence of acidic catalysts, such as H 3 P0 4 , H 2 S0 4 , HF, 
BF 3 , or HF - BF 3 . Ethylbenzene is made commercially from benzene and 
ethene using phosphoric acid as the catalyst. Cumene is made similarly from 
benzene and propene. 



CH 2 =CH 2 

h 3 po 4 



^C H 3 CH=CH 2 

H3P04 



CH 




CH, 




sec 20.4 reactions of aromatic hydrocarbons 565 

Under these conditions, the carbonium ion, which is the active substituting 
agent, is generated by protonation of the alkene. 

CH 2 =CH 2 +H® . CH 3 CH 2 ® 

CH 3 CH=CH 2 +H® . CH3CHCH3 

It is not possible to make K-propylbenzene satisfactorily by direct alkylation 
of benzene because the rc-propyl cation rearranges to the isopropyl cation as 
quickly as it is formed. Thus, cumene is the product of the reaction of benzene 
with either «-propyl chloride or isopropyl chloride. 



+ or > 1 + HC1 

%^ (CH 3 ),CHC1 



A serious drawback to Friedel-Crafts alkylation is the tendency for poly- 
substitution to occur. This is because the alkyl group enhances further substi- 
tution (Section 20-5). The use of a large excess of arene is helpful in favoring 
monosubstitution. 



F. ACYLATION 

Acylation and alkylation of arenes are closely related. Friedel-Crafts acylation 
introduces an acyl group, RCO— , into an aromatic ring, and the product is an 
aryl ketone. Acylating reagents commonly used are acid halides, RCOC1, 
or anhydrides, (RCO) 2 0. The catalyst is usually aluminum chloride, and its 
function is to generate the active substituting agent, which potentially is an 
acyl cation. 

CH3COCI + AICI3 , CH 3 CO®-CI-AICl 3 

H COCH3 

+ CH3CO— CI-AICI3 ► I ® j] + A1CU 




COCH3 




+ HC1 + AICI3 



methyl phenyl ketone 
(acetophenone) 

Acylation differs from alkylation in that the reaction is usually carried out 
in a solvent, commonly carbon disulfide or nitrobenzene. Furthermore, acyla- 



chap 20 arenes. electrophilic aromatic substitution 566 

tion requires more catalyst than alkylation because much of the catalyst is 
effectively removed by complex formation with the product ketone. 

C 6 H 5 

C 6 H 5 COCH 3 + AICI3 ■ C=0-A1C1 3 

CH 3 

1 : 1 complex 

When an acylating reagent such as carboxylic anhydride is used, still more 
catalyst is required because some is consumed in converting the acyl com- 
pound to the acyl cation. 

(RCO) 2 + 2 A1CU > RCO A1CU + RC0 2 A1C1 2 

Unlike alkylation, acylation is easily controlled to give monosubstitution 
because, once an acyl group is attached to a benzene ring, it is not possible to 
introduce a second acyl group into the same ring. For this reason, arenes are 
sometimes best prepared by acylation, followed by reduction of the carbonyl 
group with amalgamated zinc and hydrochloric acid (Section 11-4F). For 
example, rc-propylbenzene is best prepared by this two-step route since, as we 
have noted, the direct alkylation of benzene with H-propyl chloride will give 
considerable amounts of cumene and poly substitution products. 




+ CH 3 CH,COCl 



COCH 2 CH 3 CH 2 CH 2 CH 3 

AICI3 (i^l Zn(Hg), HC1 





nitrobenzene 
propanoyl chloride /z-propylbenzene 



G. SULFONATION 



Substitution of the sulfonic acid (— S0 3 H) group for a hydrogen of an aro- 
matic hydrocarbon is usually carried out by heating the hydrocarbon with a 
slight excess of concentrated or fuming sulfuric acid. 



SO,H 




H 2 S0 4 , S0 3 



30°-50° 




+ H 2 



benzenesulfonic acid 




H 2 S0 4 



.no°-i20° 




+ H 2 



p-toluenesulfonic acid 

The actual sulfonating agent is normally the S0 3 molecule, which, although 
it is a neutral reagent, has a powerfully electrophilic sulfur atom. 



sec 20.5 effect of substituents on reactivity and orientation 567 



O 

II 

/A 

o o 



o 

c© 

o' V 



+ so, 



Q e 

o x o 




o 

J. 

°o o 



S0 3 H 




H. DEUTERATION 



It is possible to replace the ring hydrogens of many aromatic compounds by 
exchange with deuteriosulfuric acid. The mechanism is analogous to other 
electrophilic substitutions. 




H + D,SO, 




=o + 



HDS0 4 



Perdeuteriobenzene can be made from benzene in good yield if a suf- 
ficiently large excess of deuteriosulfuric acid is used. Sulfonation, which might 
appear to be a competing reaction, requires considerably more vigorous 
conditions. 



20-5 effect of substituents on reactivity and 
orientation in electrophilic aromatic 
substitution 

In planning syntheses based on substitution reactions of mono- or polysubsti- 
tuted benzenes, you must be able to predict in advance which of the available 
positions of the ring are most likely to be substituted. This is now possible 
with a rather high degree of certainty, thanks to the work of many chemists 
over the last 100 years. Few, if any, other problems in organic chemistry have 
received so much attention, and there is now accumulated enough data on the 
orienting and reactivity effects of ring substituents in electrophilic substitution 
to permit the formulation of some valuable generalizations. 

Basically, three problems are involved in the substitution reactions of 
aromatic compounds : (a) proof of the structures of the possible isomers, 
o, m, and p, that are formed ; (b) the percentage of each isomer formed, if the 
product is a mixture ; and (c) the reactivity of the compound being substituted 
relative to some standard substance, usually benzene. 

Originally, the identity of each isomer formed was established by Korner's 
absolute method, which involves determining how many isomers each will give 
on further substitution, this number being diagnostic of the particular isomer 
(see Exercise 20-1). In practice, Korner's method is often very tedious and 
lengthy, and it is now primarily of historical interest except in its application 
to substitution reactions of unusual types of aromatic systems. For benzenoid 



chap 20 arenes. electrophilic aromatic substitution 568 

compounds, structures can usually be established with the aid of correlations 
between spectroscopic properties and positions of substitution, as we have 
indicated earlier in this chapter. Also, it is often possible to convert the 
isomers to compounds of known structure by reactions that do not lead to 
rearrangement. For example, trifluoromethylbenzene on nitration gives only 
one product, which has been shown to be the meta-nitro derivative by con- 
version to the known m-nitrobenzoic acid. 




hno 3 




CO,H 



H 2 S0 4 



NO, 




NO, 



A. THE PATTERN OF ORIENTATION IN AROMATIC SUBSTITUTION 

The reaction most studied in connection with the orientation problem is ni- 
tration, but the principles established also apply for the most part to the 
related reactions of halogenation, sulfonation, alkylation, and acylation. Some 
illustrative data for the nitration of a number of monosubstituted benzene 
derivatives are given in Table 20-6. The orientation data are here expressed as 
the percentage of ortho, meta, and para isomers formed, and the rate data are 



Table 20-6 Orientation and rate data for nitration of some monosubstituted 
benzene derivatives 





R 

rV 




R 

1 NO 

u 


R 

■•6 


+ 1 

V N0 2 


R 

> 


















N0 2 










ortho 


meta 




para 


substituent, R 


orientation 


relative 

reactivity 


partial rate factors 


%o 


%m 


%P 


/. 


fn. 


fp 


-CH 3 


56.5 


3.5 


40 


24 


42 


2.5 


58 


-C(CH 3 ) 3 


12.0 


8.5 


79.5 


15.7 


5.5 


4.0 


75 


-CH 2 C1 


32.0 


15.5 


52.5 


0.302 


0.29 


0.14 


0.951 


—CI 


29.6 


0.9 


68.9 


0.033 


0.029 


0.0009 


0.137 


-Br 


36.5 


1.2 


62.4 


0.030 


0.033 


0.0011 


0.112 


-NO, 


6.4 


93.2 


0.3 


~10" 7 


1.8 x 10 " 6 


2.8 x 10 


" 5 2xl0" 7 


-C0 2 C 2 H 5 


28.3 


68.4 


3.3 


0.0003 


2.5 X 10~ 4 


6x 10" 


" 4 5xl0" 5 


-CF 3 




100 




low 








-N(CH 3 ) 3 




89 


11 


low 









sec 20.5 effect of substituents on reactivity and orientation 569 

overall rates relative to benzene. Rates are also expressed as partial rate 
factors, symbolized asf ,f m , and/ p , which are, respectively, the rate of sub- 
stitution at one of the ortho, meta, and para positions relative to one of the 
six equivalent positions in benzene. Consideration of the partial rate factors 
is particularly useful, since it lets you tell at a glance if, for example, a 
substituent gives ortho, para substitution with activation (f ,f p > 1), but meta 
substitution with deactivation (f m < 1). 

Inspection of the data in Table 20-6 shows that each substituent falls into 
one of three categories : 

1. Those substituents [e.g., CH 3 and — C(CH 3 ) 3 ] which activate all the ring 
positions relative to benzene (/> 1), but are more activating for the ortho 
and para positions than for the meta position. These substituents lead to pre- 
dominance of the ortho and para isomers. As a class, they give ortho, para 
orientation with activation. Other examples, in addition to those included in 
Table 20-6, are -OH, -OCH 3 , -NR 2 , and -NHCOCH 3 . 

2. Those substituents (e.g., CI, Br, and CH 2 C1) which deactivate all of the 
ring positions (/< 1) but deactivate the ortho and para positions less than 
the meta position so that formation of the ortho and para isomers is favored. 
These substituents are classified as giving ortho, para orientation with 
deactivation. e 

3. Those substituents [e.g., -N0 2 , -C0 2 C 2 H 5 , -N(CH 3 ) 3 , and — CF 3 ] 
that deactivate all the ring positions (/< 1) but deactivate the ortho and para 
positions more than the meta position. Hence, mostly the meta isomer is 
formed. These substituents give meta orientation with deactivation. 

There is no known example of a substituent that activates the ring and, at 
the same time, directs an electrophilic reagent preferentially to the meta 
position. 

A more comprehensive list of substituents which fall into one of the three 
main categories is given in Table 20-7. It may be convenient to refer to this 
table when in doubt as to the orientation characteristics of particular substi- 
tuents. An explanation of these substituent effects follows, which should make 
clear the criteria whereby predictions of the behavior of other substituents 
can be made with considerable confidence. First, however, we must examine 
more closely the energetics of electrophilic substitution. 

The distribution of products in most aromatic electrophilic substitutions of 
benzene derivatives is determined not by the relative stabilities of the products 
but by the rates at which they are formed. Thus nitration of chlorobenzene 
gives mostly o- and ^-nitrochlorobenzene, whereas chlorination of nitro- 
benzene produces mostly the meta isomer. This means that benzene can be 
converted to one or the other set of products depending on the sequence of 
the nitration and chlorination reactions (Figure 20-7). Furthermore, there is 
virtually no isomerization of the products. 

To rationalize the orientation effects of ring substituents, then, we should 
compare the transition states leading to the various products. The rate-con- 
trolling step in electrophilic aromatic substitution is normally the first step — 
the attack of the electrophile at the activated position — not the second step 
in which a proton is lost from the intermediate ion. The energy profile for the 
attack of an electrophile Z® on benzene is shown in Figure 20-8. 



chap 20 arenes. electrophilic aromatic substitution 570 



Table 20-7 Orientation and reactivity effects of ring 
substituents 



o,p orientation 


o,p orientation 


m orientation 


■with activation 


with deactivation 


with deactivation 


-OH 


— CH 2 C1 


-N0 2 


_ e 


— F 


-NH 3 


-OR 


-CI 


e 
-NR 3 


-OC 6 H 5 


-Br 


© 
-PR3 


-NH 2 


—I 


© 
-SR 2 


-NR 2 


— CH=CHN0 2 


-IC 6 H 5 


-NHCOCH3 




-CF 3 


-alkyl(e.g.,CH 3 ) 




-CCU 


-aryl (e.g., C 6 H 5 ) 




-SO3H 

-S0 2 R 

-C0 2 H 

-C0 2 R 

-CONH 2 

-CHO 

-COR 

-C=N 



The transition states are closer in energy to the intermediate ion than they 
they are to either the products or the reactants. It is convenient to take the 
intermediate as equivalent to the transition states in the discussion that 
follows. 



B. ELECTRICAL EFFECTS 

An important effect of substituent groups on aromatic substitution is the in- 
ductive effect which we have encountered previously in connection with the 
ionization of carboxylic acids (Section 13-4B). An electron-attracting group 
( — /effect) will exert an electrostatic effect such as to destabilize a positively 

charged intermediate, while an electron-donating group ( + /effect) will have the 

e 
opposite effect. We shall illustrate this simple principle, using the (CH 3 ) 3 N— 

group as an example. This group is strongly electron-attracting. If we write 
the hybrid structure of the substitution intermediate with the group X repre- 
senting some electrophilic substituting agent, we see at once that the charge 

e 
produced in the ring is unfavorable when the (CH 3 ) 3 N— substituent is 



sec 20.5 effect of substituents on reactivity and orientation 571 



/ 



CI, 



\hno 3 



/FeCl 3 \H 2 S0 4 
CI NO, 



HN0 3 , \ 

H 2 S0 4 FeCl 3 \ci 2 




CI CI 

NO, 




CI 




NO, 



ortho 



meta 



Figure 20-7 The sequence of chlorination and nitration reactions required 
to give the three isomers of nitrochlorobenzene. 



Figure 20-8 Energetics of the reaction of an electrophile Z® with benzene 
showing the formation of the intermediate, C 6 H 6 Z ffi , and its decomposition to 
products. The transition state which determines the rate of the overall re- 
action is that of the first step. 




chap 20 arenes. electrophilic aromatic substitution 572 



present, particularly for substitution at the ortho and para positions where 
adjacent atoms would carry like charges. Thus, although all three intermediates 
should then be less stable than the corresponding intermediate for benzene, 
the ortho and para intermediates should be less favorable than the one for 
meta substitution. This should lead to meta orientation with deactivation, as 
indeed is observed. (We show the charge distribution in the intermediate ions 
in these examples.) 



N(CH 3 ) 3 
' H 



N(CH 3 ) 3 



N(CH 3 ) 3 






w X H' "X 

ortho substitution meta substitution para substitution 

Other substituents that are strongly electron attracting and that also orient 

e 
meta with deactivation include — N0 2 , — CF 3 , — P(CH 3 ) 3 , — S0 3 H, 
-C0 2 H, -C0 2 CH 3 , — CONH 2 , — CHO, -COC 6 H 5 , and — C=N. 

The activating and ortho, /?ara-orienting influence of alkyl substituents can 
also be rationalized on the basis of inductive effects. Thus, the substitution 
intermediates for ortho, para, and meta substitution of toluene are stabilized 
by the capacity of a methyl group to release electrons ( + / effect) and partially 
compensate for the positive charge. 

CH, 






H X 

ortho substitution meta substitution para substitution 

Furthermore, the stabilization is most effective in ortho and para substitu- 
tion where part of the positive charge is adjacent to the methyl substituent. 
(For other examples of stabilization of positive carbon by alkyl groups see 
Section 4-4C). The result is ortho,para orientation with activation. 

In addition to the inductive effects of substituents, conjugation effects may 
be a factor in orientation and are frequently decisive. This is especially true of 
substituents that carry one or more pairs of unshared electron pairs on the 
atom immediately attached to the ring (e.g., — OH, ~0: e , — OCH 3 , 
— NH 2 , — NHCOCH 3 , —CIO- An electron pair so situated helps to stabilize 
the positive charge of the substitution intermediate, as the extra resonance 
forms [5] and [6] will indicate for ortho and para substitution of anisole 
(methyl phenyl ether). 



ortho substitution 







sec 20. S effect of substituents on reactivity and orientation 573 

OCH 3 ®OCH 3 

para substitution 

H X 

[6] 

In meta substitution, however, the charge is not similarly stabilized as no 
resonance structures analogous to [5] or [6] can be written. Accordingly, the 
favored orientation is ortho,para, but whether substitution proceeds with 
activation or deactivation depends on the magnitude of the inductive effect 
of the substituent. For example, halogen substituents are strongly electro- 
negative and deactivate the ring at all positions ; yet they strongly orient ortho 
mdpara through conjugation of the unshared electron pairs. Apparently the 
inductive effect is strong enough to reduce the overall reactivity, but not 
powerful enough to determine the orientation. Thus for para substitution of 
chlorobenzene the intermediate stage is formed less readily than in the 
substitution of benzene itself. 

®C1 





(unfavorable, because positive 
carbon is next to chlorine) 

Other groups such as — NH 2 , — NHCOCH 3 , and — OCH 3 are electron 
attracting but much less so than the halogens, and the inductive effect is 
completely overshadowed by the conjugation effect. Therefore, substitution 
proceeds with ortho,para orientation and activation. The most activating 
common substituent is — 0' e , which combines a large electron-donating in- 
ductive effect with a conjugation effect. 



C. ORIENTATION IN DISUBSTITUTED BENZENES 

The orientation and reactivity effects of substituents discussed for the substi- 
tution of monosubstituted benzenes also hold for disubstituted benzenes ex- 
cept that the directing influence now comes from two groups. Qualitatively, 
the effects of the two substituents are additive. We would therefore expect 
/7-nitrotoluene to be less reactive than toluene because of the deactivating 
effect of a nitro group. Also, the most likely position of substitution should be, 
and is, ortho to the methyl group and meta to the nitro group. 




N0 2 e 

_____ 




NO, 



NO, 



chap 20 arenes. electrophilic aromatic substitution 574 

When the two substituents have opposed orientation effects, it is not always 
easy to predict what products will be obtained. For example, 2-methoxy- 
acetanilide has two powerful ortho, para-directing substituents, — OCH 3 and 
— NHCOCH3 . Nitration of this compound gives mainly the 4-nitro derivative, 
which indicates that the — NHCOCH 3 exerts a stronger influence than OCH 3 . 

NHCOCH3 NHCOCH3 

i. ,OCH 3 /L XDCHj 



u 



-H® 




20-6 substitution reactions of polynuclear 
aromatic hydrocarbons 

Although naphthalene, phenanthrene, and anthracene resemble benzene in 
many respects, they are more reactive than benzene in both substitution and 
addition reactions. This is expected theoretically because quantum mechanical 
calculations show that the loss in stabilization energy for the first step in 
electrophilic substitution or addition decreases progressively from benzene to 
anthracene; therefore the reactivity in substitution and addition reactions 
should increase from benzene to anthracene. 

In considering the properties of the polynuclear hydrocarbons relative to 
benzene, it is important to recognize that the carbon-carbon bonds in poly- 
nuclear hydrocarbons are not all alike nor do they correspond exactly to 
benzene bonds. This we may predict from the hybrid structures of these 
molecules, derived by considering all of the electron-pairing schemes having 
normal bonds, there being three such structures for naphthalene, four for 
anthracene, and five for phenanthrene. See Figure 20-9. If we assume that 
each structure contributes equally to its resonance hybrid, then, in the case of 
naphthalene, the 1,2 and 2,3 bonds have § and -J double-bond character, re- 
spectively. Accordingly, the 1,2 bond should be shorter than the 2,3 bond, 
and this has been verified by X-ray diffraction studies of crystalline naphtha- 
lene. 




bond lengths of naphthalene, A units 

Similarly, the 1,2 bond of anthracene should have f double-bond character 
and should be shorter than the 2,3 bond, which has only \ double-bond 
character. The 1,2 bond is indeed shorter than the 2,3 bond. 




bond lengths of anthracene, A units 



sec 20.6 substitution reactions of polynuclear aromatic hydrocarbons 575 



Naphthalene 





Anthracene 







Phenanthrene 








Figure 20-9 Resonance structures for naphthalene, anthracene, and phenanthrene. 

The trend toward greater inequality of the carbon-carbon bonds in poly- 
nuclear hydrocarbons is very pronounced in phenanthrene. Here, the 9,10 
bond is predicted to have f double-bond character, and experiment verifies 
that this bond does resemble an alkene double bond, as we shall see in sub- 
sequent discussions. 



A. NAPHTHALENE 

In connection with orientation in the substitution of naphthalene, the picture 
is often complex, although the 1 position is the more reactive (Figure 20-10). 
Sometimes, relatively small changes in the reagents and conditions change 
the pattern of orientation. One example is sulfonation, a reversible reaction 
leading to 1-naphthalenesulfonic acid at 120° but to the 2 isomer on pro- 
longed reaction or at temperatures above 160°C. Another example is supplied 
by Friedel-Crafts acylation: the major product in carbon disulfide is the 
1 isomer, while in nitrobenzene it is the 2 isomer. 



chap 20 arenes. electrophilic aromatic substitution - 576 



HN0 3 H 2 S0 4 





1-nitronaphthalene 
Br 



Br 2 



50% CH 3 C0 2 H,H 2 





1-bromonaphthalene 2-bromonaphthalene 
99% !•/ 



COCH, 



CH3COCI-AICI3 

CS 2 , -15° 



CH3COCI 
C 6 H 5 NQ 2 , AICI3 

25° 





COCH3 



1 -aeetonaphthalene 2-acetonaphthalene 

7 5% 25% 



COCH, 




Figure 20-10 Electrophilic substitution pattern of naphthalene. 



Normally, substitution of naphthalene occurs more readily at the 1 position 
than at the 2 position. This may be accounted for on the basis that the most 
favorable resonance structures for either the 1- or the 2-substituted inter- 
mediate are those which have one ring fully aromatic. We see then that 1 
substitution is favored over 2 substitution since the positive charge in the 
1 intermediate can be distributed over two positions, leaving one aromatic 
ring unchanged; but this is not possible for the 2 intermediate without 
affecting the benzenoid structure of both rings. 



/ substitution: 
H X 




rL .X 




2 substitution: 
e H 





sec 20.6 substitution reactions of polynuclear aromatic hydrocarbons 577 

B. PHENANTHRENE AND ANTHRACENE 

The substitution patterns of the higher hydrocarbons are more complex than 
for naphthalene. For example, phenanthrene can be nitrated and sulfo- 
nated, but the products are mixtures of 1-, 2-, 3-, 4-, and 9-substituted 
phenanthrenes. 





sulfonation 



nitration 



(percentages are yields of sulfonic (figures at the 9, 1, 2, 3, and 4 

acids with H 2 S0 4 at 60°. At 120°, positions are partial rate 

mostly the 2- and 3-sulfonic acids factors) 
are obtained) 

The 9,10 bond in phenanthrene is quite reactive; in fact, almost as much so 
as an alkene double bond. Addition therefore occurs readily, giving both 
9,10 addition and 9-substitution products (Scheme I). 




Br 2 



phenanthrene 




SCHEME I 



Anthracene is even more reactive than phenanthrene and has a great tendency 
to add various reagents to the 9,10 positions. The addition products 
of nitration and halogenation readily give the 9-substitution products on 
warming. 

Br 




Br 2 




-HBr 




chap 20 arenes. electrophilic aromatic substitution 578 

20-7 nonbenzenoid conjugated cyclic compounds 

A. AZULENE 

There are a number of compounds that possess some measure of aromatic 
character typical of benzene, but that do not possess a benzenoid ring. 
Appropriately, they are classified as nonbenzenoid aromatic compounds. One 
example of interest is azulene, and, like benzene, it tends to react by substitu- 
tion, not addition. It is isomeric with naphthalene and has a five- and a seven- 
membered ring fused through adjacent carbons. As the name implies, it is 




azulene 

deep blue in color. It is less stable than naphthalene and isomerizes quanti- 
tatively on heating above 350° in the absence of air. 



>350° 





Azulene can be represented as a hybrid of neutral and ionic structures. 

e 





The polarization shown above puts weight on those ionic structures having 
six electrons in both the five- and seven-membered rings (see Section 6-7). 



B. CYCLOOCTATETRAENE 

Of equal interest to azulene is cyclooctatetraene, which is a bright yellow, 
nonbenzenoid, nonaromatic compound with alternating single and double 
bonds. If the carbons of cyclooctatetraene were to occupy the corners of a 
regular planar octagon, the C— C— C bond angles would have to be 135°. 
Cyclooctatetraene does not conform to the Hiickel 4n + 2 rule (Section 6-7) 
and it is not surprising that the resonance energy gained in the planar struc- 
ture is not sufficient to overcome the unfavorable angle strain. Cycloocta- 
tetraene, instead, exists in a "tub" structure with alternating single and 
double bonds. 



tub 




♦ summary 579 

There is, however, nmr evidence that indicates that the tub form is in quite 
rapid equilibrium with a very small amount of the planar form at room tem- 
perature. Probably there is not much more than a 10-kcal energy difference 
between the two forms. 



summary 

The rules of nomenclature for arenes (aromatic hydrocarbons) are similar to 
those for aliphatic systems except that pairs of substituents at different ring 
positions may be designated ortho, meta, and para. A number of important 
arenes are C 6 H 6 (benzene), C 6 H 5 CH 3 (toluene), C 6 H 5 CH(CH 3 ) 2 (cumene), 
and C 6 H 5 CH=CH 2 (styrene). A number of important aryl or aralkyl groups 

/ 1 

are C 6 H 5 - (phenyl), C 6 H 5 CH 2 - (benzyl), C 6 H 5 CH (benzal), C 6 H 5 C— 

\ I 

(benzo), and (C 6 H 5 ) 2 CH— (benzhydryl). 

Polynuclear aromatic hydrocarbons have aromatic rings fused together so 
that the rings have one or more sides in common. Important examples of these 
are naphthalene, anthracene, and phenanthrene. Fusing an additional ring 
to an arene normally adds C 4 H 2 to the molecular formula; the extra ring is 
designated benzo or benz. 






naphthalene (C 10 H 8 ) anthracene (G 14 H 10 ) phenanthrene (C 14 H 10 ) 

Aromatic rings have characteristic absorption bands in the infrared (near 
1500, 1600, and just above 3000 cm" 1 ), in the ultraviolet (near 2000 A but 
shifting to higher wavelengths with conjugation), and in nmr spectra (6.5 to 
8.0 ppm for aromatic protons). 

Electrophilic substitution serves to introduce the following groups into an 

O 

II 
aromatic ring: — N0 2 , -CI (or Br or I), -S0 3 H, -R (alkyl), R-C— 

(acyl), and — D. The actual electrophilic reagents in these cases are electron- 
deficient cations (except for S0 3 ). Substituents already present in the ring 
determine the position taken by the incoming group. Some important sub- 
stituents that direct the incoming electrophile to the ortho and para positions 
are — R (alkyl), —OR, — NH 2 , and X (halogen). Those that direct toward 

O 

© II 

the meta position include — N0 2 , — NH 3 , — C— , and — C=N. 

The orienting effect of a substituent Y on an electrophile Z ffi is determined 

by how Y affects the dispersal of the positive charge in the transition state, 

a reasonable model for which is the intermediate ion [1] (shown for the case 



chap 20 arenes. electrophilic aromatic substitution 5S0 



of para substitution). 





fast „ 

> H e + 




All of the common ortho,para-directing substituents, with the exception of 
halogen, activate the ring toward substitution and all the meta-directing sub- 
stituents deactivate the ring; nitro to such an extent that the Friedel-Crafts 
reaction cannot be applied to nitrobenzene. 

Naphthalene is normally substituted at the 1 position (a) but sulfonation 
at high temperatures and Friedel-Crafts acylation in nitrobenzene give sub- 
stitution at the 2 position (/?). Phenanthrene gives mixtures of substitution 
products but anthracene tends to react by addition at the 9,10 positions. 
Other cyclic hydrocarbons that have been considered are phenalenyl [2], a 
resonance-stabilized radical; azulene [3], which has aromatic character; and 
cyclooctatetraene [4], which does not. There are a number of other resonance 
structures for [2] and [3]. 




[2] [3] [4] 



exercises 

20-1 How many isomeric products could each of the xylenes give on introduction 
of a third substituent? Name each isomer using chlorine as the third sub- 
stituent. 

20-2 Name each of the following compounds by an accepted system: 

a. (C 6 H 5 ) 2 CHC1 

b. C 6 H 5 CHC1 2 

c. C 6 H 5 CC1 3 

CH, 

d. 





exercises 581 

20-3 How many possible disubstitution (X,X and X,Y) products are there for 
naphthalene, phenanthrene, anthracene, and biphenyl? Name each of the 
possible dimethyl derivatives. 

20-4 A number of polynuclear hydrocarbons are shown below with their con- 
jugated rings shown as hexagons with inscribed circles. 




d. 



/• 




20-5 



(i) Determine which of these compounds are radicals by drawing Kekule 
structures, (ii) Three of the structures above represent compounds whose 
structures have been given earlier in the chapter in a somewhat different 
form. Identify them. 

Identify the two compounds with molecular formula C 7 H 7 C1 from their 
infrared spectra in Figure 20-11. 



20-6 Predict the effect on the ultraviolet spectrum of a solution of aniline in water 
when hydrochloric acid is added. Explain why a solution of sodium phen- 
oxide absorbs at longer wavelengths than a solution of phenol (see Table 
20-3). 



20 "7 Establish the structures of the following benzene derivatives on the basis of 
their empirical formulas andnmr spectra as shown in Figure 20-12. Remem- 
ber that equivalent protons do not normally split each other's resonances 
(Section 7-6B). 



a. 
b. 



CsHto 
C 8 H 7 OCl 



c. C 9 H 10 2 

d. C9H12 



20-8 Calculate from appropriate bond and stabilization energies the heats of 
reaction of chlorine with benzene to give (a) chlorobenzene and (b) 5,6- 
dich!oro-l,3-cyclohexadiene. Your answer should indicate that substitution 
is energetically more favorable than addition. 



chap 20 arenes. electrophilic aromatic substitution S82 



^V 



4 5 

i r 



~\ -r 



rAT-f^ 



wavelength, jU 



9 10 12 14 

t — i i — ~r 



\H" f \^ /Anf\ ni 



T\ 



mm 



CH-Cl 






ai: 



i i i i i _j i i_ 



3600 3200 2800 2400 2000 2000 1800 



1 600 _. 1 400 

frequency, cm 



i I I Jt 

w 

w 



I 






J I [_ 



I 1_ 



1 000 800 



~v 



4 5 

m 



r' v *A/~~ — : 



iC,H,Cl 



wavelength, jjl 



i i i ' i i ■ i . w' . 

3600 3200 2800 2400 2000 2000 1800 1600 

frequency, cm 




1000 800" 



Figure 20-11 Infrared spectra of two isomeric compounds of formula C 7 H 7 C1 
(see Exercise 20- 5). 



20-9 On what basis (other than the thermodynamic one suggested in Exercise 
20-8) could we decide whether or not the following addition-elimination 
mechanism for bromination of benzene actually takes place ? 




+ Br 2 



Br 

kJr Br 

H 




Br 



+ HBr 




g 

o 



g 



o 



u 
s- 

D 

X 

u 



E 

o 

s 

H 


u 
c 
o 





£ 


3° 


N 


Ice 


K 


1 


s 




o 




v* 




-H 


J 


(S 




C/l 


sg 


4) 


i u. 




' o. 


-H 




rs 


iO 


> 








u 




o 




-o 




u 




a 




D 




N 




S 




<U 


!<_) 


J3 



! O 



1° 



a, 



o 
to 



o 



S 

c 

s 
o 



u 

ft. 



. Ju: 



___; 



o 

CO 



3 



chap 20 arenes. electrophilic aromatic substitution 584 

20-10 Explain with the aid of an energy diagram for aromatic nitration how one 
can account for the fact that hexadeuteriobenzene undergoes nitration with 
nitric acid at the same rate as ordinary benzene. 

20-1 1 From the fact that nitrations in concentrated nitric acid are strongly retarded 
by added nitrate ions and strongly accelerated by small amounts of sulfuric 
acid, deduce the nature of the actual nitrating agent. 

20-12 Account for the fact that fairly reactive arenes (e.g., benzene, toluene, and 
ethylbenzene) are nitrated with excess nitric acid in nitromethane solution at 
a rate that is independent of the concentration of the arene (i.e., zeroth 
order). Does this mean that nitration of an equimolal mixture of benzene 
and toluene would necessarily give an equimolal mixture of nitrobenzene 
and nitrotoluenes ? Why or why not ? 

20-13 Write a mechanism for the alkylation of benzene with isopropyl alcohol 
catalyzed by boron trifluoride. 

20-14 Suggest possible routes for the synthesis of the following compounds: 
a - (~y-CU 2 -(~\ c. CH,-/ \-CH 2 CH 3 

o o 

11 f~~\ II 

b. CH 3 --C-f VC-CH 3 

20-15 Calculate the partial rate factors for each different position in the mono- 
nitration of biphenyl, given that the overall reaction rate relative to benzene 
is 40, and the products are 68% o-, 1% m-, and 31% /j-nitrobiphenyl. 
(Remember, there are two benzene rings in biphenyl.) 

20-16 Explain why the — CF 3 , — N0 2 , and — CHO groups should be meta 
orienting with deactivation. 

20-17 Explain why the nitration and halogenation of biphenyl goes with activation 
at the ortho and para positions but with deactivation at the meta position. 
Suggest a reason why biphenyl is more reactive than 2,2'-dimethylbiphenyl 
in nitration. 

20-18 Explain why the bromination of aniline gives 2,4,6-tribromoaniline, whereas 
the nitration of aniline with mixed acids gives /n-nitroaniline. 

20-19 Predict the favored positions of substitution in the nitration of the following 
compounds: 




rv< 



-CH 2 N(CH 3 ) 3 



CI 



Br 
/ % 



CH 




Br 









exercises S8S 




F 

/ 




/■ 


6 


-OCH 3 




8 

'111 




fl 


ff- 


5 




4 



Br (consider the character of 

. / the various resonance 



( V 



Br 



structures for substitution 



\ / in the 1- and 2- positions) 

20-20 Predict the orientation in the following reactions: 

a. 1-methylnaphthalene + Br 2 

b. 2-methylnaphthalene + HNO3 

c. 2-naphthoic acid + HNO3 

20-21 How would you go about proving that the acylation of naphthalene in the 
2 position in nitrobenzene solution is not the result of thermodynamic 
control? 

20-22 Show how you can predict qualitatively the character of the 1,2 bond in 
acenaphthylene. 




20-23 Write structural formulas for all of the possible isomers of C 8 Hi containing 
one benzene ring. Show how many different mononitration products each 
could give if no carbon skeleton rearrangements occur but nitration is con- 
sidered possible either in the ring or side chain. Name all of the mononi- 
tration products by an accepted system. 

20-24 Write structural formulas (more than one might be possible) for aromatic 
substances that fit the following descriptions : 

a. CsHio, which can give only one theoretically possible ring nitration 
product 

b. C 6 H 3 Br 3 , which can give three theoretically possible nitration prod- 
ucts 

c. C 6 H 3 Br 2 Cl, which can give two theoretically possible nitration prod- 
ucts 

d. C s Hs(N0 2 )2 , which can give only two theoretically possible different 
ring monobromo substitution products 

20-25 1 ,3-Cyclohexadiene cannot be isolated from reduction of benzene by hydro- 
gen over nickel. The isolable reduction product is always cyclohexane. 



chap 20 arenes. electrophilic aromatic substitution 586 

Explain why the hydrogenation of benzene is difficult to stop at the 
1,3-cyclohexadiene stage, even though 1,3-butadiene is relatively easy 
to reduce to butenes. 

How could an apparatus for determining heats of hydrogenation be 
used to obtain an accurate AH value for the reaction? 



+ H, 




Calculate a AH of combustion for benzene as 1,3,5-cyclohexatriene 
(no resonance) from bond energies and compare it with a calculated 
value for heat of combustion of benzene obtained from the experi- 
mental AH, +5.9 kcal, for the hydrogenation in (b). 



20-26 Predict the most favorable position for mononitration for each of the follow- 
ing substances. Indicate whether the rate is greater or less than for the nitra- 
tion of benzene. Give your reasoning in each case. 



a. fiuorobenzene 

b. trifiuoromethylbenzene 

c. acetophenone 

d. nitrosobenzene 

e. benzyldimethylamine oxide 
/. diphenylmethane 

g. p-methoxybromobenzene 



h. 



J- 

k. 



diphenyl sulfone 
/?-?-butyltoluene 

(C 6 H 5 ) 2 IN0 3 
m-diphenylbenzene 
(m-terphenyl) 
4-acetylaminobiphenyl 



20-27 Predict which of the following compounds have some aromatic character. 
Give your reasons. 

Q OH 




© Bre 
tropylium bromide 








aceplieadylene 



n> : i '*?•■•;;•*,{*•! 



IK 




mmmm&mmmm 




chap 21 aryl halogen compounds, nucleophilic aromatic substitution S89 

The chemical behavior of aromatic halogen compounds depends largely on 
whether the halogen is attached to carbon of the aromatic ring, as in bromo- 
benzene, C 6 H 5 Br, or to carbon of an alkyl substituent, as in benzyl bromide, 
C 6 H 5 CH 2 Br. Compounds of the former type are referred to as aryl halides, 
and those of the latter as arylalkyl halides. 

Aryl halides are expected to resemble vinyl halides to some extent since 
both have their halogen atoms attached to unsaturated carbon. 



bromobenzene vinyl bromide 

(phenyl bromide) (bromoethene) 

Consequently, it is no surprise to find that most aryl halides are usually 
much less reactive than alkyl or allyl halides toward nucleophilic reagents in 
either S N 1 or S N 2 reactions. Whereas ethyl bromide reacts easily with sodium 
methoxide in methanol to form methyl ethyl ether, vinyl bromide and bromo- 
benzene completely fail to undergo nucleophilic displacement under similar 
conditions. Also, neither bromobenzene nor vinyl bromide reacts appreciably 
with boiling alcoholic silver nitrate solution even after many hours. 

In contrast to phenyl halides, benzyl halides are quite reactive. In fact, they 
are analogous in reactivity to allyl halides (Section 9-6). 



CH,=CH-CH,-Br 



benzyl bromide allyl bromide 

Benzyl halides are readily attacked by nucleophilic reagents in both S N 1 
and S N 2 displacement reactions. The ability to undergo S N 1 reactions is 
clearly related to the stability of the benzyl cation, the positive charge of 
which is expected, on the basis of the resonance structures [la] through [Id], 
to be extensively delocalized. 





[la] [lb] [lc] [Id] 

■* 

When the halogen substituent is located two or more carbons from the aro- 
matic rings — as in 2-phenylethyl bromide, C 6 H 5 CH 2 CH 2 Br — the pronounced 
activating effect evident in benzyl halides disappears, and the reactivity of the 
halide is essentially that of a primary alkyl halide (e.g., CH 3 CH 2 CH 2 Br). 
Since, in general, the chemistry of arylalkyl halides is related more closely to 
that of aliphatic derivatives than to aryl halides, we shall defer further dis- 
cussion of arylalkyl halides to Chapter 24, which is concerned with the 
chemistry of aromatic side-chain derivatives. 



chap 21 aryl halogen compounds, nucleophilic aromatic substitution 590 
Table 21-1 Physical properties of aryl halides 



name 


bp, 

°C 


mp, 
°C 


rf 20/4 


<' 


fluorobenzene 


85 


-42 


1.024 


1.4646 22 - 8 


chlorobenzene 


132 


-45 


1.1066 


1.5248 


bromobenzene 


155 


-31 


1.4991 15 ' 15 


1.5598 


iodobenzene 


189 


-31 


1.832 


1.6214 185 


o-chiorotoluene 


159 


-34 


1.0817 


1.5238 


m-chlorotoluene 


162 


-48 


1.0732 


1.5214 19 


p-chlorotoluene 


162 


8 


1.0697 


1.5199 19 


1 -chloronaphthalene 


263 




1.1938 


1.6332 20 


2-chloronaphthalene 


265 


55 







21 -1 physical properties of aryl halogen compounds 

There is nothing unexpected about most of the physical properties of aryl 
halides. They are slightly polar substances and accordingly have boiling 
points approximating those of hydrocarbons of the same molecular weights ; 
their solubility in water is very low, whereas their solubility in nonpolar 
organic solvents is high. In general, they are colorless, oily, highly refractive 
liquids with characteristic aromatic odors and with densities greater than that 
of water. A representative list of halides and their physical properties is given 
in Table 21-1. 

With respect to the infrared spectra of aryl halides, correlations between 
structure and absorption bands of aromatic carbon-halogen bonds have not 
proved to be useful. 



21-2 preparation of aryl halides 

Many of the methods that are commonly used for the preparation of alkyl 
halides simply do not work when applied to the preparation of aryl halides. 
Thus, it is not possible to convert phenol to chlorobenzene by reagents such as 
HC1— ZnCl 2 , SOCl 2 , and PC1 3 which convert ethanol to chloroethane. In 
fact, there is no very practical route at all for conversion of phenol to chloro- 
benzene. In this situation, it is not surprising that some of the methods by 
which aryl halides are prepared are not often applicable to the preparation of 
alkyl halides. One of these methods is direct halogenation of benzene or its 
derivatives with chlorine or bromine in the presence of a metal halide catalyst, 
as discussed in Section 20-4C. 

Br 

FeCl 3 



■ci, 




sec 21.2 preparation of aryl halides 591 

Direct halogenation of monosubstituted benzene derivatives often gives a 
mixture of products, which may or may not contain practical amounts of the 
desired isomer. A more useful method of introducing a halogen substituent 
into a particular position of an aromatic ring involves the reaction of an 
aromatic primary amine with nitrous acid under conditions that lead to the 
formation of an aryldiazonium salt. Decomposition of the diazonium salt 
to an aryl chloride or bromide is effected by warming a solution of the 
diazonium salt with cuprous chloride or bromide in an excess of the corre- 
sponding halogen acid. The method is known as the Sandmeyer reaction. 



CH 




CH, 



NH, 




o-toluidine 



o-methylbenzenediazonium 

chloride 

(not isolated) 



CH 



(CuCl) 
HC1, 60° 




o-chlorotoluene 

74-79°/ 



For the formation of aryl iodides from diazonium salts, the cuprous cata- 
lyst is not necessary since iodide ion is sufficient to cause decomposition of the 
diazonium salt. Both cuprous ion and iodide ion appear to be involved in 
an oxidation-reduction process at the diazo group that promotes the de- 
composition. 




NH 2 
1 -phenanthrylamine 



1. NaNQ 2 , H 2 SQ 4 , H 2 Q ; 
2. KI-H 2 * 




1 -iodophenanthrene 

53% 



Aryl fluorides may also be prepared from diazonium salts if the procedure 
is slightly modified. The amine is diazotized in the usual way; then fiuoboric 
acid or a fluoborate salt is added, which usually causes precipitation of a 
sparingly soluble diazonium fluoborate. The salt is collected and thoroughly 
dried, then carefully heated to the decomposition point, the products being 
an aryl fluoride, nitrogen, and boron trifluoride. 



© e 
C 6 H 5 N 2 BF 4 



heat 



-> C 6 H 5 F + N 2 +BF 3 



The reaction is known as the Schiemann reaction. An example (which gives a 
rather better than usual yield) follows : 



NH, 




1. NaNQ 2 ,H 2 SQ 4 
2. HBF 4 




4-bromo-l- 
naphthylamine 



4-bromonaphthalene 
-1-diazonium fluoborate 



l-fluoro-4- 
bromonaphthalene 

97% 



chap 21 aryl halogen compounds, nucleophilic aromatic substitution 592 




2. (CuCl) 



Figure 21-1 Preparation of m-dichlorobenzene from benzene. The nitration 
reaction gives very largely meta substitution (see Table 20-6) and the m-dinitro 
product is easily purified by crystallization. 



The arylamines necessary for the preparation of aryl halides by the Sand- 
meyer and Schiemann reactions are usually prepared by reduction of the cor- 
responding nitro compounds (see Chapter 22), which in turn are usually ob- 
tained by direct nitration of an aromatic compound. For example, although 
ra-dichlorobenzene cannot be prepared conveniently by direct chlorination 
of benzene, it can be made by dinitration of benzene followed by reduction 
and the Sandmeyer reaction (Figure 21-1). In connection with this synthesis, 
it should be noted that tetrazotization (double diazotization) of 1,2- and 1,4- 
diaminobenzene derivatives is not as easy to achieve as with the 1,3 compound, 
because the 1,2- and 1,4-diaminobenzenes are very easily oxidized. 



21 -3 reactions of aryl halides 



A. ORGANOMETALLIC COMPOUNDS FROM ARYL HALIDES 

Grignard reagents can be prepared with fair ease from aryl bromides or 
iodides and magnesium metal. 

ether 
C 6 H 5 Br + Mg ► C 6 H 5 MgBr 

phenylmagnesium bromide 

Chlorobenzene and other aryl chlorides are usually unreactive unless added 
to the magnesium admixed with a more reactive halide. 1,2-Dibromoethane 
is particularly useful as the second halide because it is converted to ethene, 
which does not then contaminate the products, and it continually produces a 
fresh magnesium surface, which is sufficiently active to be able to react with 
the aryl chloride. 

The reactions of arylmagnesium halides are analogous to those of alkyl- 
magnesium halides (see Chapter 9) and require little further comment. 

Aryllithiums can usually be prepared by direct reaction of lithium metal 
with chloro or bromo compounds. 



sec 21.3 reactions of aryl halides S93 



ether 

C 6 H 5 C1 + 2 Li ► C 6 H 5 Li + LiCl 

phenyllithium 



As with the Grignard reagents, aryllithiums react as you might expect by 
analogy with alkyllithiums. 



B. NUCLEOPHILIC DISPLACEMENT REACTIONS OF ACTIVATED 
ARYL HALIDES 

While the simple aryl halides are inert to the usual nucleophilic reagents, con- 
siderable activation is produced by strongly electron-attracting substituents, 
provided these are located in either the ortho or para positions, or both. As 
one example, the displacement of chloride ion from l-chloro-2,4-dinitroben- 
zene by dimethylamine occurs measurably fast in ethanol solution at room 
temperature. Under the same conditions, chlorobenzene completely fails to 
react ; thus, the activating influence of the two nitro groups easily amounts 
to a factor of at least 10 8 . 

H 3 C CH 3 

\ / 3 

e NH Cl e • 

N °2 J^ N0 2 

+ (CH 3 ) 2 NH C2 " 5 5 o° H > 





N0 2 

A related reaction is that of 2,4-dinitrofluorobenzene with peptides and 
proteins, which is used for analysis of the N-terminal amino acids in polypep- 
tide chains. (See Section 17-3A.) 

In general, the reactions of activated aryl halides bear a close resemblance 
to S N 2 displacement reactions of aliphatic halides. The same nucleophilic re- 
agents are effective (e.g., CH 3 O e , HO e , and RNH 2 ); the reactions are second 
order overall — first order in halide and first order in nucleophile. For a given 
halide, the more nucleophilic the attacking reagent, the faster is the reaction. 
There must be more than a subtle difference in mechanism, however, since an 
aryl halide is unable to pass through the same type of transition state as an 
alkyl halide in S N 2 displacements. 

A generally accepted mechanism of nucleophilic aromatic substitution vis- 
ualizes the reaction as proceeding in two steps closely analogous to those 
postulated for electrophilic substitution (Chapter 20). The first step involves 
attack of the nucleophile Y : e at the carbon bearing the halogen substituent 
to form an intermediate anion [2]. The aromatic system is of course* destroyed 
on forming the anion, and the hybridization of carbon at the reaction site 
changes from sp 2 to sp 3 . 

X 

I 




In the second step, loss of an anion, X° or Y e , regenerates an aromatic 



chap 21 aryl halogen compounds, nucleophilic aromatic substitution 594 

system, and, if X e is lost, the reaction is one of overall nucleophilic displace- 
ment of X for Y. 



In the case of a neutral nucleophilic reagent, Y or HY, the reaction se- 
quence would be the same except for the necessary adjustments in charge of 
the intermediate. 




Formation of [2] is highly unfavorable for the simple phenyl halides, even 
with the most powerful nucleophilic reagents. It should be clear how electron- 
attracting groups, — N0 2 , —NO, — C=N, — N 2 ®, and so on, can facilitate 
nucleophilic substitution by this mechanism through stabilization of the 
intermediate. The effect of such substituents can be illustrated in the case of 
/?-bromonitrobenzene and its reaction with methoxide ion. The structure of 
the reaction intermediate can be described in terms of the resonance structures 
[3a] through [3d]. Of these [3d] is especially important because the negative 
charge can be located on oxygen, an electronegative atom. 



Br ,OCH 3 


Br ,OCH 3 


V 


kJ 


N ffi 




[3a] 


[3b] 



Br OCH3 


Br OCH 

- 


N e 


V 



[3c] [3d] 

The reason that substituents in the meta positions have much less effect on 
the reactivity of an aryl halide is the substituent's inability to contribute di- 
rectly to the derealization of the negative charge in the ring ; no structures 
can be written analogous to [3d]. 



CHXt Br CH 3 C) Br CH,0 Br 



e 



e 




e ^ s 

N ^-^ N^ ^^^N 

II II e II 

o o o 



C. ELIMINATION-ADDITION MECHANISM OF NUCLEOPHILIC 
AROMATIC SUBSTITUTION 

The reactivity of aryl halides such as the halobenzenes and halotoluenes is 
exceedingly low toward nucleophilic reagents that normally effect smooth 



sec 21.3 reactions of aryl halides 595 

displacements with alkyl halides and activated aryl halides. Substitutions, 
however, do occur under sufficiently forcing conditions involving either high 
temperatures or very strong bases. For example, the reaction of chlorobenzene 
with sodium hydroxide solution at temperatures around 340° is an important 
commercial process for the production of phenol. 

OH 

aq. NaOH ./^ 

3W * I + NaCl 



Also, aryl chlorides, bromides, and iodides can be converted to arylamines 
by amide ions, which are very strong bases. In fact, the reaction of potassium 
amide with bromobenzene is extremely rapid, even at temperatures as low as 
— 33°, with liquid ammonia as solvent. 

NH 2 

© e nh, fi^l e e 

+ KNH, "o ' + K Br 



Displacement reactions of this type, however, differ from the previously 
discussed displacements of activated aryl halides in that rearrangement often 
occurs. That is to say, the entering group does not always take up the same 
position on the ring as that vacated by the halogen substituent. For example, 
the hydrolysis of /?-chlorotoluene at 340° gives an equimolar mixture of m- 
and p-cresols. 

CH, CH, 

aq. NaOH 
340° * 




Even more striking is the exclusive formation of m-aminoanisole in the 
amination of o-chloroanisole. 



OCH, OCH 



© 6 

NaNH 2 , NH 3 





NH, 



Mechanisms of this type have been widely studied, and much evidence has 
accumulated in support of a stepwise process, which proceeds first by base- 
catalyzed elimination of hydrogen halide (HX) from the aryl halide. This 
first reaction resembles the E2 elimination reactions of alkyl halides discussed 
earlier (Section 8-12) except that the abstraction of the proton appears to 
precede loss of the bromide ion. The reaction is illustrated below for the 
amination of bromobenzene. 



chap 21 aryl halogen compounds, nucleophilic aromatic substitution 596 




6 



e 
+ :NH, 



H^y 



NH 3 

-33° 



+■ NH, + Br fc 



benzyne 
[4] 
The product of the elimination reaction is a highly reactive intermediate [4] 
called benzyne, or dehydrobenzene. Its formula is C 6 H 4 and it differs from 
benzene in having an extra bond between two ortho carbons. Benzyne reacts 
rapidly with any available nucleophile, in this example the solvent ammonia, 
to give an addition product. 



<r\ 



| + : 



:NH, 




NH, 




NH, 



[4] 



aniline 



The occurrence of rearrangements in these reactions follows from the possi- 
bility of the nucleophile's attacking the intermediate at one or the other of the 
carbons of the extra bond. With benzyne itself, the symmetry of the molecule 
is such that no rearrangement would be detected. However, this symmetry is 
destroyed if one of the ring carbons is labeled with 14 C isotope, so that two 
isotopically different products can be formed. Studies of the amination of halo- 
benzenes labeled with 14 C at the 1 position have demonstrated that essentially 
equal amounts of 1- and 2- 14 C-labeled anilines are produced, as predicted 
by the elimination-addition mechanism. 




KNH 2 



NH,.-33° 



NH 3 




NH, 




NH, 



X = CI, Br, I 

* = 14 C 



50° 



50° 



21-4 organochlorine pesticides 



The general term pesticide includes insecticides, herbicides, and fungicides. A 
number of the most important pesticides are chlorinated aromatic hydro- 
carbons or their derivatives (Figure 21-2). Prodigious quantities have been 
used throughout the world in the past 25 years with, as we now know, tragic 
consequences. 

Chlorinated hydrocarbons such as DDT have low water solubility but high 
solubility in nonpolar media such as fatty tissue. Their slow rate of decompo- 
sition causes them to accumulate in nature, and predatory birds and animals 
are particularly vulnerable. The food chain running from plankton to small 
fish to bigger fish to predatory birds results in a magnification of the residue 
concentration at each stage. As a result the world's population of falcons, 
hawks, and eagles has dropped drastically in the past decade. A remarkable 



sec 21.4 organochlorine pesticides 597 



CI 

\ 




h 




CH -CC1, 




OH 

yx 


CI 


Cl 


l,l,l-trichloro-2,2- 

bis(p-chlorophenyl)ethane (DDT) 

an insecticide 


pentachlorophenol (PCP) 
a fungicide 




Cl 

)=\ 


CI-/ \-O-CH 2 C0 2 H 


Cl-/ V-0-CH 2 C0 2 H 


\ 
CI 


\ 
Cl 


2,4-dichlorophenoxyacetic acid 
a herbicide 


2,4,5-trichlorophenoxyacetic acid 
a herbicide 



Figure 21-2 Some organochlorine pesticides. The abbreviation DDT arises 
from the semisystematic name rfichlorodiphenyl/richloroethane. Bis, tris, and 
tetrakis are used in place of di, tri, and tetra for substituents whose names contain 
two parts; thus, the compound ^-CH 3 C 6 H 4 CH 2 C 6 H4.CH 3 -p can be named 
either di-p-tolylmethane or bis (p-methylphenyl)methane. 



effect of high pesticide residues in these birds is extreme fragility of the shells 
of their eggs. Experiments have shown that DDE [5], the principal decom- 
position product of DDT, causes this effect when present in very small amount. 
It is believed to inhibit the action of the enzyme carbonic anhydrase which 
controls the supply of calcium available for shell formation. 



V 


4 




C=CC1, 

-( 


i 


) 


/ 
Cl 




,1,1 -dichloro-2,2-bis(/>-chloropheny l)ethene 
[5] 



It has been estimated that there are now a billion pounds of DDE spread 
throughout the world ecosystem and traces of it have been found in animals 
everywhere, including the Arctic and Antarctic. Even though the use of DDT 



chap 21 aryl halogen compounds, nucleophilic aromatic substitution 598 

has now been severely curtailed by legislation it will take many years for the 
level of DDE to decrease to tolerable levels. Man, like predatory birds, is at 
the top of a food chain and human beings now carry in their fatty tissue 10 
to 20 ppm of chlorinated hydrocarbon insecticides and their conversion 
products. The effect on human health is still not known with certainty. 

The herbicides 2,4-D [6] and 2,4,5-T[7] have come under fire recently because 
their indiscriminate use as defoliants threatens the ecology of large areas. 
Furthermore, 2,4,5-T is suspected of being a teratogen (a fetus-deforming 
agent). Mice given 2,4,5-T in the early stages of pregnancy have a high inci- 
dence of fetal mortality and there is a high incidence of abnormalities in the 
survivors. There is some indication that 2,3,7,8-tetrachlorodibenzodioxin [8], 
sometimes present as an impurity in commercial samples of 2,4,5-T, may 
actually be the teratogenic agent. 



OCH 2 C0 2 H 




OCH 2 C0 2 H 




^Y cl 




rV 


c 'rYYY' 


Y 


cr 


V 


ci-^^o-^^ci 


Cl 




Cl 


2,3,7,8-tetrachlorodibenzodioxin 


[6] 




[7] 


[8] 



summary 

Aryl halides such as bromobenzene, C 6 H 5 Br, are unreactive to most nucleo- 
philes unless activating groups are present in the ring. Benzyl halides such as 
C 6 H 5 CH 2 Br, on the other hand, react readily by nucleophilic displacement. 
Aryl halides can be prepared from benzene by direct halogenation or from 
aniline by the Sandmeyer reaction. 




Grignard reagents can be prepared from aryl halides and their reactions 
are analogous to those of alkylmagnesium compounds. 

A halogen which is ortho or para to one or more nitro groups is activated 
toward nucleophilic substitution. 




exercises 599 



+ N§ ► |ej > IB + X" 

1^ I 

N0 2 N0 2 

Those aryl halides that lack activating groups may suffer displacement of 
halogen under forcing conditions via an elimination-addition reaction. 

NH 2 e NH 3 
C 6 H 5 Br -* C 6 H 4 Br e — — — » C 6 H 4 ► C 6 H 5 NH 2 

( — H©) ( — Bra) 

The amino group does not necessarily occupy the position vacated by the 
bromine atom since the intermediate, benzyne (C 6 H 4 ), is symmetrical and 
ammonia can add to it in either direction. 

Many chlorinated aromatic hydrocarbons or their derivatives are widely 
used as pesticides. 



exercises 

21-1 Suggest a feasible synthesis of each of the following compounds based on 
benzene as the starting material : 

Br 





a. F-C N )-NH 2 c. (' ^>-Br 

N0 2 

CH 3 

f\a d. cnA-f\i 



b. I-f VC1 d. CH 3 -C-<' y-Br 

CH 3 ^ p 

21-2 Suggest a method for preparing the following compounds from the indicated 
starting materials and any other necessary reagents: 

a. j p-ClC 6 H4C(CH 3 ) 2 OH from benzene 

b. 1 -naphthoic acid from naphthalene 




c. H0 2 C— ^ rf— D from toluene 

21-3 Why is the following mechanism of S N 2 substitution of an alkyl halide un- 
likely for aryl halides? 

e \ !e I 8e / n 

X:+ ,.C— Y , X-C— Y . X— C. + e Y : 

7 / \ V 

transition state 



chap 21 aryl halogen compounds, nucleophilic aromatic substitution 600 

214 a. Write resonance structures analogous to structures [3a] through [3d] to 
show the activating effect of — C=N, — S0 2 R, and — CF 3 groups in 
nucleophilic substitution of the corresponding ^-substituted chloro- 
benzenes. 
b. How would you expect the introduction of methyl groups ortho to the 
activating group to affect the reactivity of p-bromonitrobenzene and 
^-bromocyanobenzene toward ethoxide ion ? 

21-5 Would you expect p-bromonitrobenzene or (p-bromophenyl)-trimethyl- 
ammonium chloride to be more reactive in bimolecular replacement of 
bromine by ethoxide ion ? Why ? 

21-6 Would you expect p-chloroanisole to be more or less reactive than chloro- 
benzene toward methoxide ion? Explain. 

21-7 Devise a synthesis of each of the following compounds from the indicated 
starting materials : 




a> //—O — C 2 H 5 fromp-nitrochlorobenzene 



CC1 3 

\ 

b, [I from toluene 




c. 2 N— / \-N > from benzene 
NO, 



21-8 In the hydrolysis of chlorobenzene-l- 14 C with 4 M aqueous sodium hydrox- 
ide at 340°, the products are 58% phenol-l- 14 C and 42% phenol-2- 14 C. 
• Calculate the percentage of reaction proceeding (a) by an elimination- 
addition mechanism, and (b) by direct nucleophilic displacement. (You may 
disregard the effect of isotopic substitution on the reaction rates.) Would 
you expect the amount of direct displacement to increase or decrease if the 
reaction were carried out (a) at 240°, and (b) in aqueous sodium acetate in 
place of aqueous sodium hydroxide ? Give the reasons on which you base 
your answers. 

21 -9 Explain the following observations : 

a, 2,6-Dimethylchlorobenzene does not react with potassium amide in 
liquid ammonia. 

b. Fluorobenzene, labeled with deuterium in the 2 and 6 positions, under- 
goes rapid exchange of deuterium for hydrogen in the presence of 
potassium amide in liquid ammonia, but does not form aniline. 



exercises 601 



21-10 Predict the principal product of the following reaction: 

^CH,CH 2 NHCH, 2 moles 

C 6 H 5 Li, ether 




21-11 Give for each of the following pairs of compounds a chemical test, preferably 
a test tube reaction, that will distinguish between the two compounds. Write 
a structural formula for each compound and equations for the reactions 
involved. 

a. chlorobenzene and benzyl chloride 

b. /7-nitrochlorobenzene and m-nitrochlorobenzene 

c. /?-chloroacetophenone and a-chloroacetophenone 

d. p-ethylbenzenesulfonyl chloride and ethyl ^-chlorobenzenesulfonate 

e. p-bromoaniline hydrochloride and p-chloroaniline hydrobromide 

21-12 Show by means of equations how each of the following substances might be 
synthesized starting from the indicated materials. Specify reagents and 
approximate reaction conditions. Several steps may be required. 

a. 1,3,5-tribromobenzene from benzene 

b. p-fluorobenzoic acid from toluene 

c. m-bromoaniline from benzene 

d. i?-nitrobenzoic acid from toluene 

e. m-dibromobenzene from benzene 
/. m-nitroacetophenone from benzene 

g. 2,4,6-trinitrobenzoic acid from toluene 
h. benzyl m-nitrobenzoate from toluene 

21-13 Write a structural formula for a compound that fits the following descrip- 
tion: 

a. an aromatic halogen compound that reacts with sodium iodide in 
acetone but not with aqueous silver nitrate solution 

b. an aryl bromide that cannot undergo substitution by the elimination- 
addition (benzyne) mechanism 

c. the least reactive of the monobromomononitronaphthalenes toward 
ethoxide ion in ethanol 



21-14 Explain why the substitution reactions of a-halonaphthalenes in Equations 
21-1 through 21-3 show no significant variation in the percentage of a- and 
/S-naphthyl derivatives produced either with the nature of the halogen 
substituent or with the nucleophilic reagent. 




N(C 2 



Li N(C;H 5 ) 2) 
ether 





N(C 2 H 5 ) 2 



(21-1) 



38' 



62°/ 



chap 21 aryl halogen compounds, nucleophilic aromatic substitution 602 




(21-2) 



(21-3) 



33% 



67' 



21-15 The conversion of DDT to DDE (Section 21-4) is catalyzed by the enzyme 
DDT dehydrochlorinase, which, as you might expect, is a protein. What 
groups normally present as substituents on protein chains (Table 17-1) might 
aid the simultaneous removal of a proton and a chloride ion? 



■'•.ft*. ST. *!•>; 



'Wfi%^0%?iM CS^^ 



&*££* ' "Jt ■■ 










i-'JsfiS 



is*5i 



btf-i.'*^JJ5? ,: 



chap 22 aryl nitrogen compounds 605 

Many of the properties of aryl halides, such as their lack of reactivity in 
nucleophilic substitution reactions, are closely related to the properties of 
vinyl halides. Attempts to make similar comparisons between vinyl oxygen 
and nitrogen compounds and the related aryl oxygen and nitrogen compounds 
are often thwarted by the unavailability of suitable vinyl analogs. Thus, 
while vinyl ethers are easily accessible, most vinyl alcohols and primary or 
secondary amines are unstable with respect to their tautomers with C=0 
and C=N bonds. (The enol forms of 1,3-dicarbonyl compounds are notable 
exceptions; see Sections 12-6 and 16TD.) 

\ / 

N-H H N 

\ / I II 

C=C ► — C — C— Atf=-16kcal 

/ \ I 

O-H H O 

\ / I II 

C = C ► — C — C— A/f=-18kcal 

/ \ I 

That the same situation does not hold for most aromatic amino and hy- 
droxy compounds is a consequence of the stability of the benzene ring. This 
stability would be almost completely lost by tautomerization. For aniline, the 
stabilization energy based on its heat of combustion is 41 kcal/mole, and we 
can expect a stabilization energy (S.E.) of about 5 kcal/mole for its tautomer, 
2,4-cyclohexadienimine. Thus the AH of tautomerization is unfavorable by 
(41 - 16 - 5) = 20 kcal/mole. 





Alcaic.) = 20 kcal 



S.E. = 41 kcal S.E. ~ 5 kcal 



Phenol is similarly more stable than the corresponding ketone by about 17 
kcal/mole. 



OH 




Alcaic.) = 17 kcal 



S.E. = 40 kcal S.E. ~ 5 kcal 



Since aromatic amino and hydroxy compounds have special stabilization, 
their behavior is not expected to parallel in all respects that of the less stable 
vinylamines and vinyl alcohols. Nonetheless, similar reactions are often 
encountered. Both enols and phenols are acidic; they react readily with halo- 
gens, and their anions undergo either C or O alkylation with organic halides. 
The qualitative differences observed in these reactions will be considered in 
more detail later in this chapter (see also Chapter 23) ; but, as already indicated, 
such differences can usually be accounted for in terms of the stabilization of 
the aromatic ring. 



chap 22 aryl nitrogen compounds 606 

aromatic nitro compounds 

22-1 synthesis of nitro compounds 

The most generally useful way to introduce a nitro group into an aromatic 
nucleus is by direct nitration, as previously discussed (Section 20-4B). This 
method is obviously unsatisfactory when the orientation determined by sub- 
stituent groups does not lead to the desired isomer. Thus, /j-dinitrobenzene 
and p-nitrobenzoic acid cannot be prepared by direct nitration, since nitra- 
tions of nitrobenzene and benzoic acid give practically exclusively m-dini- 
trobenzene and m-nitrobenzoic acid, respectively. To prepare the para 
isomers, less direct routes are necessary. The usual stratagem is to use benzene 
derivatives with substituent groups that produce the desired orientation on 
nitration and then to make the necessary modifications in these groups to 
produce the final product. Thus, /7-dinitrobenzene can be prepared from aniline 
by nitration of acetanilide (acetylaminobenzene), followed by hydrolysis to 
p-nitroaniline and replacement of amino by nitro through the action of nitrite 
ion, in the presence of cuprous salts, on the corresponding diazonium salt 
(see Section 22-8). Alternatively, the amino group of /?-nitroaniline can be 
oxidized to a nitro group by trifluoroperacetic acid. In this synthesis, acetani- 
lide is nitrated in preference to aniline itself, since not only is aniline easily 
oxidized by nitric acid, but the reaction leads to extensive meta substitution 
by nitration involving the anilinium ion. Another route to /7-nitroaniline is to 
nitrate chlorobenzene and subsequently replace the chlorine with ammonia. 
See Figure 22 T for representation of these reactions. 

The nitrations mentioned give mixtures of ortho and para isomers, but 
these are usually easy to separate by distillation or crystallization. The same 
approach can be used to synthesize /?-nitrobenzoic acid. The methyl group of 
toluene directs nitration preferentially to the para position, and subsequent 
oxidation with chromic acid yields /?-nitrobenzoic acid. 



CO,H 



HN0 3 f \ Na 2 Cr 2 Q, 

H 2 SO* ' 



In some cases, it may be necessary to have an activating group to facilitate 
substitution, which would otherwise be very difficult. The preparation of 
1,3,5-trinitrobenzene provides a good example — direct substitution of m- 
dinitrobenzene requires long heating with nitric acid in fuming sulfuric acid. 
However, toluene is more readily converted to the trinitro derivative and this 
substance, on oxidation (Section 24-1) and decarboxylation (Section 13-6), 
yields 1,3,5-trinitrobenzene. 



sec 22.1 synthesis of nitro compounds 607 



CH, 



HN0 3 

> 

H 2 SO* 



0,N 



CH, 
"^^ N ° 2 Na 2 Cr 2 7 




CO a H 
0,N^ ^L NO, 



H 2 SO» 




NO, 



heat , 



N0 2 

65°/ 



<-co 2 



v 



NO, 



NO, 



Acylamino groups are also useful activating groups and have the advantage 
that the amino groups obtained after hydrolysis of the acyl function can be 



Figure 22-1 Schemes for the preparation ofp-dinitrobenzene from aniline or 
chlorobenzene. 



NH 2 NHCOCH, NHCOCH, 

(CH 3 CO) 2 ff ^ HNO3 



CH 3 C0 2 Na Ks? 



H 2 S0 4> 0° 



aniline 



acetanilide 



N0 2 
p-nitroacetanilide 



NaOH, H 2 
100° 



NO 



e e 
N 2 BF 4 



NH, 



NaN0 2 , Cu ffi 



NO 

/>-dinitrobenzene 




CI 



HNO3 



CI 



NO, 



chlorobenzene p-nitrochlorobenzene 



chap 22 aryl nitrogen compounds 608 




NH 



/j-toluidine aceto-/)-toluidide 



c=o 

CH, 



NaOH 




Figure 22-2 Preparation of m-nitrotoluene starting with p-aminotoluene 
(p-toluidine). 

removed from an aromatic ring by reduction of the corresponding diazonium 
salt with hypophosphorous acid, preferably in the presence of copper ions. An 
example is the preparation of m-nitrotoluene from/?-aminotoluene (j?-toluidine) 
via 4-acetylaminotoluene (aceto-/?-toluidide) as shown in Figure 22-2. 

The acetylamino derivatives of the amines are usually used in the nitration 
step in preference to the amines themselves because, as mentioned in connec- 
tion with the formation of />-nitroaniline, they are less susceptible to oxidation 
by nitric acid and give the desired orientation. 

The physical properties and spectra of aliphatic and aromatic nitro com- 
pounds were touched on briefly (Section 16-5). Nitrobenzene itself is a pale- 
yellow liquid (bp 210°), which should be handled with care because like many 
nitro compounds it is toxic when inhaled or when absorbed through the skin. 

A nitro group usually has a rather strong influence on the properties and 
reactions of other substituents on an aromatic ring, particularly when it is 
in an ortho or para position. A strong activating influence in displacement 
reactions of aromatic halogens was discussed in the preceding chapter 
(Section 21-3B). We shall see later how nitro groups make aromatic amines 
weaker bases and phenols stronger acids. 



22-2 reduction of aromatic nitro compounds 

The most important synthetic reactions of nitro groups involve reduction, 
particularly to the amine level. In fact, aromatic amines are normally pre- 



sec 22.2 reduction of aromatic nitro compounds 609 

pared by nitration followed by reduction. They may also be prepared by 
halogenation followed by amination. But since amination of halides requires 
the use of either amide salts or ammonia and high temperatures, which often 
lead to rearrangements (Section 21 -3C), the nitration-reduction sequence is 
usually preferred. Direct amination of aromatic compounds is not generally 
feasible. 



A. REDUCTION OF NITRO COMPOUNDS TO AMINES 

The reduction of nitrobenzene to aniline requires six equivalents of reducing 
agent and appears to proceed through the following principal stages : 



^ N0 ^0" N= ° "^ (^N-OH 

nitrobenzene nitrosobenzene N-phenylhydroxylamine 

2[H]/ 
^-H 2 



NH 2 

aniline 

Despite the complexity of the reaction, reduction of aromatic nitro com- 
pounds to amines occurs smoothly in acid solution with a variety of reducing 
agents of which tin metal and hydrochloric acid or stannous chloride are 
often favored on a laboratory scale. Hydrogenation is also useful but is 
strongly exothermic and must be carried out with care. 



NO, 



1. Sn.HCl 

50°-100° 

' 2. NaOH 

Ni,25° 
30atm 




Ammonium (or sodium) sulfide has the interesting property of reducing 
one nitro group in a dinitro compound much faster than the other. It is 




not always easy to predict which of two nitro groups will be reduced more 
readily. 

In contrast to the reduction of 2,4-dinitroaniline, reduction of 2,4-dinitro- 
toluene leads to preferential reduction of the 4-nitro group. 



chap 22 aryl nitrogen compounds 610 

CH 3 CH 3 

N 2 NH 3 ,H 2 S 





NO, NH 2 



B. REDUCTION OF NITRO COMPOUNDS IN NEUTRAL AND 
ALKALINE SOLUTION 

In neutral or alkaline solution, the reducing power of some of the usual 
reducing agents toward nitrobenzene is less than in acid solution. A typical 
reagent is zinc, which gives aniline in the presence of excess acid, but pro- 
duces N-phenylhydroxylamine when buffered with ammonium chloride. 

" \-NO, + 3 Zn + 6 HC1 Hz ° > ( VNH, + 3 ZnCl, + 2 H,0 



// ^ H 2 / \ 

? N >— N0 2 + 2Zn + 4NH 4 C1 ► <f VNHOH + 2 Zn(NH 3 ) 2 Cl 2 + H 2 



Nitrosobenzene is too easily reduced to be prepared by direct reduction of 
nitrobenzene and is usually made by oxidation of N-phenylhydroxylamine 
with chromic acid. 



// \ H K 2 Cr 2 0, // \ 

f Vn-oh ► ( Vn=o 

H 2 SO 4 ,0° \ / 



Nitrosobenzene exists as a colorless dimer in the crystalline state. When the 
solid is melted or dissolved in organic solvents, the dimer undergoes rever- 
sible dissociation to the green monomer (see Section 16-4). 

Reduction of nitrobenzene with methanol in the presence of sodium hydrox- 
ide produces azoxybenzene. The methanol is oxidized to formaldehyde. 



// ^> NaOH // ^ I f/ \ 

2f VNO, + 3 CH,OH ► V VN=N-f x > + 3 CH 2 + 3 H 2 

\ / 2 ' \ / © \— — / 

azoxybenzene 

The reason that azoxybenzene is produced instead of aniline is partly 
because methanol in alkali is a less powerful reducing agent than tin and 
hydrochloric acid. Also, in the presence of alkali, the intermediate reduction 
products can condense with one another; thus azoxybenzene probably arises 
in the reduction by a base-induced reaction of nitrosobenzene with N-phenyl- 



sec 22.3 polynitro compounds 61 1 

hydroxylamine. In fact, azoxybenzene can be prepared separately from these 
same reagents. This condensation reaction does not occur readily in acid 



OH 



f \_n=0 + H— N-f \ 



NaOH 



O w 

* f \-N=N-/ \ + H 2 



solution; furthermore, azoxybenzene is reduced to aniline by tin and hydro- 
chloric acid. 

Reduction of nitrobenzene in the presence of alkali with stronger reducing 
agents than methanol produces azobenzene and hydrazobenzene. Both of 
these compounds are reduction products of azoxybenzene and can be formed 
from azoxybenzene as well as from nitrobenzene by the same reducing re- 
agents (Scheme I). 



CH 3 OH, NaOH 



O 6 



azoxybenzene 



SnCl 2 



SnCl z 
NaOH 



NaOH 

C 6 H 5 -N=N-C 6 H 5 Zn 



azobenzene 



Zn 



NaOH 



NaOH 



Zn, NaOH 



Sn 



HC1 



- C.H.-NH, 



FT T-f 

* C 6 H 5 -N-N-C 6 H 5 

hydrazobenzene 
SCHEME I 

When hydrazobenzene is allowed to stand in strong acid solution, it 
undergoes an extraordinary rearrangement to form the technically important 
dye intermediate, benzidine (4,4'-diaminobiphenyl). 



<Q-£-£^3 -£- h 2 n^WVnh 2 



hydrazobenzene 



benzidine 



22-3 polynitro compounds 



A number of aromatic polynitro compounds have important uses as high 
explosives (Section 16-5). Of these 2,4,6-trinitrotoluene (TNT), 2,4,6-trinitro- 
phenol (picric acid), and N,2,4,6-tetranitro-N-methylaniline (tetryl) are 
particularly important. 1,3,5-Trinitrobenzene has excellent properties as an 
explosive but is difficult to prepare by direct nitration of benzene (Section 
22-1). 





chap 22 aryl nitrogen compounds 612 



NO, 



NO, NO, 



2,4,6-trinitrotoluene 2,4,6-trinitrophenol N,2,4,6-tetranitro- 
(TNT) (picric acid) N-methylaniline 

(tetryl) 

The trinitro derivatives of 3-/-butyltoluene and l,3-dimethyl-5-f-butyl- 
benzene possess musklike odors and have been used as ingredients of cheap 
perfumes and soaps. 



2 N 



C(CH 3 ) 3 (CH 3 ) 3 C 




NO, 




^* 



1 -raethyl-3-/-butyI- 1 ,3-dimethyl-5-f-butyl- 

2,4,6-trinitrobenzene 2,4,6-trinitrobenzene 



22-4 charge-transfer and tc complexes 

An important characteristic of polynitro compounds is their ability to form 
more or less stable complexes with aromatic hydrocarbons, especially those 
that are substituted with alkyl groups or are otherwise expected to have elec- 
tron-donating properties. The behavior is very commonly observed with 
picric acid, and the complexes therefrom are often nicely crystalline solids, 
which are useful for the separation, purification, and identification of aro- 
matic hydrocarbons. These substances are often called "hydrocarbon pic- 
rates" but the name is misleading since they are not ordinary salts; further- 
more, similar complexes are formed between aromatic hydrocarbons and 
trinitrobenzene, which shows that the strongly electron-attracting nitro 
groups rather than the acidic hydroxyl group are essential to complex for- 
mation. The binding in these complexes results from attractive forces between 
electron-rich and electron-poor substances. The designation charge-transfer 
complex originates from a resonance description in which the structure of the 
complex receives contributions from resonance forms involving transfer of an 
electron from the donor (electron-rich) molecule to the acceptor (electron- 
poor) molecule. However, the name n complex is also used because usually at 
least one component of the complex has a 7i-eIectron system. Charge- 
transfer complexes between polynitro compounds and aromatic hydro- 
carbons appear to have sandwich-type structures with the aromatic rings in 
parallel planes, although not necessarily coaxial. (See Figure 22-3.) 
Charge-transfer complexes are almost always more highly colored than 



sec 22.4 charge-transfer and tt complexes 613 

their individual components. A spectacular example is shown by benzene and 
tetracyanoethene, each of which separately is colorless, but which give a 
bright-orange complex when mixed. A shift toward longer wavelengths of 
absorption is to be expected for charge-transfer complexes relative to their 
components because of the enhanced possibility for resonance stabilization 
of the excited state involving both components. (See Sections 7-5 and 26-2.) 
With good electron donors such as carbanions, nitrobenzene itself will 
undergo charge transfer by adding an electron (Equation 22-1). The resulting 
radical anion [1] has both the negative charge and the spin of the odd electron 
distributed over the nitro group and the phenyl ring. 



R:+C 6 H 5 N0 2 < > R-+C 6 H 5 N0 2 - 

[1] 
A few of the many contributing structures for [1] are shown below. 



(22-1) 







etc. 



Many autoxidation reactions occur under basic conditions and are cata- 
lyzed by nitroarenes. The base generates a carbanion which is converted to a 
radical as in Equation 22T and this combines with molecular oxygen to form 
a peroxyradical, ROO-. The next step produces the hydroperoxide and re- 
generates more radical. Thus production of one radical can account for the 
consumption of many molecules of substrate. The initiation and propagation 
steps of this, a typical chain reaction, are shown below. 

RH + OH e < > R e +H 2 



R e +C 6 H 5 N0 2 < > R-+C 6 H 5 N0 2 ' 



R- +0 2 



R—0-0- +RH 



R-O-O- 
> ROOH + R- 



initiation 



propagation 



Figure 22-3 Formulation of charge-transfer complex between 1,3,5-trinitro- 
benzene (acceptor) and l,3>5-trimethylbenzene (donor). 



2 N 




CH 




CH 3 
5 n electrons 



chap 22 aryl nitrogen compounds 614 



aromatic amines 



22- 5 general properties 

Aniline, C 6 H 5 MH 2 , is a rather musty smelling liquid which is only slightly 
soluble in water. It solidifies at —6°, boils at 184°, and is colorless when pure. 
Like most aromatic amines, however, it tends to discolor on standing because 



Table 22-1 Physical properties of some representative aromatic amines 



name 


formula 




basicity," 

*B 

H 2 0, 25°C 


ultraviolet absor 


3tion 


Amax 


£ 


Amax 


e 


aniline 


CerisNIr^ 




4.6 x 10 - 10 


2300 


8600 


2800 


1430 


N-methylaniline 


C 6 H 5 NHCH 3 




2.5 x 10 - 106 


2450 


11,600 


2950 


1800 


N, N-dimethylaniline 


C 6 H 5 N(CH 3 ) 


. 


2.42 x 10 ~ 106 


2510 


14,000 


2980 


1900 


p-toluidine 


2 N 


-NH 2 


1.48 x 10 '" 


2320 


8900 


2860 


1600 


m-nitroaniline 


h\ 




4.0 x 10- 13 


2800 


4800 


3580 


1450 




V> NH2 












p-nitroaniline 


°>"-{~} 


-NH 2 


1.1 x lO" 12 


3810 


13,500 






/j-phenylenediamine 


H 2 NhQ 


-NH 2 








3210 


1550 


benzidine c 


H 2 N-^- 


\J~ 


-NH 2 1.4X10- 12 


2840 


24,500 






diphenylamine 


(C 6 H 5 ) 2 NH 




~10~ 14 


2850 


20,600 






triphenylamine 


(C 6 H 5 ) 3 N 

NH, 

1 




d 


2950 


23,000 






1-naphthylamine 


^^\^ 




9.9 x lO" 11 






3200 


5000 


2-naphthylamine c 


ca 


NH 2 


2.0 x 10 " 10 






3400 


2000 



Often given as K Ba ®, the dissociation constant of the conjugate acid, ArNH 3 ® + H 2 O^ArNH 2 + H 3 O ffl ; 
K B = W-^/Kbh® and ftuS) for aniline is 2.2 x 10~ 5 . The K H a values for aromatic amines, corresponding to the 
reaction ArNH 2 + H 2 O^ArNH e + H 3 O s , are low, but measurable, e.g., K HA for aniline is 10 -27 . 

6 At 18°C. 

c Hazardous substances; see Section 22-9B. 

d Not measurably basic in water solution. 



sec 22.5 general properties 615 

of air oxidation. Table 22-1 gives the basicities and the ultraviolet spectral 
properties of aniline and many of its derivatives. 

The chemical properties of the aromatic amines are in many ways similar to 
those of aliphatic amines. Alkylation and acylation, for example, occur in the 
normal manner (Sections 16-1E1 and 16-1F2). We have noted before (Section 
16- ID) that aniline is a weaker base than cyclohexylamine by a factor of 10 6 . 
The stabilization which can be ascribed to derealization of the unshared 
electron pair over the aromatic ring is lost in the cation because the electron 
pair must be localized when the nitrogen-proton bond is formed. The changes 
that occur in terms of the principal electron-pairing schemes for the aniline 
and anilinium ion are shown in Figure 22-4. 

A hybrid structure [4] for aniline, deduced from the structures [2a] 
through [2e], has some degree of double-bond character between the nitrogen 
and the ring, and some degree of negative charge at the ortho and para 
positions. 



Accordingly, the ability of the amine nitrogen to add a proton should be 
particularly sensitive to the electrical effects of substituent groups on the 
aromatic ring, when such are present. Many substituents such as nitro, cyano, 
and carbethoxy have the ability to stabilize an electron pair on an adjacent 



Figure 22-4 Electron-pairing schemes for aniline and anilinium ion. 





[2a] [2b] [2c] [2d] [2e] 





3 kcal extra stabilization attributed 
to derealization of unshared electron pair 




chap 22 aryl ntirogen compounds 616 

carbon (see Section 21 -3B). Such groups, located in the ortho or para posi- 
tion, should reduce substantially the base strength of the amine nitrogen. The 
reason is that the substituted aniline, but not the anilinium ion, is stabilized by 
contributions of electron-pairing schemes such as [5]. 




To gain some idea of the magnitude of this effect, we first note that aniline 
as a base is 90 times stronger than m-nitroaniline and 4000 times stronger than 
/>-nitroaniline. In contrast the acid-strengthening effect of a nitro group on 
benzoic acid is only 5.1 times for meta nitro and 6.0 fox para nitro. Clearly, 
the nitro groups in the nitroanilines exert a more powerful electrical effect 
than in the nitrobenzoic acids. This is reasonable because the site at which 
ionization occurs is closer to the benzene ring in the anilines than in the 
acids. Even when this factor is taken into account, however, /?-nitroaniline is 
much weaker than expected, unless forms such as [5] are important. 

The contribution made by the polar form [5] becomes even more important 
on excitation of p-nitroaniline by ultraviolet radiation (see Section 7-5). The 
necessary excitation energies are therefore lower than for aniline, with the 
result that the absorption bands in the electronic spectrum of ^-nitroaniline 
are shifted to much longer wavelengths and are of higher intensity than are 
those of aniline (cf. Table 22T). Since no counterpart to [5] can be written for 
m-nitroaniline, the absorption bands of w-nitroaniline are not as intense and 
occur at shorter wavelengths than those of />-nitroaniline. 

The — NH 2 group of aniline leads to very easy substitution by electrophilic 
agents (Section 20-5A) and high reactivity toward oxidizing agents. Bromine 
reacts rapidly with aniline in water solution to give 2,4,6-tribromoaniline in 
good yield. Introduction of the second and third bromines is so fast that it is 
difficult to obtain the monosubstitution products in aqueous solution. 

Br 
Br 2 ,H 2 / = \ 

Br 



f V >„, Br 2 , n 2 u / \ 



Other facets of the substitution of aromatic amines were discussed in con- 
nection with the orientation effects of substituents (see Section 20-5). 



22-6 aromatic amines with nitrous acid 

Primary aromatic amines react with nitrous acid at 0° in a way different from 
aliphatic amines in that the intermediate diazonium salts are much more 
stable and can, in most cases, be isolated as nicely crystalline fluoborate salts 



sec 22.7 preparation and general properties 617 

(Section 21-2). Other salts can often be isolated, but some of these, such as 
benzenediazonium chloride, are not very stable and may decompose with 
considerable violence. 



NH, 



NaN0 2 , HC1 

1 

0° 



s^n a ^^ 



N=N BR 



benzenediazonium 

chloride 

(water soluble) 



benzenediazonium 

fluoborate 
(water insoluble) 



The reason for the greater stability of aryldiazonium salts compared with 
alkyldiazonium salts seems to be related to the difficulty of achieving S N 1 
reactions with aryl compounds (Section 21 -3C). Even the gain in energy, 
associated with formation of nitrogen by decomposition of a diazonium ion, 
is not sufficient to make production of aryl cations occur readily at less than 
100°. 



N = N 



H 2 0, 100° 



H 2 



OH 



This reaction has considerable general utility for replacement of aromatic 
amino groups by hydroxyl groups. In contrast to the behavior of aliphatic 
amines, no rearrangements occur. 

Secondary aromatic amines react with nitrous acid to form N-nitroso 
compounds in the same way as do aliphatic amines (Section 16-1F3). 

Tertiary aromatic amines normally behave differently from aliphatic tertiary 
amines with nitrous acid in that they undergo C-nitrosation, preferably in the 
para position. It is possible that an N-nitroso compound is formed first, which 
subsequently isomerizes to the ^-nitroso derivative. 



CH, 



w r\ 



CH, 



HONO 



CH 



CH 3 

I 

N— N=0 



0=N 



CH 3 

/ 

\ 
CH, 



CH 3 

CH 3 

^-nitroso-N, N-dimethylaniline 



diazonium salts 



22-7 preparation and general properties 

The formation of diazonium salts from amines and nitrous acid has been 
described in the previous section. Most aromatic amines react readily, unless 
strong electron-withdrawing groups are present. 



chap 22 aryl nitrogen compounds 618 

Tetrazotization of aromatic diamines is usually straightforward if the 
amino groups are located on different rings, as with benzidine, or are meta to 
each other on the same ring. Tetrazotization of amino groups para to one 
another, or diazotization of /?-aminophenols, has to be conducted carefully to 
avoid oxidation to quinones (Section 23-3). 

Diazonium salts are normally stable only if the anion is one derived from a 
reasonably strong acid. Diazonium salts of weak acids usually convert to 
covalent forms from which the salts can usually be regenerated by strong acid. 
Benzenediazonium cyanide provides a good example in being unstable and 
forming two isomeric covalent benzenediazocyanides, one with the N=N 
bond trans and the other with the N=N bond cis. Of these, the trans isomer 
is the more stable. 



C 6 H 5 _ _ 

\ / \ 



© jt~- e \ / 

C 6 H 5 N==N: CN . N=N 

raj-benzenediazocyanide /ra/w-benzenediazocyanide 



N=N 

- X CN 



In strong acid, the covalent diazocyanides are unstable with respect to 
benzenediazonium ion and hydrogen cyanide. 

The covalent forms are sometimes significant in the reactions of diazonium 
salts, since they offer a convenient path for the formation of free radicals (see 
Section 16-6B). 

o 

C 6 H 5 N=N + 2 CCH 3 , C 6 H 5 N=N-0-C-CH 3 

O 

slow II 

► C 6 H 5 -+N 2 +-0-C-CH 3 



22-8 replacement reactions of diazonium salts 

The utility of diazonium salts in synthesis is largely due to the fact that they 
provide the only readily accessible substances that undergo nucleophilic 
substitution reactions on the aromatic ring under mild conditions without the 
necessity of having activating groups, such as nitro or cyano, in the ortho or 
para position. 



A. THE SANDMEYER REACTION 

The replacement of diazonium groups by halogen is the most important 
reaction of this type and some of its uses for the synthesis of aryl halides were 
discussed previously (Section 21-2). Two helpful variations on the Sandmeyer 
reaction employ sodium nitrite with cuprous ion as catalyst for the synthesis 
of nitro compounds (Section 22-1), and cuprous cyanide for the synthesis of 
cyano compounds. 



sec 22.9 reactions of diazonium compounds 619 




Cu 2 (CN) 2 
50° 



CH, 



B. THE SCHIEMANN REACTION 

The replacement of diazonium groups by fluorine was also covered earlier 
(Section 21-2). This reaction, like the replacement of the diazonium group by 
hydroxyl (Section 22-6), may well involve aromatic cations as intermediates. 
One strong piece of evidence for this is the fact that benzenediazonium 
fluoborate yields 3-nitrobiphenyl along with fluorobenzene when heated in 
nitrobenzene. Formation of 3-nitrobiphenyl is indicative of an electrophilic 
attack on nitrobenzene. 



bf 4 s r^ 



F + BF, 



© e -N 2 
N 2 BF 4 




22-9 reactions of diazonium compounds that occur 
without loss of nitrogen 



A. REDUCTION TO HYDRAZINES 



Reduction of diazonium salts to arylhydrazines can be carried out smoothly 
with sodium sulfite or stannous chloride, or by electrolysis. 






® e 1. Na 2 S0 3 ,H 2 / \ 

N=N CI - ► { />-NH-NH 2 

2. H^H.O, 100° ^ // 2 

3. NaOH 



B. DIAZO COUPLING 



A very important group of reactions of diazonium ions involves aromatic 
substitution by the diazonium salt acting as an electrophilic agent to yield azo 
compounds. 



chap 22 aryl nitrogen compounds 620 




This reaction is highly sensitive to the nature of the substituent (X), and 
coupling to benzene derivatives normally occurs only when X is a strongly 
activating group such as — O e , — N(CH 3 ) 2 , and — OH; however, coupling 
with X = OCH 3 may take place with particularly active diazonium com- 
pounds. Diazo coupling has considerable technical value, because the azo 
compounds that are produced are colored and often useful as dyes and color- 
ing matters. A typical example of diazo coupling is afforded by formation of 
/7-dimethylaminoazo benzene from benzenediazonium chloride and N,N- 
dimethylaniline. 



© e 
N 2 C1 



Q- n x 



CH, 



-HC1 



CH, 



*\ / 



N=N 



CH, 



N 



CH, 



/>-dimethylaminoazobenzene 
(yellow) 

The product was once used to color edible fats and was therefore known as 
"Butter Yellow" but its use in foods and cosmetics has been banned by 
many countries because of its ability to cause cancer in rats. There are indi- 
cations, but no firm evidence, that it causes cancer in humans. 

Certain other nitrogen-containing compounds, particularly aromatic 
amines, have been definitely shown to be carcinogenic for man. One of the 
most dangerous of these is 2-aminonaphthalene, formerly used as an anti- 
oxidant to protect the insulation on electric cables. In Britain a study showed 
that men continually exposed to this amine during its manufacture had a 
bladder cancer incidence of 50 % at 30 years from first exposure. One parti- 
cularly unfortunate group of 15 men involved in its distillation showed a 
100% incidence. Other dangerous amines are 4-aminobiphenyl and benzidine. 



NH, 



2-aminonaphthalene 
(2-naphthylamine) 




-NH, 



-NH, 



4-aminobiphenyl 



benzidine 
(4,4'-diaminobiphenyl) 



summary 

Aryl nitrogen compounds include amines, such as aniline, C 6 H 5 NH 2 ; nitro 
compounds, such as nitrobenzene, C 6 H 5 N0 2 ; and a number of substances 



exercises 621 

with nitrogen-nitrogen bonds, benzenediazonium salts, C 6 H 5 NfX e ; 
G e 
I 
azoxybenzene, C 6 H 5 N=NC 6 H 5 ; azobenzene, C 6 H 5 N=NC 6 H 5 ; and 

hydrazobenzene, C 6 H 5 NHNHC 6 H 5 . 

Aromatic nitro compounds are prepared by direct nitration using 
HN0 3 -H 2 S0 4 mixtures. Polynitration is difficult because of the deactivating 
effect of the nitro group, and use is often made of acylamino or alkyl sub- 
stituents to counteract this effect. The groups are removed at a later stage. 

Reduction of aromatic nitro compounds in acid solution (route A) gives 
amines directly, but in neutral or basic solution (route B), a number of com- 
pounds with nitrogen-nitrogen bonds can be isolated as intermediates. 

C 6 H 5 N0 2 ^ ► QH.NH, 



I 



(B) 



QH 5 NHNHC 6 H 5 



Nitroarenes abstract electrons from certain electron donors. Some good 
donors such as carbanions convert the nitrobenzene to a radical anion and 
this reaction can be important in initiating radical processes. 

R e +C 6 H 5 N0 2 < > R-+C 6 H 5 N0 2 - 

Poly nitroarenes form complexes with many neutral donors such as 
alkylbenzenes or polynuclear hydrocarbons by charge transfer. Such com- 
plexes are usually highly colored and can be described as two radical ions held 
together by electrostatic attraction 

Aromatic amines resemble aliphatic amines in most of their reactions but 
are considerably weaker bases because of resonance interaction between the 
amino group and the ring. Aromatic amines also differ in that they give fairly 
stable diazonium ions on treatment with nitrous acid ; these undergo a number 
of useful synthetic reactions. 

ArX 

ArCN 
ArNH 2 ► ArNj" 




ArN = NAr 
Certain aromatic amines, such as 2-aminonaphthalene, are carcinogenic. 

exercises 

22-1 Show how the following compounds could be synthesized from the indicated 
starting materials. (It may be necessary to review parts of Chapters 20 and 
21 to work this exercise.) 



chap 22 aryl nitrogen compounds 622 




from toluene 



NO 



b. 2 N\/^ ;; ^-N02 from /7-toluenesulfonic acid 



/V NH2 

N0 2 



0,N^^1 




2 N,.^L^N0 2 



from chlorobenzene 



from chlorobenzene 



from p-chlorobenzenesulfonic acid 



22-2 Tetracyanoethene in benzene forms an orange solution, but when this solu- 
tion is mixed with a solution of anthracene in benzene, a brilliant blue-green 
color is produced, which fades rapidly; colorless crystals of a compound of 
composition Ci 4 Hio - C 2 (CN)4 are then deposited. Explain the color changes 
that occur and write a structure for the crystalline product. 

22-3 Anthracene (mp 217°) forms a red crystalline complex (mp 164°) with 
1,3,5-trinitrobenzene (mp 121°). If you were to purify anthracene as this 
complex, how could you regenerate the anthracene free of trinitrobenzene ? 

22-4 N,N,4-Trimethylaniline has Z B = 3xlO- 9 ; quinuclidine, K B =4x\Q-*; 
and benzoquinuclidine, Z B =6xlO~ 7 . What conclusions may be drawn 
from these results as to the cause(s) of the reduced base strength of aromatic 
amines relative to saturated aliphatic or alicyclic amines? Explain. 





benzoquinuclidine 

Would you expect a nitro group meta or para to the nitrogen in benzo- 
quinuclidine to have as large an effect on the base strength of benzoquinu- 
clidine as the corresponding substitution in aniline ? 

22-5 Pure secondary aliphatic amines can often be prepared free of primary and 
tertiary amines by cleavage of a p-mtroso-N.N-dialkylaniline with strong 
alkali to p-nitrosophenol and the dialkylamine. Why does this cleavage occur 
readily? Show how the synthesis might be used for preparation of di-w- 
butylamine starting with aniline and «-butyl bromide. 



exercises 623 

22-6 N,N-Dimethylaniline, but not N,N,2,6-tetramethylaniline, couples readily 
with diazonium salts in neutral solution. Explain the low reactivity of 
N,N,2,6-tetramethylaniline by consideration of the geometry of the transi- 
tion state for the reaction. 

22-7 Some very reactive unsaturated hydrocarbons, such as azulene(Section20-7A), 
couple with diazonium salts. At which position would you expect azulene 
to couple most readily? Explain. 

22-8 1-Naphthol couples with benzenediazonium chloride in the 2 position; 
2-methyl-l-naphthol, in the 4 position; and 2-naphthol, in the 1 position. 
However, l-methyl-2-naphthol does not couple at all under the same condi- 
tions. Why? 

22-9 Give for each of the following pairs of compounds a chemical test, preferably 
a test tube reaction, that will distinguish the two compounds. Write a struc- 
tural formula for each compound and equations for the reactions involved : 

a. p-nitrotoluene and benzamide 

b. aniline and cyclohexylamine 

c. N-methylaniline and p-toluidine 

d. N-nitroso-N-methylaniline and />-nitroso-N-methylaniline 

22-10 Show by equations how each of the following substances might be synthesized 
starting from the indicated materials. Specify reagents and approximate 
reaction conditions. 

a. o-dinitrobenzene from benzene 

b. 2,6-dinitrophenol from benzene 

c. 2-amino-4-chlorotoluene from toluene 

d. ji-cyanonitrobenzene from benzene 

e. 2-amino-4-nitrophenol from phenol 
/. m-cyanotoluene from toluene 

22-11 Write structural formulas for substances (one for each part) that fit the 
following descriptions : 

a. an aromatic amine that is a stronger base than aniline 

b. a substituted phenol that would not be expected to couple with ben- 
zenediazonium chloride in acidic, alkaline, or neutral solution 

c. a substituted benzenediazonium chloride that would be a more active 
coupling agent than benzenediazonium chloride itself 

22-12 Explain why triphenylamine is a much weaker base than aniline and why its 
absorption spectrum is shifted to longer wavelengths compared with the 
spectrum of aniline (see Table 22-1). Would you expect N-phenylcarbazole 
to be a stronger or weaker base than triphenylamine? Explain. 




N-phenylcarbazole 



aryl oxygeii compounds 



chap 23 aryl oxygen compounds 627 

In the previous chapter, we indicated that, although there are considerable 
structural similarities between vinyl alcohols (enols) and phenols, and 
between vinylamines (enamines) and aromatic amines, the enols and enamines 
are generally unstable with respect to their keto and imine tautomeric forms, 
whereas the reverse is true of phenols and aromatic amines because of the 
stability associated with the aromatic ring. 

OH *jl O 

c=c — ^^ -c-c 

/ \ I \ 

H j_| 

O" 0H ~ Qr° 

In this chapter, after considering some of the more general procedures for 
the preparation of phenols, we shall take up the effect of the aromatic ring 
on the reactivity and reactions of the hydroxyl group of phenols and the effect 
of the hydroxyl group on the properties of the aromatic ring. The chapter 
concludes with discussions of the chemistry of quinones and of some non- 
benzenoid seven-membered ring substances with aromatic properties. 



23-1 synthesis and physical properties of phenols 

Considerable amounts of phenol and cresols (o-, m-, and ^-methylphenols) 
can be isolated from coal tar, which is formed in the destructive distillation of 
coal. Phenol itself is used commercially in such large quantities that alternate 
methods of synthesis are necessary. Direct oxidation of benzene is unsatis- 
factory because phenol is much more readily oxidized than is benzene. The 
more usual procedures are to sulfonate or chlorinate benzene and then intro- 
duce the hydroxyl group by nucleophilic substitution using strong alkali. 

/^/S0 3 H 

HiSOt (T ^f 1. NaOH fusion 




CI 2 




CI 




FeCI 3 IL ^fi 250° 



These reactions are general for introduction of hydroxyl substituents on 
aromatic rings; however, in some cases, they proceed by way of benzyne 
intermediates (Section 21 -3C) and may lead to rearrangement. 

A more recent commercial synthesis of phenol involves oxidation of iso- 
propylbenzene (cumene). This is made more commercially attractive by 
virtue of acetone being formed at the same time. The sequence of reactions 
starting with benzene and propene is shown in Figure 23-1. Some interesting 
chemistry is involved in this process. The first step, conversion of benzene 
to cumene, is a Friedel-Crafts alkylation (Section 20-4D). The second step, 



chap 23 aryl oxygen compounds 628 



f j + CH 3 CH=CH, 


H 3 C 
HaPO. (j^V ^CH 3 

u 






isopropylbenzene 
(cumene) 








o 2 


II + (CH 3 ) 2 C=0 < 


H s 


CH 3 
l/OOH 

fV^CH, 




U 








cumene hydroperoxide 



Figure 23-1 Commercial preparation of phenol and acetone starting with 
benzene and propene. 



conversion of cumene to the hydroperoxide, is a radical chain reaction 
(Section 2-5B). The third step is an acid-catalyzed rearrangement that resem- 
bles the Beckmann rearrangement (Section 16-1E2). (For more on the mecha- 
nism, see Exercises 23-13 and 23-14.) 

Phenol is a colorless crystalline solid when pure, but samples of it are often 
pink or brown because, like aniline, it is subject to air oxidation. Phenols are 
more polar and are able to form stronger hydrogen bonds than the corre- 
sponding saturated alcohols. A comparison of the physical properties of phe- 
nol and cyclohexanol shown in Table 23-1 shows that phenol has the higher 
melting point, higher boiling point, and higher water solubility, and is the 
more acidic. 

The acid dissociation constants and the ultraviolet spectral properties of 
phenols are shown in Table 23-2. There is a considerable effect of substituents 
on the wavelength and intensity of the absorption maxima. 



Table 23-1 Comparative physical properties 
of phenol and cyclohexanol 





phenol 


cyclohexanol 


mp 


43° 


26° 


bp 


181° 


161° 


water solubility, 


9.3 


3.6 


g/100g,20° 






Kha. 


1.0 x 10~ 10 


~10~ 18 



sec 23.1 synthesis and physical properties of phenols 629 
Table 23*2 Physical properties of some representative phenols 



formula 



H 2 0, 25°C X„ 



phenol 



ji-cresol 



salicylaldehyde 



CfiHsOJK 



/>-nitrophenol 


o 2 nh! f~ OH 




N0 2 

/ 


picric acid 


2 N-{ VoH 




\ 
N0 2 




OH 

/ 


catechol 


{_y° H 




HO 

\ 


resorcinol 


vJ^ OH 



1.3xlO~ 10 2105 6200 2700 1450 



H,C-<1 j>~OH 1.5xl0~ 10 2250 7400 2800 1995 



6.5xl0" 8 3175 10,000 



6 x 10 - 1 3800 13,450 



3.3xlO- 10 " 2140 6300 2755 2300 



3.6xl0~ 10 " 2160 6800 2735 1900 



hydroquinone HO— 4^ />— OH 



p-aminophenol H 2 N— ^ h— OH 




\ / 



OH 



CHO 



p-hydroxybenzal- 

dehyde HO 



1-naphthol 



OH 




2-naphthol 




OH 



1 x 10 - 10 



2900 2800 



6.6X10- 9 " 2330 8000 2800 3200 



3.0xl0~ 9 2560 12,600 3240 3400 



CHO 2.2xl0~ 8 2835 16,000 



4.9X10- 10 2325 33,000 2950 5000 



2.8X10" 10 2260 76,000 2735 4500 



« it 18° 



At 18°C. 



chap 23 aryl oxygen compounds 630 

23-2 some chemical properties of phenols 



A. REACTIONS INVOLVING O-H BONDS 

The acidity of phenols compared to alcohols can be accounted for by an 
argument similar to that used to explain the acidity of carboxylic acids 
(Section 13-4). There is a small amount of resonance stabilization in phenol 
that is due to derealization of one of the unshared electron pairs on oxygen 
over the aromatic ring, as can be described in terms of the resonance struc- 
tures [la] through [lc]. 



■ e „ n — * 






[la] [lb] [lc] 



Conversion of phenol by loss of the hydroxyl proton to phenoxide anion 
leads to much greater derealization of the unshared pair because, as can be 
seen from the resonance structures [2a] through [2c], no charge separation is 
involved of the type apparent in [la] through [lc]. 





[2a] [2b] [2c] 

The greater stabilization energy of the anion makes the ionization process 
energetically more favorable than for a saturated alcohol such as cyclohexa- 
nol. 

The reactions of the hydroxyl groups of phenols that involve breaking 
the O— H bonds and formation of new bonds from oxygen to carbon are 
generally similar to those of alcohols. It is possible to prepare esters with 
carboxylic acid anhydrides and to prepare ethers by reaction of phenoxide 
anions with halides, sulfate esters, sulfonates, and so on, which react well by 
S N 2 mechanisms. 

o 

OH JCH^i^ ^O-C-CH, 



\ ff acid or base \ // 

catalyst 



phenyl acetate 



/ \_ oh NaOH ) / \_0 N CH3l » /~~\— OCH 
\ — / \ — V \ (CH 3 0) 2 S0 2 s* \ 'J 



methoxybenzene 
(anisole) 



sec 23.2 some chemical properties of phenols 631 

Phenols are sufficiently acidic to be converted to methoxy derivatives with 
diazomethane (Section 16-6C) with no need for an acidic catalyst. 



OH + CH,N, 



ether 



<Q-OCH 3 



However, they are weaker than carboxylic acids (by a factor of 10 5 ) and 
this is the basis for separating phenols and carboxylic acids by extraction with 
aqueous bicarbonate solution. Carboxylic acids can be extracted from ether 
or benzene solution by this reagent whereas phenols can not (Section 13-1). 

Almost all phenols and enols (such as those of 1,3-diketones) give colors 
with ferric chloride in dilute water or alcohol solutions. Phenol itself produces 
a violet coloration with ferric chloride and the cresols give a blue color. The 
products are apparently ferric phenoxide salts, which absorb visible light to 
give an excited state having electrons delocalized over both the iron atom and 
the unsaturated system. 



B. C- vs. O-ALKYLATION OF PHENOLS 

The same type of problem with respect to O- and C-alkylation is encountered 
with phenoxide salts as with enolate anions (Section 12-2B). Normally, only 
O-alkylation is observed. However, with allyl halides either reaction can be 
made essentially the exclusive reaction by proper choice of solvent. With 
sodium phenoxide, more polar solvents such as acetone tend to lead to 
phenyl allyl ether while in nonpolar solvents, such as benzene, o-allylphenol is 
the favored product. 

I V()CH 2 CH=CH 2 
phenyl allyl ether 



polar 



solvents 




\ / 



-O Na + CH, = CH-CH 2 Br 



nonpolar 



solvents 




\ 



o 



CH 2 
I 
CH=CH, 



CH 2 CH=CH 2 

o-allylphenol 



Apparently, in nonpolar solvents, the lack of dissociation of the — ONa 
part of the phenoxide salts tends to increase the steric hindrance at oxygen 
and makes attack on the ring more favorable. 



chap 23 aryl oxygen compounds 632 

The C-allylation product is thermodynamically more stable than the O- 
allylation product, as shown by the fact that phenyl allyl ether rearranges 
to o-allylphenol above 200°. Such rearrangements are quite general and are 
called Claisen rearrangements. 



£H,CH=CH 2 





CH 2 CH = CH, 



It should be noted that C-allylation of sodium phenoxide as observed in 
nonpolar solvents is not the result of O-allylation followed by rearrangement, 
because the temperature of the allylation reaction is far too low to obtain the 
observed yield of o-allylphenol by rearrangement. 



C. REACTIONS INVOLVING THE C— O BONDS 

It is very difficult to break the aromatic C— O bond in reactions involving 
phenols or phenol derivatives. Thus, concentrated halogen acids do not 
convert phenols to aryl halides, and cleavage of phenyl alkyl ethers with 
hydrogen bromide or hydrogen iodide produces the phenol and an alkyl 
halide, not an aryl halide and an alcohol. Diaryl ethers, such as diphenyl ether, 
do not react with hydrogen iodide even at 200°. 

f VoH + CH 3 Br 



rv<> 




CH 3 + HBr 

^ / \-Br + CH3OH 

Such behavior is very much in line with the difficulty of breaking aromatic 
halogen bonds in nucleophilic reactions (Chapter 21). There is no very 
suitable way for converting phenols to aryl halides, except when activation is 
provided by ortho or para nitro groups. Thus, 2,4-dinitrophenol is smoothly 
converted to 2,4-dinitrochlorobenzene with phosphorus pentachloride. 




PCI5 




NO, 



D. REACTIONS OF THE AROMATIC RING 

The —OH and — O e groups of phenol and phenoxide ion make for easy 
electrophilic substitution. The situation here is very much like that in aniline 
(see Section 22-5; see also Section 20-5A). Phenols react rapidly with bromine 



sec 23.2 some chemical properties of phenols 633 

in aqueous solution to substitute the positions ortho or para to the hydroxyl 
group, phenol itself giving 2,4,6-tribromophenol in high yield. 



Br 




\ / 



OH 



Br 2 



H 2 



Br A f~ 0H 
Br 



A number of important reactions of phenols involve electrophilic aromatic 
substitution of phenoxide ions. One example, which we have discussed in the 
previous chapter, is the diazo coupling reaction (Section 22-9B). Another 
example, which looks quite unrelated, is the Kolbe reaction (Figure 23-2) in 
which carbon dioxide reacts with sodium phenoxide to give the sodium salt 
of o-hydroxybenzoic acid (salicylic acid). 

Sodium phenoxide absorbs carbon dioxide at room temperature to form 
sodium phenyl carbonate and, when this is heated to 125° under a pressure of 
several atmospheres of carbon dioxide, it rearranges to sodium salicylate. 
However, there is no reason to expect that this reaction is anything other than 
a dissociation-recombination process, in which the important step involves 
electrophilic attack by carbon dioxide on the aromatic ring of phenoxide ion. 





e e 
CO,Na 



HO 



e e 
CO,Na 



With sodium phenoxide and temperatures of 125° to 150°, ortho substi- 
tution occurs ; at higher temperatures (250° to 300°) and particularly with the 
potassium salt, the para isomer is favored. 



Figure 23-2 The reaction of sodium phenoxide with carbon dioxide (the 
Kolbe reaction). 



O e Na® 


0H e e 


itS 


^L/ c °2 Na 


K^ ™° 


u 




!■ 


sodium salicylate 


co 2 






O 

II e © 
O-C — O Na 

A, 









sodium phenyl carbonate 





chap 23 aryl oxygen compounds 634 

Many substances such as salicylaldehyde, salicylic acid, and o-nitrophenol 
that have hydroxyl groups ortho to some substituent to which they can 
form hydrogen bonds have exceptional physical properties compared with the 
meta or para isomers. This is because formation of intra- rather than inter- 
molecular hydrogen bonds reduces intermolecular attraction, thus reducing 
boiling points, increasing solubility in nonpolar solvents, and so on. Com- 
pounds with intramolecular hydrogen bonds are often said to be chelated 
(Gk. chele, claw) and the resulting ring is called a chelate ring. 





OH 



intramolecular only intermolecular 
hydrogen bond hydrogen bonds 



Table 23-3 Physical properties of some o, m, and p 
disubstituted benzene derivatives 



compound 



bp,°C 



mp, 



°C 



volatility with 
steam 



NO, 



ortho 
meta 
para 

ortho 
meta 
para 

ortho 
meta 
para 

ortho 
meta 
para 

ortho 
meta 
para 



191 

203 
202 



196.5 

240 



244 
230 
248 



211 2 



216 

j 9470mm 



31 
12 
35 



-7 
108 
117 



38 

2.5 



158 
201 
215 

45 

97 

114 



+ + 
+ + 



+ 



+ 
+ 
+ 

+ 



+ 



sec 23.3 polyhydric phenols 635 

The physical constants for the different isomers of some substances that can 
and cannot form reasonably strong intramolecular hydrogen bonds are given 
in Table 23-3. It will be seen that intramolecular hydrogen bonding between 
suitable ortho groups has the effect of reducing both the melting and boiling 
points. An important practical use of this is often made in isomer separations, 
because many of the substances which can form intramolecular hydrogen 
bonds turn out to be volatile with steam, whereas the corresponding meta and 
para isomers are much less so. 

Formation of intramolecular hydrogen bonds shows up clearly in nmr 
spectra, as we have seen before in the case of the enol forms of 1,3-dicarbonyl 
compounds (Section 12-6). Figure 23-3 shows that there is a difference of 2.3 
ppm between the O— H resonance positions of o-nitrophenol and /7-nitro- 
phenol. Intramolecular hydrogen-bond formation also influences the OH 
stretching frequencies in the infrared. 

Phenols generally can be successfully reduced with hydrogen over nickel 
catalysts to the corresponding cyclohexanols. A variety of alkyl-substituted 
cyclohexanols can be prepared in this way. 




OH 




H 2 (Ni) 



H3C-C-CH3 H3C-C-CH3 



CH 3 CH 3 



23.3 polyhydric phenols 

A number of important aromatic compounds have more than one phenolic 
hydroxyl group. These are most often derivatives of the following dihydric 
and trihydric phenols, all of which have commonly used but poorly descrip- 
tive names. 



H(X /^ /OH HO^ /-^ ^OH 






OH OH 

catechol resorcinol hydroquinone pyrogallol phloroglucinol 

The polyhydric phenols with the hydroxyls in the ortho or para relationship 
are normally easily oxidized to quinones — the chemistry of which substances 
will be discussed shortly. 

L JL m 8 so 4 L 1 

^-^OH ether ^-^O 

o-benzoquinone 



chap 23 aryl oxygen compounds 636 



600 



...u 



OH 




500 



400 117 



1 




ijfri 



ft^Y-*"** 



(b) 



IK.) 



**Jr***-***~**j~»*^ l jj-**» r H* t ^ 




W w 



■» • l W»^v<> r « 1 ^A» 



(C) 



Mill 



^■ ^ ^* ••+m*'*ir , +m*-*1* 




10.0 



7.0 ppm 



Figure 23-3 Nmr spectra at 60 MHz of o-nitrophenol (a), m-nitrophenol (b), 
and p-nitrophenol (c) in diethyl ether solution (the solvent bands are not 
shown). 



sec 23.4 quinones 637 




OH 

p-benzoquinone 

The m-dihydroxybenzenes undergo oxidation but do not give m^quinones, 
since these are substances for which no single unstrained planar structure can 
be written. Oxidation of resorcinol gives complex products — probably by way 
of attack at the 4 position, which is activated by being ortho to one hydroxyl 
and para to the other. The use of hydroquinone and related substances as 
reducing agents for silver bromide in photography will be discussed later. 

Substitution of more than one hydroxyl group on an aromatic ring tends to 
make the ring particularly susceptible to electrophilic substitution, especially 
when the hydroxyls are meta to one another, in which circumstance their 
activating influences reinforce one another. For this reason, resorcinol and 
phloroglucinol are exceptionally reactive toward electrophilic reagents, 
particularly in alkaline solution. 



23-4 quinones 



Strictly speaking, quinones are conjugated cyclic diketones rather than aro- 
matic compounds ; hence a discussion of the properties of quinones is, to a 
degree, out of place in a chapter covering aromatic oxygen compounds, even 
though quinones have more stability than expected on the basis of bond 
energies alone. Thus, /7-benzoquinone has a stabilization energy of 5 kcal, 
which can be ascribed largely to resonance structures such as [3], there being a 
total of four polar forms equivalent to [3]. The fact that quinones and poly- 





hydric phenols are normally very readily interconvertible results in the 
chemistry of either class of compound being difficult to disentangle from the 
other; consequently, we shall discuss quinones at this point. 

A variety of quinones have been prepared, the most common of which are 
the 1,2- and 1,4-quinones as exemplified by o-benzoquinone and ^-benzo- 
quinone. Usually the 1,2-quinones are more difficult to make and are more 
reactive than the 1,4-quinones. A few 1,6- and 1,8-quinones are also known. 





chap 23 aryl oxygen compounds 638 



CI 

,5-dichloro-2,6-naphthoquinone 3, 1 0-pyrenequinone 



A. REDUCTION OF QUINONES 

The most characteristic and important reaction of quinones is reduction to 
the corresponding dihydroxyaromatic compounds. 



2e 

(23-1) 





-2e 



These reductions are sufficiently rapid and reversible to give easily repro- 
ducible electrode potentials in an electrolytic cell. The position of the quinone- 
hydroquinone equilibrium (Equation 23-1) is proportional to the square of the 
hydrogen ion concentration. The electrode potential is therefore sensitive to 
pH, a change of one unit of pH in water solution changing the potential by 
0.059 volt. Before the invention of the glass electrode pH meter, the half-cell 
potential developed by the quinone-hydroquinone equilibrium was widely 
used to determine pH values of aqueous solutions. The method is not very 
good above pH 9 or 10 because quinone reacts irreversibly with alkali. 

Numerous studies have been made of half-cell potentials for the reduction 
of quinones. As might be expected, the potentials are greatest when the 
greatest gain in resonance stabilization is associated with formation of the 
aromatic ring. 

The hydroquinone-quinone oxidation-reduction system is actually some- 
what more complicated than presented above. This is evident in one way 
from the fact that mixing alcoholic solutions of hydroquinone and quinone 
gives a brown-red solution, which then deposits a crystalline green-black 1 : 1 
complex known as quinhydrone. This substance is apparently a charge-transfer 
complex (of the type discussed in Section 22-4) with the hydroquinone acting 
as the electron donor and the quinone as the electron acceptor. Quinhydrone 
is not very soluble and dissociates considerably to its components in solution. 

The reduction of quinone requires two electrons, and it is of course possible 
that these electrons could be transferred either together or one at a time. The 
product of a single electron transfer leads to what is appropriately called a 
semiquinone [4] with both a negative charge and an odd electron. The forma- 



sec 23.4 quinones 639 




+ e 





dimer 



semiquinone 
[4] 

tion of relatively stable semiquinone radicals by electrolytic reduction of 
quinones has been established by a variety of methods. Some semiquinone 
radicals undergo reversible dimerization reactions to form peroxides. 



B. PHOTOGRAPHIC DEVELOPERS 

A particularly important practical use of the hydroquinone-quinone oxida- 
tion-reduction system is in photography. Exposure of the minute grains of 
silver bromide in a photographic emulsion to blue light (or any visible light 
in the presence of suitable sensitizing dyes ; see Chapter 26) produces a stable 
activated form of silver bromide, the activation probably involving generation 
of some sort of crystal defect. Subsequently, when the emulsion is brought 
into contact with a developer, which may be an alkaline aqueous solution of 
hydroquinone and sodium sulfite, the particles of activated silver bromide are 
reduced to silver metal much more rapidly than the ordinary silver bromide. 
Removal of the unreduced silver bromide with sodium thiosulfate ("fixing") 
leaves a suspension of finely divided silver in the emulsion in the form of the 
familiar photographic negative. 



OH 




+ 2 AgBr + 2 OH 



OH 




+ 2 Ag + 



e 
2 Br 



+ 2H 2 



AgBr* = activated silver bromide 



C. ADDITION REACTIONS OF QUINONES 

Being oe,/?-unsaturated ketones, quinones are expected to have the possibility 
of forming 1,4-addition products in the same way as their open-chain analogs 
(Section 12-3). ^-Benzoquinone itself undergoes such additions rather readily. 
Two examples are provided in Figure 23-4 by the addition of hydrogen 
chloride and the acid-catalyzed addition of acetic anhydride. In the second 
reaction, the hydroxyl groups of the hydroquinone are acetylated by the 
acetic anhydride. Hydrolysis of the product affords hydroxyhydroquinone. 



chap 23 aryl oxygen compounds 640 



HC1 



OH 




H® 



(CH 3 CO) 2 



HO 




OH 




O-C-CHj 




O— C — CH, 



+ CH,CO,H 



Figure 23-4 Some addition reactions ofp-benzoquinone. 



D. NATURALLY OCCURRING QUINONES 

Many naturally occurring substances have quinone-type structures, one of 
the most important being the blood antihemorrhagic factor, vitamin K l5 
which occurs in green plants and is a substituted 1,4-naphthoquinone. The 
structure of vitamin K t has been established by degradation and by syn- 
thesis. Surprisingly, the long alkyl side chain of vitamin K t is not necessary for 
its action in aiding blood clotting because 2-methyl- 1,4-naphthoquinone is 
almost equally active on a molar basis. 




CH, 

I 



CH, 



CH,CH=C-eCH 2 -CH 2 -CH 2 -CH^CH 3 



vitamin K, 




CH, 



2-methyl- 1, 4-naphthoquinone 
(menadione) 

A molecule that is structurally similar to vitamin K L is coenzyme Q. There 
are, in fact, several of these differing in the length of the side chain. Coenzyme 
Q 10 is shown below, the subscript in the name revealing the number of repeat- 
ing C 5 units in the side chain. The Q coenzymes are widely distributed in 
nature and are particularly important constituents of mitochondria. They are 



sec 23. S tropolones and related compounds 641 

links in the so-called electron-transport chain (Section 184), appearing 
between the flavins and the cytochromes, and seem to be able to function 
both as one- and two-equivalent oxidants. 



CH,0 




CH 2 CH=C-fCH,-CH 2 -CH=C^- CH 3 



coenzyme Q 10 

The repeating C 5 moieties in the side chains of vitamin K x and coenzyme Q 
are referred to as isoprenoid units. These will be met again in Chapter 29. 



23-5 tropolones and related compounds 

The tropolones make up a very interesting class of nonbenzenoid aromatic 
compounds which were first encountered in several quite different kinds of 
natural products. As one example of a naturally occurring tropolone, the 
substance called /?-thujaplicin or hinokitiol has been isolated from the oil of 
the western red cedar. The wood of these trees rots extremely slowly and this 
characteristic has been traced to the presence of this compound, and its y 
isomer, which are natural fungicides. The outer butt heartwood of older 
cedars contains as much as 1 % thujaplicins whereas young trees lay down 
very little of these materials. This accounts for the hollow center often seen 
in very old cedars — the core, which has a low thujaplicin content, rots. 




/Mhujaplicin 
(4-isopropyltropolone) 

Tropolone itself can be prepared in a number of ways, the most convenient 
of which involves oxidation of 1,3,5-cycloheptatriene (" tropilidene ") with 
alkaline potassium permanganate. The yield is low but the product is readily 
isolated as the copper salt. 



Cu 2S> 
KMnO^ [/ ^ ^=± copper salt 

H 2 S 

tropilidene tropolone 

Tropolone is an acid with an ionization constant of 10~ 7 , which is inter- 





chap 23 aryl oxygen compounds 642 

mediate between that of acetic acid and that of phenol. Like phenols, tropo- 
lones form colored complexes with ferric chloride solution. Tropolone has 
many properties which suggest that it has some aromatic character. Thus, it 
resists hydrogenation, undergoes diazo coupling, and can be nitrated, sul- 
fonated, and substituted with halogens. Its stability can be attributed to 
resonance involving the two nonequivalent structures [5] and [6] and to the 
several structures such as [7] and [8] which correspond to the stable tropylium 
cation with six n electrons (Section 6-7). There is a strong intramolecular 
hydrogen bond between the carbonyl oxygen and the hydroxylic proton 
and this fact reflects the importance of structure [6]. 



&~Cf 



[5] 



[6] 




OH 



[7] 



e O 

® 
[8] 



etc. 



The tropylium cation itself is easily prepared by transfer of hydride ion 
from tropilidine to triphenylmethyl cation in sulfur dioxide solution. 




(QH 5 ) 3 C® 



so 2 




I A + (C 6 H 5 ) 3 CH 



tropylium cation 



Seven equivalent resonance structures can be written for the cation so that 
only one-seventh the positive charge is expected to be on each carbon. Since 
the cation also has just six n electrons, it is anticipated to be unusually stable 
for a carbonium ion. 




hybrid structure for 
tropylium ion 



summary 

Phenols, ArOH, although enols, are stable because of the stabilization energy 
of the benzene ring. Phenol can be prepared from benzene via benzenesulfonic 
acid or via halobenzenes. 



Phenols resemble alcohols in forming esters and ethers but their considera- 



summary 643 

bly greater acidity (midway between alcohols and carboxylic acids) allows them 
to form salts with sodium hydroxide, though not with sodium bicarbonate. 

O 

II 
ArCOR 



ArOH > ArOR 



^ ArO e 

Ether formation via phenoxide ion may be accompanied by C-alkylation. 
Cleavage of an aryl alky] ether with hydrogen halide always gives alkyl halide 
and a phenol. 

The hydroxyl group in phenols activates the ring toward electrophilic 
attack. Ionization of the hydroxyl causes further activation and enables feeble 
electrophiles such as carbon dioxide to react (Kolbe reaction). 





Intramolecular hydrogen bonding in or^o-substituted phenols is revealed 
by downfield nmr spectral shifts and by higher vapor pressures in comparison 
to analogous para compounds. 

A number of polyhydric phenols are known; those with two hydroxyl 
groups ortho or para to one another can usually be oxidized to quinones. A 
quinone can undergo a number of addition reactions. 

O e 



OH 

[O] 





(semiquinone) 



OH 

Two important naturally occurring quinones are vitamin Kj and coenzyme 
Q. Tropolones, including the natural product /Mhujaplicin, are seven-mem- 
bered cyclic enols. A major factor in stabilizing the enol form is intramolecu- 
lar hydrogen bonding. Resonance stabilization as in the tropylium ion may 
also be important. 





(tropolone) (tropylium ion) 



chap 23 aryl oxygen compounds 644 

exercises 

23*1 Would you expect phenyl acetate to be hydrolyzed more readily or less 
readily than cyclohexyl acetate in alkaline solution ? Use reasoning based on 
the mechanism of ester hydrolysis (Section 13 '8). 

23-2 Rearrangement of phenyl allyl-3- 14 C ether at 200° gives o-allyl-l- l4 C- 
phenol. What does this tell you about the rearrangement mechanism? Can 
it be a dissociation-recombination process ? What product(s) would you ex- 
pect from a para Claisen rearrangement of 2,6-dimethylphenyl allyl-3- l4 C 
ether? From 2,6-diallylphenyl allyl-3- 14 C ether? 

23-3 Explain why phenol with bromine gives tribromophenol readily in water 
solution and o- and ^-monobromophenols in nonpolar solvents. Note that 
2,4,6-tribromophenol is at least a 300-fold stronger acid than phenol in 
water solution. 

23-4 The herbicide 2,4-D is 2,4-dichlorophenoxyacetic acid (Figure 21-2). Show 
how this substance might be synthesized starting from phenol and acetic 
acid. 

23-5 How much difference in physical properties would you expect for o- and 
p-cyanophenol isomers ? Explain. 

23-6 Resorcinol (/n-dihydroxybenzene) can be converted to a carboxylic acid with 
carbon dioxide and alkali. Would you expect resorcinol to react more or less 
readily than phenol ? Why ? Which is the most likely point of monosubsti- 
tution ? Explain. 

23-7 Arrange the following quinones in order of increasing half-cell potential 
expected for reduction: />-benzoquinone, 4,4'-diphenoquinone, cw-2,2'-di- 
phenoquinone, 9,10-anthraquinone, and 1 ,4-naphthoquinone. Your reason- 
ing should be based on difference in stabilization of the quinones and the 
hydroquinones, including steric factors (if any). 

23-8 Tropone (2,4,6-cycloheptatrienone) is an exceptionally strong base for a 
ketone. Explain. 

23-9 At which position would you expect tropolone to substitute most readily 
with nitric acid ? Explain. 

23-10 Give for each of the following pairs of compounds a chemical test, preferably 
a test tube reaction, that will distinguish between the two compounds. Write 
a structural formula for each compound and equations for the reactions 
involved. 

a. phenol and cyclohexanol 

b. methyl p-hydroxybenzoate and i>-methoxybenzoic acid 

c. hydroquinone and resorcinol 

d. hydroquinone and tropolone 



exercises 645 

23-11 Show by means of equations how each of the following substances might 
be synthesized, starting from the indicated materials. Specify reagents and 
approximate reaction conditions. 

a. methyl 2-methoxybenzoate from phenol 

b. 2,6-dibromo-4-/-butylanisole from phenol 

c. 2-hydroxy-5-nitrobenzoic acid from phenol 

d. 4-cyanophenoxyacetic acid from phenol 

e. cyanoquinone from hydroquinone 

23-12 Write structural formulas for substances (one for each part) that fit the 
following descriptions : 

a. a phenol that would be a stronger acid than phenol itself 

b. that isomer of dichlorophenol that is the strongest acid 

c. the Claisen rearrangement product from a-methylallyl-2,6-dimethyl- 
phenyl ether 

d. the Claisen-type rearrangement product from allyl 2,6-dimethyl-4- 
(j8-methylvinyl)-phenyl ether. 

e. a quinone that would be a better charge-transfer agent than quinone 
itself 

/. the expected product from addition of hydrogen cyanide to mono- 
cyan oquinone 

g. a nonbenzenoid, quinone-like substance with its carbonyl groups in 
a 1,3 relationship 

23-13 The chain reaction involved in the conversion of isopropylbenzene (cumene) 
to its hydroperoxide (Figure 23-1) can be written as follows, using a 7-butoxy 
radical (formed by the decomposition of di-^-butyl peroxide) as the initiator. 

C 6 H 5 CH(CH 3 ) 2 -M-BuO- ► A + /-BuOH initiation step 



A + 2 > B 

OOH 

I 
B + C 6 H 5 CH(CH 3 ) 2 ► A + C 6 H 5 C(CH 3 ) 2 



propagation 
steps 



a. Deduce the structure of A and B and suggest a reason for A, rather 
than one of its structural isomers, being formed in the initiation step. 

b. Suggest a termination step (Section 2-5B) for the chain. 

23-14 The acid-catalyzed rearrangement involved in the conversion of cumene 
hydroperoxide to phenol and acetone (Figure 23-1) can be written as follows : 



OOH OOH 2 

I I -H 2 © 

C 6 H 5 C(CH 3 ) 2 +H ffl < > C 6 H 5 C(CH 3 ) 2 ► C 6 H s O=C(CH 3 ) 2 

+ H 2 -H«> 
► C ► D ► C 6 H 5 OH + (CH 3 ) 2 C=0 



a. Write a structure for the transition state for the rearrangement that 
occurs in the second step. 

b. What are the structures of the cation C and the neutral molecule D 
and to what class of compounds does D belong? 



chap 23 aryl oxygen compounds 646 

23-15 Reduction of 9,10-anthraquinone with tin and hydrochloric acid in acetic 
acid produces a solid, light-yellow ketone, mp 156°, which has the formula 
Ci<tH 10 O. This ketone is not soluble in cold alkali but does dissolve when 
heated with alkali. Acidification of cooled alkaline solutions of the ketone 
precipitates a brown-yellow isomer of the ketone of mp 120°, which gives 
a color with ferric chloride, couples with diazonium salts, reacts with bro- 
mine, and slowly is reconverted to the ketone. 

What are the structures of the ketone and its isomer ? Write equations for 
the reactions described. 

23-16 Devise syntheses of each of the following photographic developing agents 
based on benzene as the aromatic starting material. Give approximate 
reaction conditions and reagents. 

a. hydroquinone 

b. /)-aminophenol 

c. p-amino-N,N-diethylaniline 

d. (p-hydroxyphenyl)-aminoacetic acid 

e. 2,4-diaminophenol 

23-17 Addition of hydrogen chloride to /?-benzoquinone yields some 2,3,5,6-tetra- 
chloroquinone. Explain how the latter could be formed in the absence of 
an external oxidizing agent. 

23-18 Consider the possibility of benzilic acid-type rearrangements of 9, 1 0-phenan- 
threnequinone and anthraquinone. Give your reasoning. 

23-19 When quinone is treated with hydroxylamine and phenol is treated with 
nitrous acid, the same compound of formula C 6 H 5 2 N is produced. What 
is the likely structure of this compound and how would you establish its 
correctness ? 

23-20 How would you expect the properties of 3- and 4-hydroxy-2,4,6-cyclohepta- 
trienone to compare with those of tropolone ? Explain. 

23-21 Make an atomic orbital model of phenol, showing in detail the orbitals and 
electrons at the oxygen atom (it may be desirable to review Chapters 6 and 20 
in connection with this problem). From your model, would you expect either 
or both pairs of unshared electrons on oxygen to be delocalized over the 
ring ? What would be the most favorable orientation of the hydrogen of the 
hydroxyl group for maximum derealization of an unshared electron pair ? 



aromatic side~chai» 



chap 24 aromatic side-chain derivatives 649 

The pronounced modification in the reactivity of halogen, amino, and 
hydroxyl substituents when linked to aromatic carbon rather than saturated 
carbon was discussed in Chapters 21, 22, and 23. Other substituents, par- 
ticularly those linked to an aromatic ring through a carbon-carbon bond, are 
also influenced by the ring, although usually to a lesser degree. Examples 
include -CH 2 OH, -CH 2 OCH 3 , -CH 2 C1, -CHO, -COCH 3 , -C0 2 H, 
and — CN, and we shall refer to aromatic compounds containing substi- 
tuents of this type as aromatic side-chain derivatives. 



preparation of aromatic side-chain compounds 

Since the utility of any method of synthesis is limited by the accessibility of the 
starting materials, we may anticipate that the most practical methods for the 
preparation of benzenoid side-chain compounds will start from benzene or an 
alkylbenzene. These methods may be divided into two categories — those that 
modify an existing side chain, and those by which a side chain is introduced 
through substitution of the aromatic ring. We shall consider first the reactions 
that modify a side chain and for which the obvious starting materials are the 
alkylbenzenes, especially toluene and the xylenes. 



24-1 aromatic carboxjlic acids 



An alkylbenzene can be converted to benzoic acid by oxidation of the side 
chain with reagents such as potassium permanganate, potassium dichromate, 
or nitric acid. 

C0 2 H 

KMnO„,H 2 0, OH 



(reflux) 

benzoic acid 

Under the conditions of oxidation, higher alkyl or alkenyl groups are 
degraded and ring substituents, other than halogen and nitro groups, often 
fail to survive. 

In fact, the presence of a hydroxyl or amino substituent causes the whole 
ring to be degraded long before the alkyl group undergoes appreciable 
oxidation. 



CH, CO,H 




H,N ^ H,N 




By contrast, prolonged oxidation of 5-nitro-2-indanone gives a good yield 
of product with the substituent untouched and the ring intact. 




dil. hno 3 




, ^ ^ ^ 24hr, 25° , ^ ^ _ 

2 N ^-" 2 N ^^ "C0 2 H 

5-nitro-2-indanone 4-nitrophthalic acid 



chap 24 aromatic side-chain derivatives 650 

To retain a side-chain substituent, selective methods of oxidation are 
required. For example, /7-toluic acid may be prepared from p-tolyl methyl 
ketone by the haloform reaction (Section 12-1C). 

H 3 C-^y C OCH3 ^^ H,C^Q-CO a H 

The Cannizzaro reaction (Section 11-4H) is sometimes useful for the prep- 
aration of substituted benzoic acids and (or) benzyl alcohols, provided that 
the starting aldehyde is available. 



H 2 0, SQH 






^OH "^ "OH ^ OH 

2-iodo-3-hydroxy- 2-iodo-3-hydroxy- 2-iodo-3-hydroxy- 

benzaldehyde benzoic acid benzyl alcohol 

80% 80% 

24-2 preparation of side-chain aromatic 
halogen compounds 

Although many side-chain halogen compounds can be synthesized by reac- 
tions that are also applicable to alkyl halides, there are several other methods 
especially useful for the preparation of arylmethyl halides. The most important 
of these are the radical halogenation of alkylbenzenes and chloromethylation 
of aromatic compounds (Section 24-4B). 

The light-induced, radical chlorination or bromination of alkylbenzenes 
with molecular chlorine or bromine gives substitution on the side chain rather 
than on the ring. Thus, toluene reacts with chlorine to give successively benzyl 
chloride, benzal chloride, and benzotrichloride. 



C6H5CH3 


Cl 2 


C.6H5CH2CI 


Ch 


C6H5CHC12 


ci 2 




hv 


hv 


hv 


' ^etis^^-h 


toluene 




benzyl 
chloride 




benzal 
chloride 




benzo- 
trichloride 



This reaction was met when the chlorination of methane was discussed in 
Chapter 2. The major effect of the phenyl ring is to facilitate the reaction by 
making the intermediate benzyl radicals more stable. 

CH,- CH, 



24-3 side-chain compounds derived from 
arylmethyl halides 

Arylmethyl halides, such as benzyl chloride (C 6 H 5 CH 2 C1), benzal chloride 
(C 6 H 5 CHC1 2 ), and benzotrichloride are quite reactive compounds. They are 



sec 24.4 preparation of aromatic side-chain compounds 6S1 



CH,C1 




benzyl chloride 



/ VCHC1 2 + H 2 
benzal chloride 



f VCH 2 OH 
benzyl alcohol 

f VCH 2 CN 
phenylacetonitrile 



100° 



CHO + 2 HC1 
benzaldehyde 



CC1 3 

benzotrichloride 



+ 2H 2 10 °° » <f V"C0 2 H + 3HC1 



benzoic acid 



Figure 24-1 Reactions of benzyl chloride, benzal chloride, and benzotri- 
chloride. 



readily available or easily prepared and are useful intermediates for the 
synthesis of other side-chain derivatives. See Figure 24-1 for examples. 

24-4 preparation of aromatic side-chain compounds 
by ring substitution 

A. FRIEDEL-CRAFTS REACTION 

The Friedel-Crafts alkylation and acylation reactions have been discussed 
(Sections 20-4D and 20-4E). For alkylation, catalytic amounts of A1C1 3 are 



+ RCl 



AICI, 



+ HC1 



usually sufficient and polysubstitution may be an important side reaction 
because of the activating effect of the R group. 

For acylation, large amounts of A1C1 3 are required and only monosubsti- 

O 

II 
tution occurs because of the deactivating effect of the — C— R group. 



o 

II Aids 
+ RC-Cl 




O 

II 
CR 



+ HC1 



chap 24 aromatic side-chain derivatives 652 



B. CHLOROMETHYLATION 



The reaction of an aromatic compound with formaldehyde and hydrogen 
chloride in the presence of zinc chloride as catalyst results in the substitution 
of a chloromethyl group, — CH 2 C1, for a ring hydrogen. 



QH 6 + CH 2 + HC1 



ZnCh 



-» C 6 H 5 CH 2 C1 + H 2 



The mechanism of the chloromethylation reaction is related to that of 
Friedel-Crafts alkylation and acylation and probably involves an incipient 
chloromethyl cation, ®CH 2 C1. 



O 



OH 



H— C + HC1 ;= 



\ 
H 

8e 
HOZnCl 2 



H CH,C1 



+ HOZnCl, 



Se 
HOZnCl, 



I ZnCI 2 : 

H — C— CI ^- ' H — C— CI 



CH,C1 



+ ZnCl, 



C. ALDEHYDES BY FORMYLATION 

Substitution of the carboxaldehyde group (— CHO) into an aromatic ring is 
known as formylation. This is accomplished by reaction of an aromatic 
hydrocarbon with carbon monoxide in the presence of hydrogen chloride and 
aluminum chloride. Cuprous chloride is also required for reactions proceeding 
at atmospheric pressure but is not necessary for reactions at elevated pressures. 



CH(CH 3 ) 2 



+ CO 



HCI, AICI3 (1 mole) 
500 psi, 25°-30° 



CH(CH 3 ) 2 



CHO 



p-isopropylbenzaldehyde 
60% 



CO 



HCI, AICI3 
Cu 2 CI 2 , 35°-40° ' \ / 



-CHO 



p-phenylbenzaldehyde 

73% 



Formylation of reactive aromatic compounds such as phenols, phenolic 
ethers, and certain hydrocarbons can be brought about by the action of 
hydrogen cyanide, hydrogen chloride, and a catalyst, usually zinc chloride or 
aluminum chloride. A convenient alternative is to use zinc cyanide and 
hydrogen chloride. The product is then hydrolyzed to an aldehyde. 



sec 24.5 arylmethyl halides 653 



CHO 




OH 



+ HCN + HC1 



I. ZnCl 2 , ether. 



2. H 2 




OH 



+ NHX1 



2-naphthol 



2-hy droxy- 1 -naphthaldehyde 



properties of aromatic side-chain derivatives 

24-5 arylmethyl halides. stable carbonium ions, 
carbanions, and radicals 

The arylmethyl halides of particular interest are those having both halogen 
and aryl substituents bonded to the same saturated carbon. Typical examples 
and their physical properties are listed in Table 24-1. 

We noted in Chapter 21 that benzyl halides (C 6 H 5 CH 2 X) are comparable 
in both S N 1 and S N 2 reactivity to allyl halides (CH 2 =CHCH 2 X) and, because 
high reactivity in S N 1 reactions is associated primarily with exceptional car- 
bonium ion stability, the reactivity of benzyl derivatives can be ascribed 
mainly to resonance stabilization of the benzyl cation. Diphenylmethyl or 
benzhydryl halides, (C 6 H 5 ) 2 CHX, are still more reactive than benzyl halides in 
S N 1 reactions, and this is reasonable because the diphenylmethyl cation has 
two phenyl groups over which the positive charge can be delocalized and is, 
therefore, more stable relative to the starting halide than is the benzyl cation. 




"^ 





CH 





CH 




etc. 



diphenylmethyl cation 



Table 24-1 Physical properties of* arylmethyl halides 







bp, 


mp, 


d 20 '*, 


compound 


formula 


°C 


°C 


g/ml 


benzyl fluoride 


CeHsCH^F 


140 


-35 


1.0228 25 '* 


benzyl chloride 


CgHsCH^Cl 


179 


-43 


1.1026 18 '* 


benzyl bromide 


CeH 5 CH^Br 


198 


-4.0 


1.438"'° 


benzyl iodide 


C 6 H 5 CH 2 I 


0310mm 


24 


1.733"'* 


benzal chloride 


C6H5CHCI2 


207 


-16 


1.2557 1 * 


benzotrichloride 


C 6 H 5 CC1 3 


214 


-22 


1.38 


benzotrifluoride 


C 6 H 5 CF 3 


103 


-29.1 


1.1 886 20 


benzhydryl chloride 


(C 6 H 5 ) 2 CHC1 


173 19mm 


20.5 




(diphenylmethyl chloride) 










triphenylmethyl chloride 


(C 6 H 5 ) 3 CC1 




112.3 




(trityl chloride) 











chap 24 aromatic side-chain derivatives 654 

Accordingly, we might expect triphenylmethyl or trityl halides, (C 6 H 5 ) 3 C 
—X, to be more reactive yet. In fact, the C— X bonds of such compounds are 
sufficiently labile that reversible ionization occurs in solvents that have reason- 
ably high dielectric constants but do not react irreversibly with the carbonium 
ion. An example of such a solvent is liquid sulfur dioxide, and the degrees of 
ionization of a number of triarylmethyl halides in this solvent have been 
determined by electrical-conductance measurements, although the equilibria 
are complicated by ion-pair association. 

so 2 , o° 
(C 6 H 5 ) 3 C-C1 -r=- (C 6 H 5 ) 3 C®Cie ;=± (C 6 H S ) 3 C® + Cje 

ion pair dissociated ions 

Triarylmethyl cations are among the most stable carbonium ions known. 
They are intensely colored and are readily formed when the corresponding 
triarylcarbinol is dissolved in strong acids. 

H2SO4 e (-H20) 

(QH 5 ) 3 C-OH . (C 6 H 5 ) 3 C-OH 2 . (C 6 H 5 ) 3 C® 

triphenylcarbinol triphenylmethyl cation 

(colorless) (orange-yellow) 

If electron-donating para substituents such as amino groups are placed in 
each ring [1] the energy of the carbonium ion is lowered to such an extent that 
it is stable in water at pH 7. 

Q>-NH 2 C 6 H 4 ) 3 C® 

[1] 

In addition to forming stable cations, triarylmethyl compounds form stable 
carbanions. Because of this, the corresponding hydrocarbons are relatively 
acidic compared to simple alkanes. They react readily with strong bases such 
as sodamide, and the resulting carbanions are usually intensely colored. 

ether 

(QH 5 ) 3 CH + Na e NH 2 e , (C 6 H 5 ) 3 C: e Na® +NH 3 

triphenylmethane sodium triphenylmethide 

(colorless) (blood red) 

This carbanion can also be generated by the action of less basic reagents 
such as sodium ethoxide, provided polar aprotic solvents such as dimethyl 
sulfoxide or hexamethylphosphoramide are used (Sections 8-1 ID and 19-2C). 
Just as the positive charge in triarylmethyl cations can be distributed over the 
ortho and para positions of each ring, so can the negative charge in the tri- 
phenylmethide ion. 

Triarylmethyl compounds also form stable triarylmethyl radicals, and 
indeed the first stable carbon radical to be reported was the triphenylmethyl 
radical, (C 6 H 5 ) 3 0, prepared inadvertently by Gomberg in 1900. Gomberg's 
objective was to prepare hexaphenylethane by a Wurtz coupling reaction of 
triphenylmethyl chloride with metallic silver; but he found that no hydro- 
carbon was formed unless air was carefully excluded from the system. 

benzene 
2 (C 6 H 5 ) 3 C-C1 + 2 Ag ► (C 6 H 5 ) 3 C-C(C 6 H 5 ) 3 + 2AgCl 

hexaphenylethane (C 38 H 3 o) 



sec 24.5 arylmethyl halides 655 

In the presence of atmospheric oxygen, the product is triphenylmethyl 
peroxide, (C 6 H 5 ) 3 COOC(C 6 H 5 ) 3 , rather than hexaphenylethane. In the 
absence of oxygen a compound, C 38 H 30 , assumed to be hexaphenylethane, 
was obtained and shown to dissociate slightly to triphenylmethyl radicals at 
room temperature in inert solvents {K=2.2 x 10 ~ 4 at 24° in benzene). 
However, equilibrium between this compound and triphenylmethyl radicals is 
rapidly established so that oxygen readily converts the ethane into the rela- 
tively stable triphenylmethyl peroxide. 

benzene 24° 

2(C 6 H 5 ) 3 C- 



A = 2.2xlO- 4 

triphenylmethyl radical 

(C 6 H 5 ) 3 + 2 =i (C 6 H 5 ) 3 COO- (C<,H5)3C '» (C 6 H 5 ) 3 COOC(C 6 H 5 )3 

triphenylmethyl 
peroxide 

While these reactions may now seem entirely reasonable, Gomberg's 
suggestion that the triphenylmethyl radical could exist as a fairly stable 
species was not well received at the time. Today, the stability of the radical has 
been established beyond question by a variety of methods such as electron 
paramagnetic resonance (epr) spectroscopy, which is discussed briefly at the 
end of this chapter (Section 24-8). This stability can be attributed to stabili- 
zation of the odd electron by the attached phenyl groups. 

// \—C- < ► •< >=C < ► <f y— C < ► etc. 

^5 ^b ^5 

A curious sequel to Gomberg's work has occurred. The compound 
C 38 H 30 that is in equilibrium with the radical was shown in 1968 to be not 
hexaphenylethane [2] but its structural isomer [3] resulting from coupling of 
two radicals through a para carbon in one of them. Presumably [3] is more 









[2] [3] 

stable than [2] because of lower steric strain. It is interesting that the difference 
is sufficient to overcome the loss of the resonance energy of one benzene ring. 
Some strain undoubtedly remains in [3] and is relieved by dissociation, al- 



chap 24 aromatic side-chain derivatives 656 

though the main driving force for this is the resonance stabilization of the 
radical so formed. 

The stability of a carbon radical, R 3 0, is reflected in the ease with which 
the C— H bond of the corresponding hydrocarbon, R 3 CH, is broken homo- 
lytically. A hydrogen atom bonded to a tertiary carbon is replaced in radical 
chlorination faster than hydrogen at a secondary or primary position (see 
Section 3-3B) showing that the order of stability of the resulting carbon 
radicals is tertiary > secondary > primary. Hydrogen-abstraction reactions 
by radicals other than chlorine atoms have been investigated to obtain some 
measure of hydrocarbon reactivity and radical stability. 



24-6 aromatic aldehydes 



Most of the reactions of aromatic aldehydes involve nothing new or surpris- 
ing in view of our earlier discussion on the reactions of aldehydes (Chapters 1 1 
and 12). One reaction, which is rather different and is usually regarded as 
being characteristic of aromatic aldehydes (although, in fact, it does occur 
with other aldehydes having no a hydrogens), is known as the benzoin con- 
densation. It is essentially a dimerization of two aldehyde molecules through 
the catalytic action of sodium or potassium cyanide. 

HO O 

2 (' V-CHO ► (' ^)—C—C—(' V 

K_7^ n,J C 2 H 5 OH,H 2 \ /| C \ / 

N ' (reflux) N ' jlj x==/ 

benzoin 
90% 

The dimer so formed from benzaldehyde is an a-hydroxy ketone and is 
called benzoin. Unsymmetrical or mixed benzoins may often be obtained in 
good yield from two different aldehydes. 

O OH 

(reflux) jj 

anisaldehyde benzaldehyde 4-methoxybenzoin 

In naming an unsymmetrical benzoin, substituents in the ring attached to 
the carbonyl group are numbered in the usual way while primes are used to 
number substituents in the ring attached to the carbinol carbon. 



Cl 



3-nitro-4'-chlorobenzoin 

The first step in the benzoin condensation involves conversion of the alde- 
hyde to the cyanohydrin by attack of cyanide ion at the carbonyl group. 







O 


OH 


c 


> 


II 


1 
-CH 


>= 


=/ 






NO, 









sec 24.7 natural occurrence and uses of aromatic side-chain derivatives 657 

The cyanohydrin [4] thus formed has a relatively acidic a hydrogen because 
the resulting carbanion is stabilized by both a phenyl and a cyano group. At 
the pH of a cyanide solution, a benzyl-type carbanion [5] is readily formed 
and, in a subsequent slow step, attacks the carbonyl carbon of a second 
aldehyde molecule. Loss of HCN from the addition product [6] leads to 
benzoin. 

O O e OH 

C b H 5 C^ + CN e ^=± C 6 H 5 -C-H ^=± C 6 H 5 -C:e 

H C=N C=N 

[4] [5] 

OH o HO O e 

I \ II 

+ C-C 6 H 5 ► C 6 H 5 -C-C-C 6 H 5 

C=N H NC H 

[6] 



I 



O OH 



C 6 H 5 -C-C-C 6 H 5 -^» C 6 H 5 -C-CH-C 6 H 5 
NC H benzoin 

The unique catalytic effect of cyanide ion is due to its high nucleophilicity 
which leads to the production of an adduct such as [4]; the electron- 
withdrawing power of the cyano group that stabilizes ion [5] ; and the ease 
with which the cyanide ion can be eliminated in the final step. 

The benzoin condensation is a useful synthetic reaction when you want to 
prepare a compound with the Ar-C-C-Ar skeleton. The carbonyl and hydroxyl 
groups in benzoins are subject to the usual reactions of ketones and alcohols. 

24-7 natural occurrence and uses of aromatic 
side-chain derivatives 

Derivatives of aromatic aldehydes occur naturally in the seeds of plants. For 
example, amygdalin is a substance occurring in the seeds of the bitter almond; 
it is a derivative of gentiobiose, which is a disaccharide made up of two glucose 
units. One of the glucose units is bonded through the OH group of benzal- 
dehyde cyanohydrin by a /?-glucoside linkage. 




CHCN 



amygdalin 



chap 24 aromatic side-chain derivatives 658 

The flavoring vanillin occurs naturally as glucovanillin (a glucoside) in the 
vanilla bean, although it is also obtained commercially as a byproduct from 
the treatment of lignin waste liquor (Section 15-7) and by oxidation of 
eugenol, a constituent of several essential oils. 




OCH, 



= CH, 



eugenol 



OH 



3 OH 




OCH, 



CH = CH— CH 

isoeugenol 



OH 




OCH, 



CHO 

vanillin 



H 2 0, H s 



(-CH 3 C0 2 H) 



OCH, 



CH=CHCH, 




CHO 



Methyl salicylate, the major constituent of oil of wintergreen, occurs in 
many plants, but it is also readily prepared synthetically by esterification of 
salicylic acid, which in turn is made from phenol (see Section 23-2D). 



C0 2 H 




C0 2 CH 3 


rV H 


CH 3 OH,H 2 SO„. (| ^f 


Kj 


(-H 2 0) 


\^ 


salicylic acid 




methyl salicylate 
(oil of wintergreen) 



The acetyl derivative of salicylic acid is better known as aspirin and is 
prepared from the acid with acetic anhydride using sulfuric acid as catalyst. 



o 



C0 2 H C0 2 H 

/L X>H /L.OCCH3 

f^f (CH,CO) 2 0, H 2 S0 4 (| ^T 



acetylsalicylic acid 
(aspirin) 



The structures of several other side-chain compounds used as flavorings, 
perfumes, or drugs are shown in Figure 24-2. 

Numerals are used to indicate positions both on an alkane chain and in a 
benzene ring and this can sometimes be awkward, as in the systematic name 
for adrenaline shown in Figure 24-2. With a single substituent on a ring this 
difficulty can be avoided by using the symbol 0, m, or p. 



sec 24.8 electron paramagnetic resonance (epr) spectroscopy 659 



CO,CH, 



NH, 



6 



methyl 2-aminobenzoate 

(methyl anthranilate) 

grape flavoring 

and perfume 



CH 3 

CH 2 -C-NH 2 
I 
H 



l-phenyl-2-aminopropane 
(benzedrine) 

central nervous 

system stimulant; 

decongestant 




CO,C 2 H 5 



NH, 



5-ethyl-5-phenylbarbituric acid 

(phenobarbital) 

sedative 



CHO 



ethyl 4-aminobenzoate 

(benzocaine) 

local anesthetic 




OH 

I 
H-C-CH 2 NHCH 3 



OH 




OH 



1 -(3,4-dihydroxyphenyl)- 

2-methylaminoethanol 

(adrenaline, epinephrine) 

central nervous 

system stimulant; 

blood pressure 

raising principle 

of adrenal glands 



NHCOCH, 



4-ethoxyacetanilide 
(phenacetin) 

analgesic 




CH, 



3,4-methylenedioxybenzaldehyde 
(piperonal) 
perfume ingredient 



OCH, 



/?-(3,4,5-trimethyoxyphenyl)ethylamine 

(mescaline) 

a hallucinogen 



Figure 24-2 Some aromatic side-chain compounds with physiological effects. 



24- 8 electron paramagnetic resonance (epr) 
spectroscopy 

One of the most important methods of studying radicals that has yet been 
developed is electron paramagnetic resonance (epr) or, as it is sometimes 



chap 24 aromatic side-chain derivatives 660 

called, electron-spin resonance (esr) spectroscopy. The principles of this 
form of spectroscopy are in many respects similar to nmr spectroscopy, even 
though the language used is often quite different. The important point is that 
an unpaired electron, like a proton, has a spin and a magnetic moment such 
that it has two possible orientations in a magnetic field corresponding to 
magnetic quantum numbers +| and —\. The two orientations define two 
energy states which differ in energy by about 1000 times the energy difference 
between corresponding states for protons, and therefore the frequency of 
absorption of electrons is about 1000 times that of protons at the same mag- 
netic field. At magnetic fields of 3600 gauss, the absorption frequency of free 
electrons is about 10,000 MHz, which falls in the microwave, rather than the 
radiowave, region. 

The basic apparatus for epr spectroscopy differs from that shown in Figure 
7-11 for nmr spectroscopy by having the sample located in the resonant cavity 
of a microwave generator. The spectrum produced by epr absorption of 
unpaired electrons is similar to that shown in Figure 24-3a, except that epr 
spectrometers are normally so arranged as to yield a plot of the first derivative 
of the curve of absorption against magnetic field, rather than the absorption 
curve itself, as shown in Figure 24-3b. This arrangement is used because it 
gives a better signal-to-noise ratio than a simple plot of absorption against 
magnetic field. 

The sensitivity of epr spectroscopy for detection of radicals is high. Under 
favorable conditions, a concentration of radicals as low as 10 ~ 12 M can be 
readily detected. Identification of simple hydrocarbon free radicals is often 
possible by analysis of the fine structure in their spectra. This fine structure 

Figure 24-3 Plots of absorption (a) and derivative (b) epr curves. 




sec 24.9 linear free-energy relations 661 

arises from spin-spin splittings involving protons, which are reasonably close 
to the centers over which the unpaired electron is distributed. The multiplicity 
of hydrogens and their location in the ortho, meta, and para positions of the 
triphenylmethyl radical produces an extremely complex epr spectrum with at 
least 21 observable absorption lines. Other radicals may give simpler spectra. 
Methyl radicals generated by X-ray bombardment of methyl iodide at 
— 196° show four (n + 1) resonance lines, as expected, for interaction of the 
electron with three (n) protons (see Section 7-6B). 

One of the most exciting uses of epr is in the study of radical intermediates 
in organic reactions. Thus, in the oxidation of hydroquinone in alkaline 
solution by oxygen, the formation of the semiquinone radical (Section 23-4) 
can be detected by epr. The identity of the intermediate is shown by the 
fact that its electron spectrum is split into five equally spaced lines by the 
four equivalent ring protons. The radical disappears by disproportionation 
reactions and has a half-life of about 3 seconds. 

Similar studies have shown that radicals are generated and decay in 
oxidations brought about by enzymes. Radicals have been detected by 
epr measurements in algae "fixing" carbon dioxide in photosynthesis. The 
character of the radicals formed has been found to depend on the wavelength 
of the light supplied for photosynthesis. 



24-9 linear Jree-energy relations 

Can we predict the effect a substituent on a benzene ring will have on the rate 
or the position of equilibrium of a reaction taking place elsewhere in the 
molecule ? Yes, to a very considerable degree, provided the substituent is in a 
meta or para position and certain other information is available. Some idea 
of the regularity of substituent effects can be seen in Figure 24-4, which shows 
a plot of the logarithm of the equilibrium constants for the dissociation of 
a series of benzoic acids against the logarithm of the rate constants for 
the alkaline hydrolysis of the corresponding ethyl benzoates (Equations 24- 1 
and 24-2). 

/^\- C 2 H +=*=> /~\-cO? + H e (24-1) 

z z 



/~\-C0 2 Et + OH e — *— /"VcO? + EtOH (24-2) 

A plot of log k against log Kis really a plot of free energies since log k cc A G % 
and log K oc AG. Thus, we can say that the two reaction series (the meta and 
para compounds, at least) obey a "linear free-energy relation." 

Relations like this are not usually observed with o/t/zo-substituted com- 
pounds (see Figure 24-4) or with aliphatic systems because of steric and other 



chap 24 aromatic side-chain derivatives 662 



60 

o 



-1.0 


- 




o/p-N0 2 

'm-N0 2 


-2.0 


— 


Jbm-CL 
p-Vo/P- CX 


o-N0 2 • 
«o-F 

mo-Cl 


-3.0 




h/ 

p-CH 3 o/ m " CH 3 




-4.0 




Xp-OCH 390 „ch 3 

1 


i 



1.0 2.0 

lOgl0 5 *HA 



Figure 244 Plot of log k for the rates of alkaline hydrolysis of substituted 
ethyl benzoates in 85% ethanol at 30° against log 10 5 K HA for the dissociation 
of substituted benzoic acids in water at 25°. 



Table 24-2 Substituent constants 



substituent 


0" 




substituent 


a 


meta 


para 


meta 


para 


o e 


-0.708 


-1.00 


SH 


+0.25 


+0.15 


NH 2 


-0.161 


-0.660 


CI 


+0.373 


+0.227 


OH 


+0.121 


-0.37 


C0 2 H 


+0.355 


+0.406 


OCH 3 


+0.115 


-0.268 


COCH3 


+0.376 


+0.502 


CH 3 


-0.069 


-0.170 


CF 3 


+0.43 


+0.54 


(CH 3 ) 3 Si 


-0.121 


-0.072 


N0 2 


+0.710 


+0.778 


C 6 H 5 


+0.06 


-0.01 


N(CH 3 ) 3 


+0.88 


+0.82 


H 


0.000 


0.000 


S(CH 3 ) 2 


+ 1.00 


+0.90 


SCH 3 


+0.15 


0.00 


N 2 ® 


+ 1.76 


+ 1.91 


F 


+0.337 


+0.062 









sec 24.9 linear free-energy relations 663 

proximity effects. Meta- and para-substituted series can be related this way 
because the substituents are far enough away from the reaction sites so that 
steric and proximity effects are diminished and only the electrical effect of the 
substituent is important. A p-nitro group is always electron withdrawing and 
a p-amino group always electron donating and we can assign to these and 
other substituents a value that represents their electron-donating or electron- 
withdrawing ability relative to hydrogen taken as zero. These substituent 
constants, symbol <x, are obtained from the ionization constants of substituted 
benzoic acids. If a substituent increases the acidity of benzoic acid, a is 
positive; if it decreases the acidity of benzoic acid, a is negative. The larger 
the effect of the substituent on the benzoic acid ionization, the larger is the 
absolute value of a. Table 24-2 lists the substituent constants for a number of 
common groups. 

To predict the effect of a substituent on a given rate or equilibrium con- 
stant, a further factor must be considered — the sensitivity of the reaction site 
to electrical effects of the substituents. The slope of the line in Figure 24-4 is 
+ 2.2. The positive sign means that the reaction sites in both ester hydrolysis 
and acid ionization have the same kind of response to substituent electrical 
effects. Thus, p-mtxo speeds up the hydrolysis of the ester and also increases 
the degree of dissociation of the acid. For the ester hydrolysis, the electron- 
withdrawing group helps to make the carbonyl carbon more positive and 
hence better able to attract hydroxide ion (Section 13-8); in ionization, it 
helps to stabilize the anion by electron attraction. Since the slope is greater 
than unity (2.2) the ester hydrolysis is more sensitive to the effects of substi- 
tuents than is the acid dissociation. 

The sensitivity of a reaction to substituents is given by the reaction constant, 
symbol p, and is obtained by measuring the slope of the line when log k or 
log K is plotted against log K HA for benzoic acids. If, as for the benzoic acid 
dissociation, the reaction is facilitated by electron-withdrawing groups, p is 
positive. If the reaction is more sensitive than the benzoic acid dissociation 
(slope > 1), then p is greater than unity; if it is less sensitive, p is less than 
unity. A reaction with the opposite electronic requirements has a negative p. 

Values of p are obtained by plots such as that in Figure 24-4, and a number 
of these reaction constants are listed in Table 24-3. 

The combination of the two independent variables p and a defines the 
effect of the meta or para substituent on the rate constant — that is, the dif- 
ference between log k and log k , where k refers to the unsubstituted 
compound. 

log k = pa + log k 

This equation (or its analog for equilibrium constants) is known as the 
Hammett equation. It describes in a quantitative way the effect of substi- 
tuents on reaction rates (or equilibrium constants) oimeta ox para substituted 
aromatic compounds. 



chap 24 aromatic side-chain derivatives 664 

Table 24-3 Reaction constants 



equilibria 



ff\ H 2 Q // \ 

((, V-CO,H . </ >-co 2 e + H® 1.00 

(, V-CH 2 C0 2 H . H2 ° ' /, \-CH 2 C0 2 e + H® 

^— * 95% C 2 H 5 OH /— \ 

<£- VCHO + HCN ■ J/J)-CH(OH)CN 



/~V nu -J^L r\ 



0.489 



1.492 



OH ^= 6 Vo e + H® 2.113- 

25 ° R' 



reaction rates 



/7^S-C0 2 C 2 H 5 + OH e CzH ^° H - /TA-C0 2 e+ C 2 H 5 OH 2.431 

/TS-CH^A + C 2 H 5 OH — ^- J^V? 1 *"^!^ + HC1 ~ 5 '° 90 * 

R" X==/ a _ R _ OC 2 H 5 ~~ 

R £>- COCH > + ** "^r R Jc> COCH2Br + HBr 

(7S + NO/ «»§>£ r ^-N0 2 + H* 



" Nitro and similar groups have somewhat exalted orj>«™ values in reactions such as this in which 
direct resonance interaction is possible between the anionic reaction site and the para position. 

* Amino and similar groups have somewhat exalted opara values in reactions such as this in 
which direct resonance interaction is possible between the cationic reaction site (in the transition 
state in this case) and the para position. 



summary 665 

summary 

Aromatic side chains are joined by carbon-carbon bonds to the aromatic 
ring. Such compounds have reactions that are very like those of their aliphatic 
analogs, but their methods of preparation are usually different. Vigorous 
oxidation of any of these compounds (including those with longer side 
chains) produces benzoic acid, [l]-[8] -> [9], although the presence of hydroxyl 
or amino substituents on the ring causes the ring to be degraded. 

O 

II 
ArCH 3 ArCH 2 X ArCHX 2 ArCX 3 ArCH 2 OH ArCHO ArCCH 3 ArC=N ArC0 2 H 

[1] [2] [3] [4] [5] [6] [7] [8] [9] 

X = halogen 

Side-chain halogenation occurs readily, [1] -» [2] + [3] + [4], and the 
resulting halo compounds can be readily hydrolyzed to [5], [6], and [9]. 
Friedel-Crafts methylation of arenes gives [1] (RX gives ArR) but poly- 

O 

II 
substitution also occurs; Friedel-Crafts acylatioh by CH 3 CC1 gives [7] 

O O 

II II 

(RCC1 gives ArCR) and only monosubstitution occurs. Introduction of the 

formyl group, — CHO, is only successful with activated rings. Aromatic 

O 

II 
aldehydes undergo the benzoin condensation to give benzoins, ArCHOHCAr. 

Multiple substitution of aryl rings on one carbon atom allows carbonium 

ions, carbanions, and radicals to be formed. 



Ar 3 C e Z = OH, Y = H® 

Ar 3 C-Z+Y ► Ar 3 C e Z = H, Y = NHf 

Ar 3 C- Z = CI, Y = Ag 



Steric and electronic effects are important in stabilizing these three enti- 
ties. All absorb light in the visible region of the spectrum. Radicals, in addi- 
tion, can be detected and identified by epr spectroscopy. 

The reactions of meta- and /?ora-substituted aromatic systems obey linear 
free-energy relationships, one of which is the Hammett equation, log k = 
pa + log k . This expresses a rate constant for the reaction of a meta- or 
/?«ra-substituted aryl compound in terms of k (the rate constant for the 
rate of the unsubstituted compound), p (the reaction constant, which is 
independent of the substituent), and a (the substituent constant, which is 
independent of the reaction). A similar relationship holds for equilibrium 
constants. 



chap 24 aromatic side-chain derivatives 666 

exercises 

24-1 Suggest a practical synthesis of each of the following compounds from a 
readily available aromatic hydrocarbon : 

/^-\ /^ /^ .C0 2 H 

a. H 2 N-f )-C0 2 H 



b. H0 2 C-f VC0 2 H 





CO,H 



24-2 Write a mechanism for the formation of benzyl chloride by photochemical 
chlorination of toluene with molecular chlorine. What other products would 
you anticipate being formed? At what position would you expect ethyl- 
benzene to substitute under similar conditions ? 

24-3 Outline a suitable synthesis of each of the following compounds, starting 
with benzene : 



a. C 6 H 5 CH 2 COC 6 H 5 

b. C 6 H 5 CH 2 CONHCH 2 CH 2 -4 />~ N0 2 




c. ClHk J^-CHO 

24-4 Suggest a reason why zinc chloride is used in preference to aluminum chloride 
as a catalyst for chloromethylation reactions. 

24-5 Give the principal product(s) of chloromethylation of the following 
compounds : 

a. 1-methylnaphthalene c. />-methoxybenzaldehyde 

b. 1-nitronaphthalene d. anisole (using acetaldehyde in 

place of formaldehyde) 

24-6 Suggest a possible mechanism for formylation of arenes by carbon monoxide 
(Section 24 -4C). 

24-7 Formulate the steps that are probably involved in the formylation of a phenol 
by the action of HCN, HC1, and ZnCl 2 . 

24-8 How would you synthesize the following compounds from the indicated 
starting materials ? 



«■ h 3 c y^/ 



OH 

CHO 

from toluene 



exercises 



667 



b. CH 3 CH 2 



/~\ 




CH 2 OH from benzene 



CI 
c. CH 3 CH 2 0— i \-CHO from benzene 

24-9 Write resonance structures for O-NHaCeH^C®. Would meta amino groups 
be as effective in stabilizing the ion ? 

24-10 a. Suggest why the extent of ionic dissociation of triarylmethyl chlorides 
in liquid sulfur dioxide decreases for compounds [1], [2], and [3] in the 
order [1] > [2] > [3]. Use of models may be helpful here. 



(C 6 H 5 ) 3 C-C1 
[1] 




CI 

I 
C(C 6 H 5 ) 2 



C1-C(C 6 H 5 ) 2 




b. Which alcohol would you expect to give the more stable carbonium ion 
in sulfuric acid, 9-fluorenol [4] or 2,3,6,7-dibenzotropyl alcohol [5]? 
Explain. 



OH 





c. When triphenylcarbinol is dissolved in 100% sulfuric acid, it gives a 
freezing-point depression that corresponds to formation of 4 moles of 
particles per mole of carbinol. Explain. 

24-11 . Write the important resonance structures for the triphenylmethide ion and 
for the carbanion formed by proton loss from 4-cyanophenyldiphenyl- 
methane. 



24-12 Which of the following pairs of compounds would you expect to be the more 
reactive under the specified conditions ? Give your reasons and write equa- 
tions for the reactions involved. 

a. p-N0 2 C 6 H 4 CH 2 Br orp-CH 3 OC6H 4 CH 2 Br on hydrolysis in aqueous 
acetone 

b. (C 6 H 5 ) 3 CH or C 6 H 5 CH 3 in the presence of phenyllithium 

c. (C 6 H 5 ) 3 C-C(C 6 H 5 ) 3 or (C 6 H 5 )2CH-CH(C 6 H 5 ) 2 on heating 

d. (C 6 H 5 ) 2 N-N(C6H 5 ) 2 or (C 6 H 5 ) 2 CH-CH(C 6 H 5 ) 2 on heating 



chap 24 aromatic side-chain derivatives 668 




Figure 24-5 Electron paramagnetic resonance spectrum of cycloheptatrienyl 
radical produced by X-irradiation of 1,3,5-cycloheptatriene. 



24-13 Draw structures and name all the possible benzoins that could be formed 
from a mixture of (a) />-tolualdehyde and o-ethoxybenzaldehyde, and (b) 
8-methyl-l-naphthaldehyde and anisaldehyde. An unsymmetrical benzoin 
such as 4-methoxybenzoin is rather readily equilibrated with its isomer, 
4'-methoxybenzoin, under the influence of bases. Explain. 

24-14 The epr spectrum shown in Figure 24-5 is of a first-derivative curve of the 
absorption of a radical produced by X-irradiation of 1,3,5-cycloheptatriene 
present as an impurity in crystals of naphthalene. Make a sketch of this 
spectrum as it would look as an absorption spectrum and show the structure 
of the radical to which it corresponds. Show how at least one isomeric 
structure for the radical can be eliminated by the observed character of the 
spectrum. 

24-15 The ionization constants of m- and p-cyanobenzoic acids at 20° are 2.51 x 
10 ~ 4 and 2.82 x 10 ~ 4 , respectively. Benzoic acid has K HA of 6.76 X 10~ 5 at 
20°. Calculate a meta and a para for the cyano substituent. 



24-16 The effects produced by substituents are explained in terms of inductive, 
conjugative (resonance), and steric influences. Show how it is possible, 
within this framework, to account for the following facts : 

a. The a constant of the methoxy group (— OCH 3 ) in the meta position 

is positive and in the para position negative. 
© 

b. The — N(CH 3 ) 3 group has a larger positive a constant in the meta 
position than in the para position, but the reverse is true for the — N 2 e 
group. 



chap 25 heterocyclic compounds 671 

Heterocyclic organic compounds have cyclic structures in which one or more 
of the ring atoms are elements other than carbon. In this chapter we shall 
confine our attention to a discussion of the chemistry of heterocyclic nitro- 
gen, oxygen, and sulfur compounds, and of these we shall be concerned 
primarily with the aromatic heterocycles rather than their saturated analogs. 
The chemistry of saturated heterocycles, such as ethylene oxide and the other 
compounds shown in Figure 25-1, has been dealt with in earlier chapters. In 
general, the properties of such substances can be correlated with those of their 
open-chain analogs, provided appropriate account is taken of the strain and 
conformational effects that are associated with ring compounds. 

The importance of heterocyclic compounds is apparent from the wealth 
and variety of such compounds that occur naturally or are prepared on a 
commercial scale by the dye and drug industries. Many of these compounds 
fulfill important physiological functions in plants and animals. We have 
already encountered some of the important naturally occurring heterocycles 
in earlier chapters. Thus, the carbohydrates may be classified as oxygen 
heterocycles, whereas the nucleic acids and some amino acids, peptides, and 
proteins possess nitrogen-containing ring systems. 

We shall begin with a discussion of four important unsaturated hetero- 
cyclic compounds: pyrrole, furan, thiophene, and pyridine (Figure 25-2; 
their systematic names are shown in parentheses). The prefixes az-, ox-, and 
thi- refer, respectively, to nitrogen, oxygen, and sulfur heterocycles. Suffixes 
indicate the size of the ring, for example, -ole for five-membered unsaturated 
rings, and -ine for six-membered unsaturated rings. 

These four compounds are all liquid at room temperature; their boiling 
points range from 32° for furan to 130° for pyrrole. The fairly high boiling 
point of the latter is presumably the result of intermolecular hydrogen 
bonding. 

The conjugated bonding in pyridine bears an obvious resemblance to that 
in benzene and it is no surprise to find that pyridine is an " aromatic" com- 
pound; for instance, it reacts with electrophiles by substitution rather than by 
addition. It is less obvious that pyrrole, furan, and thiophene should have 
aromatic character and yet we shall see that these compounds, too, resemble 
benzene in many ways. 



Figure 25- 1 Some important saturated heterocycles. 









/ 


/> 


H 2 C-CH 2 

V 


Q 





o 


O h 


ethylene 
oxide 


tetrahydrofuran 


1,4-dioxane 


lactones 


lactams 



chap 25 heterocyclic compounds 672 



HC-CH g—z HC-CH 

// w 

HC .. CH 



H O 



H H 

pyrrole furan 

(azole) (oxole) 



HC-CH r -» HC^ ^CH ^ 

HC^CH V HC^CH . N 

thiophene pyridine 

(thiole) (azine) 



Figure 25-2 The structures of four important unsaturated heterocycles. 

25' 1 aromatic character of pyrrole, Jur an, and 
thiophene 

The five-membered heterocyclic compounds, pyrrole, furan, and thiophene, 
possess some degree of aromatic character because of the derealization of 
four carbon % electrons and the two unshared electrons of the heteroatom. 
This combination constitutes a sextet of delocalized electrons. We learned 
earlier (Section 6-7) that cyclic systems with this electronic arrangement tend 
to have special stabilization — for example, benzene, cyclopentadienide anion, 
and cycloheptatrienyl (tropylium) cation. 

The structure of each heterocycle can be described as a hybrid of several 
electron-pairing schemes, as shown here for pyrrole. We shall have more to 
say later about the degree of contribution of the dipolar resonance forms, 
[lb] to [le]. 



H 

[la] 


H e 
[lb] 


[lc] 
or 

86,^^8 6 

8e4 .^8e 
N 

j>8a> 


H ffl 
[Id] 


H® 

[le] 



[1] 



In terms of atomic orbitals, the structure of each of these heterocycles may 
be regarded as a planar pentagonal framework of C— H, C— C, and C— Y a 
bonds (Y being the heteroatom) made up of trigonally bonded (sp 2 ) atoms, 
each with one p orbital perpendicular to the plane of the ring. The % system 
formed by overlap of the p orbitals perpendicular to the plane contains four 



sec 2S.2 chemical properties of pyrrole, furan, thiophene, and pyridine 673 





C— C a boncH 




,' > " ? ~^^ / 




,,'''n -*^ >*"/ H^^-/'"' — H a bond-? 




r f~-^ ^\- y L^-r^ / 


,-*"* ~~^^^'- > <^/ ^ / 


H- 


-<T. /- ^0^» 




/: .^i ; fe ,^V W'ff ^x 












v'V- "_* \ V, ■■■■-■; 




<. 1 ■•■ \ >-'• 




N I \ -"" 




1 \ / 




s ■■■■. \ / 




1 V 




\ -■:" -'■■' ; : x' V C— N a bond 




m ••: ,y \_ ff bonding 




#" 




H 



Figure 25-3 Atomic orbital description of pyrrole. 



electrons from the carbons and two from the heteroatom. The overall 
formulation is illustrated in Figure 25-3 for pyrrole. 

The stabilization energies of pyrrole, furan, and thiophene obtained from 
experimental and calculated heats of combustion are only about half of the 
stabilization energy of benzene. However, the heterocycles differ from benzene 
in that each has only one resonance structure with no formal bonds or charge 
separation. Furthermore, on the basis of relative electronegativities of sulfur, 
nitrogen, and oxygen, we may anticipate that the structures analogous to 
[lb] to [le] should be important in the order thiophene > pyrrole > furan. 
As a result, it is reasonable to expect furan to be the least aromatic of the 
three heterocycles, and indeed it is. 



25-2 chemical properties of pyrrole, furan, thiophene, 
and pyridine 

In discussing the reactivity of these heterocycles, we shall be interested pri- 
marily in their degree of aromatic character, as typified by their ability to 
exhibit electrophilic and nucleophilic substitution reactions rather than under- 
go addition reactions. First, however, we shall consider their acidic and basic 
properties. 

A. ACIDIC AND BASIC PROPERTIES 

Pyrrole, furan, thiophene, and pyridine are potential bases because each can 
accept a proton at the heteroatom. Thiophene and furan, however, are too 
weak to form salts with aqueous acid. Pyrrole and pyridine are somewhat 
stronger bases, as might be expected from the lower electronegativity of 
nitrogen. Pyrrole, however, polymerizes in acid solution (as does furan). 



chap 25 heterocyclic compounds 674 

Pyridine (K B = 1.1 x 1CT 9 ) is thus the only one of these heterocycles which 
forms stable salts with aqueous acid. 



HBr 

^ H2 ° : "X o 
H®Br e 

pyridine pyridinium bromide 

Turning to the acidic properties of these four compounds, a glance at their 
structures tells us that pyrrole is the only one that is likely to be acidic because 
it is the only one in which there is a hydrogen attached to the heteroatom. 
Pyrrole is a rather weak acid (X HA = 10~ 15 ) and reacts completely only with 
strong bases such as hydroxide ion or Grignard reagents. 




0+H 2 
pyrrylpotassium 



+ ch * 

MgBr 
pyrrylmagnesium bromide 

Although pyrrole is only weakly acidic it is a stronger acid than aliphatic 
amines by a factor of about 10 18 (see Section 16-1D) and stronger than 
aromatic amines by a factor of about 10 12 (see Table 22T, footnote). This 
reflects the stability of the 7i-electron system of the resultant anion [2] relative 
to that of pyrrole itself [1] where charge separation is associated with all but 
one of the resonance structures. 



o — o — o — O ,o — „ 

N N N N N 

e 

[2a] [2b] [2c] [2d] [2e] 

The resonance structures [2a]-[2e] are useful for indicating charge dispersal 
in the anion but do not reveal the special stabilization that is associated with 
derealization of six n electrons. 



B. ELECTROPHILIC SUBSTITUTION REACTIONS 

Electrophilic substitution of pyrrole is rapid at the 2 position, although if this 
site (and the 5 position) is blocked, the 3 position is attacked readily. As with 
electrophilic substitution in benzene (Figure 20-8) a convenient model for the 
transition state is the intermediate cation formed by addition of an electro- 
phile X® to the pyrrole ring. 
The stability that the heteroatom confers on the intermediate and, by 



sec 25.2 chemical properties of pyrrole, furan, thiophene, and pyridine 675 

analogy, on the transition state for substitution is illustrated in [3a]-[3c] for 
the case of 2 substitution. 



H 



H X 

[3a] 



H X 

[3b] 



"H X 

[3c] 



The positive charge is located on nitrogen in [3c] and this is the most 
important contributor to the resonance hybrid, even though nitrogen is more 
electronegative than carbon. It is important to recognize the reason for this, 
which is simply that there is more bonding in [3c] than in [3a] or [3b] — two n 
bonds instead of one. 

The analogous intermediates [4] or [5] that result from electrophilic attack 
on benzene or pyridine do not have structures in which there are different 
numbers of bonds. 



+ z a 







+ jy- 





z x H 


Z H 


z x H 




e ^S . 


- A e - 


~6 




k) 


u 








« 




[4a] 


[4b] 


[4c] 




Z H 


z x H 


Z H 




e fil ■ 


^Av 


^0 




u 


u 




[5a] 


[5b] 


[5c] 



Although the nitrogen atoms in [3c] and [5c] are each cationic, that in [3c] 
has an octet and that in [5c] only a sextet of electrons. Because nitrogen is more 
electronegative than carbon, [5c] will make only a small contribution to the 
resonance hybrid and, for this reason, electrophilic substitution in pyridine 
occurs at the 3 position (although very slowly). 

The principal electrophilic substitution reactions of pyrrole are summarized 
in Figure 25-4 and a study of these reactions provides an opportunity to 
review the reactions discussed earlier in our study of benzene. Note that 
2 substitution predominates; nitration and sulfonation of pyrrole are pos- 
sible, but only if strongly acidic conditions which would lead to polymerization 
are avoided ; and pyrrole is sufficiently reactive for halogenation and Friedel- 
Crafts acylation to proceed without a catalyst. 

Furan resembles pyrrole in its behavior toward electrophilic reagents, and 
its principal reactions of this type are summarized in Figure 25-5. Direct 
chlorination and bromination of furan are hard to control and can lead to 
violent reaction, possibly caused by the halogen acid that is formed. 

Related reactions of thiophene are summarized in Figure 25-6. Because 
thiophene is less subject to acid-induced polymerization than either pyrrole or 
furan, it can be sulfonated or nitrated under strongly acidic conditions. In 
fact, sulfonation and extraction are used as a means of freeing commercial 



chap 25 heterocyclic compounds 676 



CH 3 C0 2 N0 2 



(CH 3 CO) 2 0, 5° 



S0 3 , pyridine 
90° 



2 N-/ \-N 2 CI 



1. HCN, HC1 

2. H 2 



(CH 3 CO) 2 Q 

250° 



Br 2 , C 2 H 5 OH 



"N N0 2 



^ SO3H 



H 



N N=NC 6 H 4 N0 2 



^W CHO 
H 



X N C 



H 



O 



Br Br 

Br N Br 

H 



nitration 



sulfonation 



diazo coupling 



formylation 



Friedel-Crafts 
acylation 



bromination 



Figure 25-4 Electrophilic substitution reactions of pyrrole. 



Figure 25-5 Electrophilic substitution reactions of furan. 



CH 3 C0 2 NO 



3 «~u 2 r«_> 2 



(CH 3 CO) 2 0, chilled 



- n 



-o 



■NO, 



so 3 



o 

N o^ 



pyridine 

Cl-f VN 2 e Cl e 



(1 
N Q ,>~SO,H 



nitration 



sulfonation 



-» // \ diazo coupling 

< n X~-N = NC 6 H 4 C1 



(CH 3 CO) 2 



BF 3 



1. HCN, HC1 

2. H 2 



V >~COCH 3 



Friedel-Crafts 
acylation 



formylation 



sec 25.2 chemical properties of pyrrole, furan, thiophene, and pyridine 677 



H 2 SO» 




^ s /~~S0 3 H 






benzene Br '^ o/^Br 



r\ 



sulfonation 



nitration 



bromination 



iodination 



Friedel-Crafts 
CC 6 H 4 C0 2 H acylation 







<. C A-CH,C1 



chloromethylation 



Figure 25-6 Electrophilic substitution reactions of thiophene. 



Figure 25-7 Electrophilic substitution reactions of pyridine. In most of these 
reactions the yields are low. 



H 2 S0 4 




Br 2 



2O0°-3O0° 



Br, 



300°-500° 



KN0 3 , H 2 S0 4 

370° 



NO, 



H 



S0 3 , H 2 S0 4 , HgS0 4 

220°, 24 hr 



SO,H 



Br Br^^^Br 



a + xj 



(probably a 
thermally 
induced 
"N^^Br Br" ^f>r"Br radical 

substitution) 



chap 25 heterocyclic compounds 678 

benzene from thiophene, with which it is often contaminated. Thiophene is 
sulfonated much more readily than benzene and the resulting product, being 
acidic, can be extracted by aqueous base. 

Electrophilic substitution of pyridine is hard to achieve, partly because of 
deactivation of the ring by the heteroatom and partly because under acidic 
conditions, as in sulfonation and nitration, the ring is further deactivated by 
formation of the pyridinium ion. Three pertinent substitution reactions are 
listed in Figure 25-7; their most striking feature is the vigorous conditions 
necessary for successful reaction. Note that the Friedel-Crafts reaction does 
not take place with pyridine. 



C. NUCLEOPHILIC SUBSTITUTION REACTIONS 

The most important substitution reactions of the pyridine ring are effected by 
nucleophilic reagents. Thus, pyridine can be animated on heating with soda- 
mide, hydroxylated with potassium hydroxide, and alkylated and arylated 
with alkyl- and aryllithiums (Figure 25-8). Since related reactions with 
benzene either do not occur or are relatively difficult, we can conclude that 
the ring nitrogen in pyridine has a pronounced activating effect for nucleo- 
philic attack at the ring analogous to the activation produced by the nitro 
group in nitrobenzenes. The reason for this activation is that addition to the 



Figure 25-8 Nucleophilic 


substitution reactions of pyridine. 




1. NaNHj, 100° 


► + H 2 (Tschitchibabin reaction) 

^N^-NH 2 


2. H 2 


/ 




2-aminopyridine 


// 


KOH 


, n >f\ 


320° [O] 


H 




\\ 


C 6 H s Li 

110°, toluene 


2-pyridinol 2-pyridone 

if] + LiH 
k N^-c 6 H 5 
2-phenylpyridine 


v 


«-C 4 H 9 Li 


f] +LiH 
^N^c 4 H 9 


100° 






2-butylpyridine 



sec 25.3 polycyclic and polyhetero systems 679 

2 or 4 but not the 3 position permits the charge to reside at least partially on 
nitrogen rather than on carbon (see Equations 25-1 and 25-2). 



:NH, 



favorable 



*N- 



1ST 



, unfavorable 



H 
NH, 



H 



H 



N NH, 



H 



: N /-e 



V N 



H 

NH, 



H 
NH, 



(25-1) 



"N 



(25-2) 



In the case of amination, the reaction is completed by loss of hydride ion and 
subsequent formation of hydrogen (Equation 25-3). 




e 

(Th 



' N NH, 



+ :H 



*N- 



NH, 



(25-3) 



+ H, 



^N 



NH 



H a> 



N NH, 



25-3 polycyclic and polyhetero systems 

A number of heterocycles that have fused benzene rings attached are shown 
in Figure 25-9. As might be expected, electrophilic substitution occurs in the 



Figure 25-9 Some important polycyclic and polyhetero ring systems. 




N' 
H 

indole 
(benzopyrrole) 




benzofuran 




quinoline 




%>^^N: 

isoquinoline 






imidazole 



thiazole pyrimidine 



H 

purine 



pteridine 



chap 25 heterocyclic compounds 680 

hetero ring in the case of indole and benzofuran but in the benzenoid ring 
in the case of quinoline and isoquinoline. 

There are several five- and six-membered heterocyclic ring systems con- 
taining two or more heteroatoms within the ring that are particularly impor- 
tant in that they occur in many natural products and in certain synthetic 
drugs and synthetic dyes. The parent compounds of the most commonly 
encountered ring systems are shown in Figure 25-9. 

Each five- and six-membered ring compound has a delocalized sextet of n 
electrons; the bicyclic compounds, purine and pteridine, resemble naph- 
thalene in having 10 delocalized n electrons. 



heterocyclic natural products 

Some of the heterocyclic ring systems mentioned in this chapter are of special 
interest and importance because certain of their derivatives are synthesized 
naturally as part of the life cycles of plants and animals. The structures of 
these naturally occurring compounds are often extremely complex and eluci- 
dation of their structures has been and continues to be a major challenge to 
organic chemists and biochemists alike. The approach to solving the struc- 
ture of a complex natural product is discussed in some detail in Chapter 29 ; 
at this point we shall briefly describe only a few natural products of known 
structure which can be classified as heterocyclic compounds and which are of 
some biological or physiological importance. 



25-4 natural products related to pyrrole 

An interesting compound having a fully conjugated cyclic structure of four 
pyrrole rings linked together through their 2 and 5 positions by four methine 
(=CH— ) bridges is known as porphyrin [6]. 




Although porphyrin itself does not exist in nature, the porphyrin or related 
ring system is found in several very important natural products, notably 
hemoglobin, chlorophyll, and vitamin B 12 . 

Hemoglobin is present in the red corpuscles of blood and functions to carry 
oxygen from the lungs to the body tissue ; it consists of a protein called globin 
bound to an iron-containing prosthetic group called heme. Acid hydrolysis 
of hemoglobin liberates the prosthetic group as a complex iron(III) salt called 



sec 25.4 natural products related to pyrrole 681 

hemin [7]. The structure of hemin was established by 1929 after years of work, 
notably by W. Kiister, R. Willstatter, and H. Fischer. A complete synthesis of 
hemin was achieved by Fischer in 1929, and his contributions were rewarded 
with a Nobel Prize (1930). The structure of hemin [6] shows that the iron 
(as Fe m ) is complexed to all four of the pyrrole nitrogens. 

The poisonous action of carbon monoxide is due to its reaction with 
hemoglobin to form carboxyhemoglobin. Carbon monoxide makes up about 
4% of undiluted cigarette smoke and can produce up to 15% carboxyhemo- 
globin in the blood. 



CH=CH 2 CH 3 
H,C— < 1 "I ^CH=CH, 




CH, 



CH 2 
CH 2 
CO,H 



CH 2 

CH 2 
I 
CO,H 



Cl fe 



hemin 

[7] 

Certain pigments in the bile of mammals, the so-called bile pigments, are 
pyrrole derivatives. They contain four pyrrole rings linked in a chain through 
a methine bridge between the 2 position of one ring and the 5 position of an- 
other. As one might suspect, bile pigments are degradation products of he- 
moglobin. 




basic structure of a bile pigment 



Chlorophyll was briefly mentioned in connection with photosynthesis, and 
its structure is shown in Figure 15T. It is a porphyrin derivative in which the 
four pyrrole nitrogens are complexed with magnesium (as Mg u ). The struc- 
ture was established largely through the work of R. Willstatter, H. Fischer, 
and J. B. Conant. A total synthesis was completed by R. B. Woodward and 
co-workers in 1960. 

The structure of vitamin B 12 [8], known also as the antipernicious anemia 
factor and as cyanocobalamin, was finally established in 1955 as the result 
of both X-ray diffraction and chemical studies. The vitamin has a reduced 
porphyrin ring in which one methine bridge is absent and the nitrogen 
heteroatoms are complexed with a cyanocobalt group. It also has a ribo- 
furanoside ring and a benzimidazole ring. Intensive efforts have been under- 
way for some time to achieve a synthesis of vitamin B 12 in the laboratory. 



chap 25 heterocyclic compounds 682 



o 

II 

H 2 NCCH 2 
HjC'N 

O 

|| J 


CH 2 CH 2 CONH 2 
| CH 3 

| I^^C cH 2 CONH 2 

/ ^y\^\-- CH 2 CH,CONH 2 

" CN > / " 

^N | NJ 

Co® \ 
^N ! N-A CH 3 


HjNCchAJJ JL a 

IN-CCH.CH^/T ^H 2 CH 2 CONH 2 

1 ° / 3 
CH 2 


CHCH 3 

o^ x o HO 


• , "\^CH 3 


H 2 C ^o^ 




OH 


vitamin B 12 




[8] 



25- 5 natural products related to indole 

The indole ring system is common to many naturally occurring compounds, for 
example, the essential amino acid tryptophan which is a constituent of almost 
all proteins. 




CH 2 CHC0 2 H 

I 

\ N " 2 



tryptophan 

There are also many compounds related to indole that occur in plants. They 
are part of a class of natural products known as alkaloids — the term being 
used to designate nitrogen-containing compounds of vegetable origin com- 
monly having heterocyclic ring systems and one or more basic nitrogen atoms. 
Their physiological activity is often pronounced and their structures complex. 
Alkaloids related to indole are called indole alkaloids, and some of these are 
described here. 

An indole derivative commonly known as serotonin, which is actually 
5-hydroxytryptamine, is of interest because of its apparent connection with 
mental processes. It occurs widely in plant and animal life, but its presence in 
the brain and the schizophrenic state that ensues when its normal concen- 



sec 25. S natural products related to indole 683 

tration is disturbed indicates that it may have an important function in 
establishing a stable pattern of mental activity. 



HO 




serotonin 

The ergot alkaloids are produced by a fungus known as ergot, which 
grows as a parasite on cereals, particularly rye. They are amides of the indole 
derivative known as lysergic acid and their levorotatory forms are physiolog- 
ically active in minute amounts. Ergot poisoning, or St. Anthony's fire, has 
been known for centuries and still occurs occasionally. Several deaths follow- 
ing fits of madness were reported in a village in France a few years ago and 
were traced to bread baked with ergot-containing rye. 




lysergic acid 

The diethylamide of lysergic acid, while not itself a naturally occurring 
compound, has achieved notoriety as a drug (LSD) that can produce a 
temporary schizophrenic state, although permanent damage to the brain can 
also result. Current theory suggests that the diethylamide of lysergic acid 
upsets the balance of serotonin in the brain. 

Another indole alkaloid called reserpine has important clinical use in the 
treatment of high blood pressure (hypertension) and also as a tranquilizer for 
the emotionally disturbed. The tranquilizing action is thought to be the 
result of a reduction in the concentration of brain serotonin. 




Two other alkaloids, strychnine and brucine, have been discussed (Section 
14-6). The problem of elucidating their structures was solved only after more 
than a century of research, the major contributions in recent years being made 
by R. Robinson and R. B. Woodward. 



chap 25 heterocyclic compounds 684 

25-6 natural products related to pyridine, 
quinoline, and isoquinoline 

Among the natural products related to pyridine, we have already mentioned 
the coenzyme nicotinamide-adenine dinucleotide (NAD®, Figure 18-3). Other 
important pyridine derivatives include nicotine, nicotinic acid (niacin, anti- 
pellagra factor), and pyridoxine (vitamin B 6 ). Coniine is a toxic alkaloid 
which occurs in the shrub poison hemlock; it has a reduced pyridine (piperi- 
dine) ring. 



, — . 




CH 2 OH 




(l N 

^*T CH 3 


^yC0 2 H 


ho^^ch 2 oh 

H 3 C^N^ 


^N^CH 2 CH 2 CH 3 

H 2 2 3 


nicotine 
(from tobacco) 


nicotinic acid 
(niacin) 


pyridoxine 


coniine 

(poisonous component 

of hemlock) 



A group of related and rather poisonous compounds known as the tropane 
alkaloids are derivatives of reduced pyrroles and reduced pyridines. Two of 
the more important tropanes are atropine and cocaine. 



CH, 



II 



CH, 



N N CO z CH 



atropine 

(from Atropa belladonna plant, 

" deadly nightshade " 



( — )-cocaine 
(from coca plant) 



The cinchona alkaloids are quinoline derivatives which occur in cinchona 
bark and have medicinal value as antimalarials. The most notable example is 
quinine, the structure of which is shown in Section 14-6. 

Many alkaloids have isoquinoline and reduced isoquinoline ring systems. 
The opium alkaloids are prime examples, and include the compounds nar- 
cotine, papaverine, morphine, codeine, and several others, all of which occur 
in the seed of the opium poppy. 



CH,0 




JO- 



H,C 



O 




N-CH, 



CH 



-O 

I 



OCH, 



6; c=0 

S^OCH 3 




N-CH, 



HO 



OCH 3 

papaverine 



OCH, 



narcotine 



morphine, R = H 
codeine, R = CH, 



sec 25.7 natural products related to pyrimidine 68S 

25-7 natural products related to pyrimidine 

The pyrimidine ring system occurs in thymine, cytosine, and uracil, which are 
component structures of the nucleic acids and certain coenzymes. A detailed 
account of the structures of nucleic acids is given in Section 17-7. Thiamine 
[9] is both a pyrimidine and a 1,3-thiazole derivative. The pyrophosphate of 
thiamine is the coenzyme of carboxylase — the enzyme that decarboxylates 
a-ketoacids ; thiamine is also known as vitamin B 1( and a deficiency of it in the 
diet is responsible for the disease known as beri-beri. 



NH 2 CH, 



CH 3 -( VCH 2 -N 



H 

OH OH 

thiamine, R = H | | 

thiamine pyrophosphate, R = — P — O — P — OH 

II II 

O O 

[9] 

There are, in addition to the above-mentioned naturally occurring pyrim- 
idine derivatives, many pyrimidines of synthetic origin which are widely 
used as therapeutic drugs. Of these, we have already mentioned the sulfona- 
mide drug, sulfadiazine (see Section 19-2D). Another large class of pyrimidine 
medicinals is based on 2,4,6-trihydroxypyrimidine. Most of these substances 
are 5-alkyl or aryl derivatives of 2,4,6-trihydroxypyrimidine — which is better 
known as barbituric acid and can exist in several tautomeric forms. 

OH 
HN^H, 

A N A 0H 




i\ k HN^ 

**k N A 0H O^N"\> 



(predominant 
form) 



barbituric acid 



For simplicity we shall represent barbituric acid as the triketo tautomer. Two 
of the more important barbituric acids are known as veronal (5,5-diethyl- 
barbituric acid) and phenobarbital (5-ethyl-5-phenyl barbituric acid). 



o 


O 


3 H 


i r c2Hs 

H 


veronal 


phenobarbital 



chap 25 heterocyclic compounds 686 

Barbituric acids are readily synthesized by the reaction of urea with substi- 
tuted malonic esters. 



O 



R 



II HNT XH 

NH 2 -C-NH 2 + C 2 H 5 2 CCHC0 2 C 2 H 5 ► | | + 2C 2 H 5 OH 

I X. X 

r Q S ^ Nq 



25-8 natural products related to purine and pteridine 

Heterocyclic nitrogen bases (other than pyrimidines) present in nucleic acids 
are the purine derivatives adenine and guanine (Section 17-6). Adenine is also 
a component of the trinucleotide adenosine triphosphate (ATP), whose 
structure is shown in Section 15-5. Nicotinamide-adenine dinucleotide 
(NAD®, Figure 18-3) is also a derivative of adenine. 

A number of alkaloids are purine derivatives. Examples include caffeine, 
which occurs in the tea plant and coffee bean, and theobromine, which occurs 
in the cocoa bean. The physiological stimulation derived from beverages such 
as tea, coffee, and many soft drinks is due to the presence of caffeine. 




O CH 3 

I 
CH 3 

theobromine 



25-9 natural products related to pjran 

The six-membered oxygen heterocycles, a-pyran and y-pyran, do not have 
aromatic structures and are rather unstable. Of more interest are the a- and 
y-pyrones, which differ from the corresponding pyrans in having a carbonyl 
group at the a- and y-ring positions, respectively. 



a-pyran -y-pyran a-pyrone 




The pyrones may be regarded as pseudoaromatic compounds — they are 
expected to have considerable electron derealization through n overlap of 
orbitals of the double bond, the ring oxygen, and the carbonyl group. Thus, 



sec 25.9 natural products related to pyran 687 

■y-pyrone should have at least some stabilization associated with contribu- 
tions of the electron-pairing schemes [10a] to [10e]. 




[10e] 

It is significant in this connection that y-pyrone behaves quite differently than 
might be expected from consideration of structure [10a] alone. For example, 
it does not readily undergo those additions characteristic of a,/?-unsaturated 
ketones and does not form carbonyl derivatives. 

The benzo derivatives of the pyrones are known as coumarin for benzo-a- 
pyrone and chromone for benzo- y-pyrone. 



o^o 





coumarin 

Coumarins occur in grasses, citrus peel, and the leaves of certain vegetables. 
Coumarin itself occurs in clover and is used as a perfume; it can be prepared 
by condensation of salicylaldehyde with acetic anhydride. 

,T ^V CH,CO,Na r^>S^C^ 



I 1 



CH 3 C0 2 Na (f Y C* 
(CH 3 CO) 2 — > [ 1 ^ 



CHO ^^ CH 



Chromones or benzo- y-pyrones are widely distributed in plant life, mostly 
as pigments in plant leaves and flowers. Particularly widespread are the 
flavones (2-phenylchromones), quercetin being the flavone most commonly 
found. 



HO 



OH O 

quercetin 

The beautiful and varied colors of many flowers, fruits, and berries are due 
to the pigments known as anthocyanins. Their structures are closely related 
to the flavones, although they occur as glycosides, from which they are 
obtained as salts on hydrolysis with hydrochloric acid. The salts are called 
anthocyanidins. Two examples follow: 





chap 25 heterocyclic compounds 688 



OH 



HO. 




OH 



e 
CI 



OH 

pelargonidin chloride 




CI 



delphinidin chloride 
(delphinium) 



Glycoside formation is through the 3-hydroxy group of the anthocyanidin. 



25-10 polyhetero natural products 

Of the many important polyhetero natural products with two or more 
different heteroatoms in one ring we have already mentioned penicillin 
(Section 17-2) and thiamine (Section 25-7). Another interesting example is 
luciferin, which is a benzothiazole derivative. 



HO 




N. N. 



-"^ 



XJ 



H 

C0 2 H luciferin 



Enzymic oxidation of luciferin is responsible for the characteristic lumi- 
nescence of the firefly. The luminescence arises because the oxidation product 
is formed in an excited state which liberates its excess energy as light rather 
than as heat. The relation between ground states and excited states of mole- 
cules will be pursued in the next chapter. 



summary 

In heterocyclic compounds at least one of the ring members is an atom other 
than carbon. Some of the important unsaturated heterocyclic ring systems 
containing nitrogen, oxygen, and sulfur are pyrrole [1], furan [2], thiophene 
[3], and pyridine [4]. 



Q Q Q O 



H 

[1] 



-o 

[2] 



S' 

[3] 



[4] 



All have aromatic character, [4] because of its benzenoid bonding, and [1], 
[2], and [3] because their heteroatoms have unshared pairs of electrons that 
can interact with the n electrons in the ring, giving a stable aromatic sextet. 

Strong acids cause polymerization of [1] and [2] and form salts with [4]. 
Strong bases form salts only with [1]. Electrophilic substitution occurs rapidly 
with [1], [2], and [3] and substitution occurs preferentially at the 2 position. 



exercises 689 



The activation of the ring results because the unshared pair of electrons on the 
heteroatom is able to stabilize the cationic intermediate [5]. 



o 



[5] 



All of the substitution reactions of activated benzene compounds, such as 
aniline and phenol, also occur with pyrrole and furan. Thiophene is not quite 
so reactive. Pyridine [4], on the other hand, is deactivated toward electro- 
philic substitution because in the analogous intermediate [6] the electro- 
negative nitrogen atom has a share in only six valence electrons. 



z 

[6] 

Substitution under vigorous conditions occurs at the 3 position. Pyridine, 
however, is subject to nucleophilic attack at the 2 position, and if hydride can 
be removed, either directly or by oxidation, substitution is achieved. 



z - H9 , 



N/ ^N^H ^N 



e 



Some important polycyclic and polyhetero compounds are indole, benzo- 
furan, quinoline, isoquinoline, imidazole, thiazole, pyrimidine, purine, and 
pteridine. 

A number of important natural products contain heterocyclic rings. Hemo- 
globin, bile pigments, chlorophyll, and vitamin B 12 all contain pyrrole rings. 

Virtually all alkaloids contain heterocyclic rings. The indole ring is present 
in serotonin, lysergic acid, and reserpine, all of which affect mental processes. 
Pyridine, quinoline, and isoquinoline are present in the structures of the 
tropane, cinchona, and opium alkaloids. The pyrimidine ring is present in 
barbituric acid and its derivatives. The purine ring system (fused pyrimidine 
and imidazole rings) occurs in nucleic acids and in caffeine. 

The six-membered unsaturated oxygen heterocycles (pyrans) are not aro- 
matic. Their keto derivatives (pyrones) are present in a number of natural 
products, many of which are highly colored. 

Compounds with two different heteroatoms in the same ring are also known 
and include penicillin, thiamine, and luciferin. 



exercises 

25-1 Suggest a feasible synthesis of each of the following compounds from the 
indicated starting material: 



chap 25 heterocyclic compounds 690 

a. II \ from thiophene 



^ from thiophene 




from benzothiophene 

from furan 
"CT "CH 2 CH 3 

25-2 m-Dinitrobenzene in the presence of potassium hydroxide and oxygen yields 
potassium dinitrophenoxide. Write a reasonable mechanism for this reaction 
with emphasis on determining the most likely arrangement of groups in the 
product. 

25-3 Predict the product(s) of the following reactions: 



0. 


LI 


MgBr 




KOH CH 3 I 








d. 


quinoline - 


knh 2 




e. 


isoquinoline 


KNH2 

> 



254 Would you expect porphyrin to possess aromatic character and to what 
degree would you expect the four N-Fe bonds to be equivalent ? Explain. 

25-5 The dextrorotatory ergot alkaloids are amides of isofysergic acid. Hydrolysis 
of these amides with aqueous alkali gives lysergic acid. What is the most 
likely structure of isolysergic acid ? Why does rearrangement occur on basic 
hydrolysis ? 

25-6 Uric acid, a purine derivative found mainly in the excrement of snakes and 
birds, has the molecular formula C5H4N4O3. On nitric acid oxidation it 
breaks down to urea and a hydrated compound called alloxan of formula 
C4.H 2 N204"H 2 0. Alloxan is readily obtained by oxidation of barbituric 
acid. What is the probable structure of alloxan and uric acid? Why is 
alloxan hydrated ? 



exercises 691 

25-7 2,6-Dimethyl-y-pyrone is converted by treatment with dimethyl sulfate and 
then with perchloric acid to [C 8 Hi 1 2 ]®C104 e . Simple recrystallization of 
this salt from ethanol converts it to [CgUnOz^ClO^. What are the 
structures of these salts, and why does the reaction with ethanol occur so 
readily ? 

25-8 Show the probable mechanistic steps involved in the preparation of cou- 
marin by the condensation of acetic anhydride with salicylaldehyde in the 
presence of sodium acetate as catalyst (Section 25-9). 



Si "IB 
Btt| ; 

8i§issiiii§ smS '-WW! 

HHrhwR 





f 11|SJ#gi§ «i^ till 










ifl^ffe* ...» 

^fill it i 



'& ; -tB 



i #*S^t tt!i# «SfSP pSm PiJbJ 
I te^ft |p|i IPSNI Itipl |f%PI ■ 



MM #tfe*S«li 












re 1$ 

Ml 



.0f<mSSSP^ 



chap 26 photochemistry 695 

Photochemistry deals with the chemical changes that are brought about by 
the action of visible or ultraviolet light. In Chapter 7 we discussed organic 
spectroscopy — how radiation from the various regions of the electromagnetic 
spectrum interacts with organic molecules. In that chapter and in subsequent 
references to light absorption we have dealt with it primarily as an analytical 
tool. We turn now to the chemical changes that sometimes result from the 
absorption of radiation. We shall see that the chemistry of electronically 
excited states is often quite different from that of ground states and that we 
can sometimes change the course of a reaction completely by activating the 
reactants by light rather than by heat. 

The regions of the electromagnetic spectum that we described in Chapter 7 
and the changes they bring about in organic molecules are listed in Table 
26-1. 

Absorption of infrared or microwave radiation produces vibrationally or 
rotationally excited molecules whose chemistry is almost unchanged. After all, 
in any sample of compound at room temperature, there will be a distribution 
of molecules among the various vibrational and rotational states. As the 
temperature is increased, the higher states become more and more populated. 
Although molecules do have more energy at higher temperatures and chemical 
reactions do occur at faster rates, it should be clear that we will not expect to 
encounter much new or unusual chemistry as a result of absorption of radia- 
tion quanta of such low energy as to only duplicate the effects of increasing 
temperature. 

As we proceed to higher-energy radiation we reach the visible and then the 
ultraviolet region of the spectrum. Here we find activation of a sort that can- 
not be duplicated by heat. Absorption of the higher-energy light quanta 
produces electronically excited states, and even the lowest of these are 
normally so far above the electronic ground states in energy that they are 
essentially completely unpopulated, even at temperatures of 100° or more. 
Although the lifetimes of excited states are normally short they are often 
quite long enough for important chemical reactions to ensue. 

The energy of quanta of visible light — light of wavelength 7500 to 4000 A — 
varies between 38 and 71 kcal/mole. These are energies of the same order of 
magnitude as bond energies (Section 2-4 A). Although red light (A near 7500 
A) is able to produce little in the way of chemical activation, the greater 



Table 26-1 Effects of various kinds of radiation 



type of radiation effect on absorbing molecule 



incre 
ene 


asing 
rgy 


X-rays and y-rays 
ultraviolet and 

visible light 
infrared radiation 
microwaves and 

radio waves 



bond rupture, decomposition" 
electronic excitation 

vibrational excitation 
rotational excitation 



' The results of X-ray absorption, not X-ray diffraction. 



chap 26 photochemistry 696 

energy of blue light {k near 4000 A) is often able to induce chemical change. 
Ultraviolet light is, of course, still more effective. As we pass to still higher 
energy quanta, such as X rays, we find more and more molecular disintegra- 
tion occurring as bonding electrons are not merely excited but are completely 
removed from the influence of the molecule's nuclei. We shall, therefore, 
concentrate on the effects of visible and ultraviolet light on organic mole- 
cules in the remainder of the chapter. Before beginning our discussion of the 
chemical reactions of excited states, however, we will review the principles of 
spectroscopy since activation can only occur if the light is actually absorbed. 
We will also examine more closely the ordinary fate of excited states ; that is, 
when their deactivation does not produce chemical change. 

26-1 light absorption, fluorescence, and 
phosphorescence 

When a quantum of visible or ultraviolet light is absorbed by a molecule, the 
time required to produce the new electronic arrangement is extremely short 
(~ 10~ 15 sec) and consequently the electronically excited state will be formed 
with the atoms essentially in their original positions (Franck-Condon prin- 
ciple). (The absorption of rf radiation in nmr spectroscopy is quite different. 
The time required to absorb such low-energy quanta is actually much longer 
than the vibrational times and is even longer than the reaction times of some 
chemical processes — see Section 7-6D.) 

What is the fate of the electronically excited molecule ? We have seen that 
in the first instant it is produced, it is just like the ground-state molecule as far 
as positions and kinetic energies of the atoms go, but has a very different 
electronic configuration. What happens at this point depends on several 
factors, some of which can be best illustrated by energy diagrams. We shall 
talk in terms of diatomic molecules, but the argument is easily extended to 
more complicated systems. 

Consider the diagram of Figure 26-1, which shows schematic potential 
energy curves for a molecule A — B in the ground state (A — B) and in an 
excited electronic state (A— B*). Each curve represents the energy of the 
molecule A— B as a function of the distance r between the two atoms. At 
very large values of r there is no interaction ; at very small values of r there is 
enormous repulsion and the energy of these configurations is very high; and 
at intermediate values there are energy minima which correspond to bonding 
between A and B. The vibrational levels of the bonds are represented by 
horizontal lines in the figure. 

The two curves in Figure 26-1 do not have identical shapes. The weaker 
bonding in the excited state tends to make the average distance r e between 
the nuclei at the bottom of the "potential well" greater in the excited state 
than in the ground state. 

The high-energy transition marked 1 in Figure 26-1 corresponds to ab- 
sorption of energy by an A— B molecule existing in a relatively high vibra- 
tional level. The energy change occurs with no change in r (Franck-Condon 
principle), and the electronic energy of the A— B* molecule so produced is 



sec 26.1 light absorption, fluorescence, and phosphorescence 697 



•-A + B 




Figure 26-1 Schematic potential energy diagram for ground and excited 
electronic singlet states of a diatomic molecule, A — B and A — B*, respectively. 
The horizontal lines represent vibrational energy levels. The wavy lines 
represent the arrival or departure of light quanta. 



seen to be above the level required for dissociation of A— B*. The vibration of 
the excited molecule therefore has no restoring force and leads to dissociation 
to A and B atoms. On the other hand, the somewhat lower-energy transition 
marked 2 leads to an excited vibrational state of A— B* which is not expected 
to dissociate but which can lose vibrational energy to the surroundings and 
come down to a lower vibrational state. This is called " vibrational relaxation " 
and usually requires about 10" 12 sec. The vibrationally "relaxed" excited 
state can now undergo several processes. It can return to ground state with 
emission of radiation (transition 3) ; this is known as fluorescence, the wave- 
length of fluorescence being different from that of the original light absorbed. 
Normally, fluorescence, if it occurs at all, occurs in 10~ 9 to 10~ 7 sec after 
absorption of the original radiation. In many cases, the excited state can also 
return to the ground state by nonradiative processes, the electronic excited 
state being converted to a vibrationally excited ground state of the original 
molecule which by vibrational relaxation proceeds to the ground state. This 
means, in effect, that the excess energy is shuffled into vibrational modes and 
thence by molecular collisions to the system as a whole. Sometimes, however, 
decay to a triplet state occurs before either complete vibrational relaxation or 
fluorescence takes place. Formation of a triplet state is of particular chemical 
interest because triplet states, even though of high energy, are often relatively 
long lived, up to a second or so, and can lead to important reaction products. 
To understand these processes we must consider in more detail the nature of 
singlet and triplet electronic states. 

We have already noted that in the ground states of ordinary molecules all 
the electrons are paired; we can also have excited states with all electrons 



chap 26 photochemistry 698 

paired. States with paired electrons are called singlet states. A schematic 
representation of the ground state (S ) and lowest excited singlet (S t ) elec- 
tronic configurations of a molecule with four electrons and two bonding and 
two antibonding molecular orbitals is shown in Figure 26-2. The 7i-electron 
system of butadiene (Section 6-7) provides a concrete example of this type of 
system. 

A triplet state has two unpaired electrons and is normally more stable than 
the corresponding excited singlet state because, by Hund's rule, less inter- 
electronic repulsion is expected with unpaired than paired electrons. An 
example of the lowest energy triplet electronic configuration (7\) is shown in 
Figure 26-2. The name "triplet" arises from the fact that two unpaired 
electrons turn out to have three possible energy states in an applied magnetic 
field. " Singlet" means that there is only one possible energy state in a mag- 
netic field. 

Conversion of the lowest singlet excited state to the lowest triplet state 
(S l -> T t ) is energetically favorable but usually occurs rather slowly, in 
accord with the so-called spectroscopic selection rules, which predict that 
spontaneous changes of electronic configuration of this type should have 
very low probabilities. Nonetheless, if the singlet state is sufficiently long 
lived, the singlet-triplet change, Sj -* T 1 (often called intersystem crossing), 
may occur for a very considerable proportion of the excited singlet molecules. 
The triplet and singlet states are actually different chemical entities. 

The triplet state, like the singlet state, can return to the ground state by a 
nonradiative process or by a radiative transition (T l -> S ). Such radiative 
transitions result in emission of light of considerably longer wavelength than 
either that absorbed originally or that emitted by fluorescence. This type of 
radiative transition is called phosphorescence. Because phosphorescence is a 
process with a low probability, the 7\ state may persist from fractions of a 
second to many seconds. For benzene at — 200°, absorption of light at 2540 A 
leads to fluorescence centered on 2900 A and phosphorescence at 3400 A 
with a half-life of 7 sec. 

There are a number of possible fates for excited singlet or triplet molecules. 



Figure 26-2 Schematic representation of the electronic configurations of 
ground and lowest excited singlet and triplet states of a molecule with four 
electrons and four molecular orbitals (i/j l , i/r 2 , <]i z , and i/r 4 ). 





ground 


lowest excited 


lowest triplet 




state (S ) 


singlet state (S l ) 


state (Ti) 




^ O 


o 


antibonding 


a 


^3 O 


© 


/~\ orbitals 


^ © 





{ij bonding 








orbitals 




*i © 


© 


© 



sec 26.2 light absorption and structure 699 

A singlet state can lose some vibrational energy and radiate the remainder 
(fluoresce) ; decay to a high vibrational state of the electronic ground state 
and then undergo vibrational relaxation to dissipate its energy; undergo a 
chemical reaction ; lose energy by vibrational relaxation and come down to a 
lower singlet state (if there is one) ; or decay to a triplet state. A triplet state 
can undergo similar changes except that intersystem crossing to an excited 
singlet state is seldom possible because the latter is usually of higher energy. 
We are chiefly concerned in this chapter with the chemical reactions of 
excited states, but we must first examine more closely the relation between 
molecular structure and the wavelength and intensity of absorption of ultra- 
violet and visible light, because photochemical changes depend on the system 
first being activated by the absorption of light quanta. 



26-2 light absorption and structure 

In Chapter 7 it was shown how the changes in wavelength of absorption of 
conjugated systems could be accounted for in terms of differences in the 
degree of resonance stabilization between ground and excited states. If the 
conjugated system of multiple bonds is long enough the absorption occurs in 
the visible part of the spectrum. Thus, 1 ,2-diphenylethene is colorless 
(/l max 3190 A), whereas l,10-diphenyl-l,3,5,7,9-decapentaene is orange (A max 
4240 A). 



\ /- CH=CH A / 

colorless 
(A max 3190A) 






I j>-CH=CH-CH=CH-CH=CH— CH=CH-CH=CH-^ J 

orange 
(A max 4240A) 

In general, the more extended a planar system of conjugated bonds is, the 
smaller is the energy difference between the ground and excited states. The 
importance of having the bonds coplanar should be obvious from considera- 
tion of preferred geometries of the resonance structures with formal bonds 
and (or) charge separations. The contribution of such structures to stabiliza- 
tion of the ground state increases with an increase in length of the con- 
jugated system; but stabilization of the excited state increases even more 
rapidly, so that the overall energy difference between ground and excited 
states decreases and hence the wavelength required for excitation shifts to 
longer wavelengths. 

The effect of substituents on the colors associated with conjugated systems 
is of particular interest in the study of dyes because, with the exception of 
compounds such as /^-carotene, which is an orange-red conjugated polyene 
occurring in a variety of plants and which is commonly used as a food color, 



chap 26 photochemistry 700 

most dyes have relatively short conjugated systems and would not be intensely 
colored in the absence of substituent groups. 



CH, 




-carotene (all trans double bonds) 
(A max 4500 A, £140,000) 



A typical dyestuff is 2,4-dinitro-l-naphthol (Martius Yellow), a substance 
that is used to dye wool and silk. (In addition to possessing color, dyes must be 
able to interact with the fibers of these substances. The chemistry of dyestuffs 
is considered in Chapter 28, which deals with polymers.) 



OH 




NO, 



N0 2 

Martius Yellow 



Although naphthalene is a conjugated system it is colorless to the eye, 
as is pure 1-naphthol. 2,4-Dinitronaphthalene is pale yellow, so the red- 
orange color of crystalline Martius Yellow is clearly due to a special com- 
bination of substituents with a conjugated system. 

Substitution of a group on a conjugated system which is capable of either 
donating or accepting electrons usually has the effect of extending the con- 
jugation. This is particularly true if an electron-attracting group is connected 
to one end of the system and an electron-donating group to the other. Thus, 
withp-nitrophenolate ion, we can expect considerable stabilization because of 
interaction between the strongly electron-donating — O e group and the 
strongly electron-accepting — N0 2 group. 



o 






e 
O 


N- 


■f> 


r\ i 


\ 


y) < 




6/ 


\ / 




/ 


o 






O 

e 






The high degree of electron derealization which can be associated with this 
system is clearly related to its absorption spectrum because, while p-nitxo- 
phenolate ion in water gives a strongly yellow solution (2 max 4000 A, e 15,000), 
/7-nitrophenol produces a less intensely colored greenish-yellow solution 
(l max 3200 A, e 9000). The important structural difference here is that the 
—OH group is not nearly so strong an electron-donating group as an — O e 
group, and electron derealization is therefore likely to be less important. 



O 

N- 
Oe 


o 


i~ll 1 . 


e 
O 

\ 


vJri < 


ed 






OH 



sec 26.2 light absorption and structure 701 

There are several connections in which changes of color resulting from 
interconversions of such substances as p-nitrophenol and the /?-nitropheno- 
late ion are important, including the practical one of the colors of acid-base 
indicators as a function of pH. However, before discussing some of these, it 
will be well to make clear that, in visual comparisons of the colors of sub- 
stances, it must be remembered that both wavelength and absorption coeffi- 
cient are involved in judgments of color intensities. The change from />-nitro- 
phenolate ion to j?-nitrophenol provides an excellent example of how both 
wavelength and absorption coefficient are affected by a structural change. It is 
possible for wavelength to shift without changing the degree of absorption 
and vice versa, so in speaking of one substance as being more "highly 
colored" than another we must be careful as to which features of the spectra 
we are actually comparing. 

Furthermore, a compound with an absorption maximum in the ultraviolet 
may be colored if the absorption band extends into the visible (Figure 26-3). It 
should be remembered that colored substances remove part of the white light 
spectrum and the color that registers in the eye is complementary to that 
which was absorbed. Thus compounds that absorb violet light (light of 
wavelength near 4000 A) appear green-yellow and those that absorb red 
light (light of wavelength near 7000 A) appear blue-green. 

The problems of correlating wavelength and absorption coefficient with 
structure can be approached in a number of ways. In an earlier discussion 
(Section 7-5), the excited electronic states were considered to have hybrid 
structures with important contributions from dipolar electron-pairing schemes. 
Thus for the first excited singlet state of butadiene, we may write these contrib- 
uting structures : 

e e ® e 

CH 2 -CH=CH-CH 2 < ► CH 2 -CH=CH— CH 2 

[1] [2] 

?. © 
< ► CH 2 =CH— CH — CH 2 < > etc. 

[3] 

Extension of this approach to benzene suggests the importance of reso- 
nance structures such as [4], [5], and so on for the excited singlet state of 
benzene. 



// vo 4 * etc. 



[4] [5] 

It is not unreasonable to suppose that substitution of an electron-attracting 
group at one end of such a system and an electron-donating group at the other 
end should be particularly favorable for stabilizing the excited state relative 
to the ground state wherein [4], [5], and so on are of negligible importance. 
At the same time, we would not expect that two electron-attracting (or two 
electron-donating) groups at opposite ends would be nearly as effective. 

The problem of predicting the absorption intensities is a difficult and 
complicated one. An important factor is the ease of displacement of charge 



chap 26 photochemistry 702 




visible - 



3000 



4000 



5000 



X, A 

Figure 26-3 Absorption in the visible region by a substance that has A^ 
in the ultraviolet. 



during the transition. On this account we expect substances such as j?-nitro- 
phenolate ions, which have roughly equivalent electron-donating and electron- 
attracting groups in each of their principal resonance structures, to absorb 
particularly strongly. 



e 

° / \ 

O 

electron 
attracting 



e 
O 



O 

e 



electron electron 

donating donating 



O 



electron 
attracting 



Figure 26-4 Two of the many resonance structures of the dye Crystal Violet. 




sec 26.3 photodissociation reactions 703 

We expect this factor to be especially favorable where equivalent resonance 
structures may be written. Many useful and intensely colored dyes have 
resonance structures of this general sort. A typical example is Crystal Violet, 
shown in Figure 26-4. 



26-3 photodissociation reactions 



We have already mentioned (Section 2-5B) that the chlorine molecule under- 
goes dissociation with near-ultraviolet light to give chlorine atoms and thereby 
initiates the radical chain chlorination of saturated hydrocarbons. Photo- 
chemical chlorination is an example of a photochemical reaction which can 
have a high quantum yield — that is, many molecules of chlorination product 
can be generated per quantum of light absorbed. The quantum yield, <J>, of a 
reaction is said to be unity when 1 mole of reactant ( s ) is converted to product(s) 
per mole of photons absorbed. (This quantity of light is called one einstein.) 

Acetone vapor undergoes a photodissociation reaction with 3130-A light 
with <I> somewhat less than unity and is of interest in illustrating some of the 
things which are taken into account in the study of photochemical processes. 

Absorption of light by acetone results in the formation of an excited state 
which has sufficient energy to undergo cleavage of a C— C bond (the weakest 
bond in the molecule) and form a methyl radical and an acetyl radical. 



o 

ho 



CH3-C-CH3 



o 

II 
CH3-C-CH3 



o 

II 

CH 3 -C- +CH3 



At temperatures much above room temperature, the acetyl radical breaks 
down to give another methyl radical and carbon monoxide. 

o 

II 

CH3-O ► CH 3 -+C=0 

If this reaction goes to completion, the principal reaction products are 
ethane and carbon monoxide. 

2 CH 3 - > CH3-CH3 

If the acetyl radical does not decompose completely, then some biacetyl is 
also formed. This reaction is quite important at room temperature or below. 

O OO 

II II II 

2 CH3-O ► CH 3 -C-C-CH 3 

biacetyl 

Lesser amounts of methane, hydrogen, ketene (see Section 12-4), and so on 
are also formed in the photochemical dissociation of acetone. 



chap 26 photochemistry 704 

A variety of dissociation-type photochemical reactions has been found to 
take place with other carbonyl compounds. Two examples are 



o o 

II hv II 

CH 3 — C— CH 2 CH 2 CH 3 ► CH 3 — C— CH 3 + CH,=CH, 

vapor phase 



O 

II 



CH 2 — CH 2 

hv 



CH, CH, r — ► + CO 

2 2 vapor phase 



\ / 

CHi CHj 



CH 2 — CH 2 



26-4 photochemical reduction 



One of the classic photochemical reductions of organic chemistry is the 
formation of benzopinacol, as brought about by the action of light on a 
solution of benzophenone in isopropyl alcohol. The yield is quantitative. 



/=\ 1 /=\ .. f\.. „ _.M V... f • 





2 4 /-C-\ / + H— C — OH > HO C C OH + C=0 

CH 3 /=K )=\ CH 3 



\J \J 



The light functions to energize the benzophenone, and the activated ketone 
removes a hydrogen from isopropyl alcohol. 



o 



CH 3 OH CH 3 

H-j-OH ► {J^^^} + -{-° H 



benzhydrol radical 
[6] 



Benzopinacol results from dimerization of benzhydrol radicals [6]. 



_ OH _ 

2 



<( )>— C— L \ ► benzopinacol 



The initial light absorption process involves excitation of one of the un- 
shared electrons on the carbonyl oxygen atom (« -> 71*). However, the excited 
singlet state undergoes facile intersystem crossing to give the longer-lived 
triplet state (5 t -»■ 7^) and it is the latter that abstracts the hydrogen atom from 
the alcohol. 

Because the quantum yields of acetone and benzopinacol are both nearly 
unity when the light intensity is not high, it is clear that two benzhydrol 
radicals [6] must be formed for each molecule of benzophenone that becomes 



sec 26.5 photochemical oxidation 705 

activated. This is possible if the hydroxyisopropyl radicals formed by Equa- 
tion 26- 1 react with benzophenone to give benzhydrol radicals. 

CH 3 O CH 3 OH 

i /=\ ii / = \ i r=\ i 



C-OH + i^J-C-^J > C=0 + ^J-^^J (26-1) 

CH 3 CH 3 

This reaction is energetically favorable because of the greater possibility 
for derealization of the odd electron in the benzhydrol radical than in the 
hydroxyisopropyl radical. 

Photochemical formation of benzopinacol can also be achieved from ben- 
zophenone and benzhydrol. 





O O 

The mechanism is similar to that for isopropyl alcohol as the reducing agent 
except that now two benzhydrol radicals are formed. 



O 

/ = \ I' / = \ 



OH . . OH 



\ r?A / — > 2 V_y ? u 



H 



t 



benzopinacol 

The reduction is believed to involve the triplet state of benzophenone by 
the following argument. Benzopinacol formation is reasonably efficient even 
when the benzhydrol concentration is low; therefore, whatever excited state 
of benzophenone accepts a hydrogen atom from benzhydrol, it must be a 
fairly long-lived one. Because benzophenone in solution shows no visible 
fluorescence, it must be converted to another state in something like 10~ 10 
sec, but this is not long enough to seek out benzhydrol molecules in dilute 
solution. The long-lived state is then most reasonably a triplet state. 



26-5 photochemical oxidation 



Molecular oxygen, 2 , is unusual in that its most stable electronic configura- 
tion is that of a triplet, • O— O • , in which the spins of the odd electrons are the 
same. The reactions of oxygen are in accord with this arrangement, for 
example, rapid reaction with radicals (Sections 2-5B and 22-4) or paramag- 
netic metal ions (Section 18-4). When oxygen is raised to its excited state by 
absorption of energy we find a different set of reactions entirely. The first 
excited state is a singlet and corresponds to the structure 0=0. Because 



chap 26 photochemistry 706 

oxygen does not absorb light in the accessible region of the visible or ultra- 
violet spectrum we cannot easily obtain singlet oxygen by simply irradiating 
oxygen. However, it is possible to obtain this chemical species if there is a 
sensitizer present in the system. A sensitizer is a compound which can absorb 
a light quantum and then transfer some of this energy to a second substance. 
Benzophenone is particularly useful for this purpose. We saw in the previous 
section how the excited singlet state of benzophenone changes to the excited 
triplet which can then abstract a hydrogen atom from alcohols. The triplet 
is also extremely efficient at transferring energy to oxygen. The sequence of 
reactions is shown. 

hv 
(C 6 H 5 ) 2 C=0 > (C 6 H 5 ) 2 C=0* (singlet) 

3665A 

(C 6 H 5 ) 2 C=0* (singlet) ► (C 6 H 5 ) 2 C=0* (triplet) 

(C 6 H 5 ) 2 C=0* (triplet) + 2 * (C 6 H 5 ) 2 C=0 + 0=0 (singlet) 

The reactions of the excited state of oxygen, 0=0, are characteristic of a 
molecule with paired electrons rather than of a diradical. It reacts with 
alkenes in a concerted manner by attaching itself to one carbon atom of the 
double bond while abstracting a hydrogen from the allylic position. 



H 3 C CH 3 H 3 C^ ^CH 2 

C C 

II + 0=0 ► I 

c c 

HjC" ^CH 3 HjC^I^O-OH 

33 3 CH 3 

The concerted nature of the reaction is shown by the fact that the abstracted 
hydrogen is always cis to the position of oxygen attack; see Exercise 26T0. 
This suggests that the reaction proceeds through a cyclic transition state such 
as the following : 

. X 

C' H 

li i 

Singlet oxygen reacts with conjugated dienes to form bicyclic compounds. 



+ 0=0 



^^ 



o 

I 
o 



We shall encounter this type of cyclization reaction again in the next chapter 
and we shall see that there is a distinct resemblance between the reactions of 
0=0 and CH 2 =CH 2 . 

Singlet oxygen can also be generated chemically by the oxidation of 
hydrogen peroxide in alkaline solution with two-electron oxidizing agents 
(Section 18-4) able to remove hydride ion from hydrogen peroxide. Hypo- 
chlorite is useful for this purpose. 

3 OCl ► 0=0 + HO e + Cl e 



sec 26.7 photochemical cycloadditions 707 

26-6 photochemical isomerization of cis- and trans- 
unsaturated compounds 

An important problem in many syntheses of unsaturated compounds is to 
produce the desired isomer of a cis-trans pair. In many cases, it is necessary 
to utilize an otherwise inefficient synthesis because it affords the desired 
isomer, even though an efficient synthesis of the unwanted isomer or of an 
isomer mixture may be available. An alternative way of attacking this prob- 
lem is to use the most efficient synthesis and then to isomerize the undesired 
isomer to the desired isomer. In many cases this can be done photochemically. 
A typical example is given by cis- and ?ra«^-stilbene. 





\=/ H H 

trans cis 

Here the trans form is easily available by a variety of reactions and is much 
more stable than the cis isomer because it is less sterically hindered. However, 
it is possible to produce a mixture containing mostly the cis isomer by 
irradiating a solution of the trans isomer in the presence of a suitable photo- 
sensitizer. This process in no way contravenes the laws of thermodynamics, 
because the input of radiant energy permits the equilibrium point to be shifted 
from what it would be normally. 

Another example is provided by the equilibration of l-bromo-2-phenyl-l- 
propene. The trans isomer is formed to the extent of 95 % in the dehydro- 
halogenation of l,2-dibromo-2-phenylpropane. 






W Br \_J H W Br 

\ / NaOCH 3 \ / \ / 

Br-C-C-H ^„ » C=C + C = C 

/ \ HOCH 3 / \ / \ 

H 3 C H H 3 C Br H 3 C H 

trans cis 

95% 5% 

Photoisomerization of the elimination product with 2-naphthyl methyl ketone 
as sensitizer produces a mixture containing 85% of the cis isomer. 

In the practical use of the sensitized photochemical equilibrium of cis and 
trans isomers, it is normally necessary to carry out pilot experiments to 
determine what sensitizers are useful and the equilibrium point which each 
gives. 



26-7 photochemical cycloadditions 

In the next chapter cyclization reactions of various kinds will be discussed, 
including those resulting from absorption of light. We shall only point out 



chap 26 photochemistry 708 

here how photochemistry has in recent years enabled some cyclic compounds 
with unusual structures to be prepared. 




R = 7-butyl [7] [8] [9] 

Compounds [8] and [9] are moderately stable at room temperature 
although they are clearly of higher energy than benzene derivatives. They will 
be recognized as two of the structures that were suggested a century ago for 
the structure of benzene by Dewar and Ladenburg (Chapter 6). The terms 
Dewar benzene and Ladenburg benzene are often used for [8] and [9], 
respectively; [9] is also called prismane, because of its resemblance to a 
prism. Although the structure of Dewar benzene is often drawn as in [8] 
it is preferable to show the molecule's partly folded geometry and the close- 
ness of the bridgehead carbons, as in [10]. Thus [10] should not be regarded 
as contributing to the resonance hybrid of benzene because the geometries of 
[10] and benzene are very different. 



H 
H 




[10] 



summary 

Absorption of visible or ultraviolet light by an organic molecule in a ground 
singlet state, S , produces a short-lived electronically excited singlet state, 
S (S t if the lowest singlet state), which may undergo a chemical reaction or 
return to the ground state by one of the following means : it may fluoresce (lose 
some energy by vibrational relaxation and then radiate the remainder) ; it may 
undergo vibrational relaxation to a lower singlet state, if there is one, or all 
the way to the ground state (5 ) ; or it may undergo intersystem crossing to 
produce a longer-lived triplet state (5 1 -»7' 1 ) which may, in turn, either phos- 
phoresce (lose energy by vibrational relaxation and then radiate the remain- 
der), undergo a chemical reaction, or act as a photosensitizer by transferring 
its energy to a second molecule. 



exercises 709 

Photoactivation requires that light be absorbed; photosensitizers (e.g., 
benzophenone) are useful for activating other compounds that do not absorb 
light in an accessible part of the spectrum. 

Conjugation stabilizes excited states more than ground states and conju- 
gated compounds thus tend to absorb at the longer wavelengths. 

Some of the reactions undergone by compounds that have been raised to 
their excited electronic states (either by direct light absorption or by energy 
transfer from a photosensitizer) include dissociation, oxidation or reduction, 
trans-cis isomerization, and cycloaddition. Many of the sensitized reactions 
involve the triplet state of aromatic ketones such as benzophenone. 



HO OH 
R 2 CHOH I I 



Ar 2 C= 0*(triplet) 



RCH=CHR((rans) 



Ar 2 C-CAr 2 + R 2 C=0 
-*■ singlet 0=0 
■*- RCH=CHR(ra) 



Singlet oxygen (0=0), which can also be generated chemically, undergoes 
a number of reactions unknown to ground-state (triplet) oxygen, including 
the following: 



0=0- 



«f I (cis mechanism) 

' OOH 



Derivatives of both Dewar benzene and Ladenburg benzene (prismane) 
have been prepared by photochemical isomerization of derivatives of ordinary 
benzene. 



exercises 

26-1 The fluorescence of many substances can be "quenched" (diminished or 
even prevented) by a variety of means. Explain how concentration, tempera- 
ture, viscosity, and presence of dissolved oxygen and impurities might affect 
the degree of fluorescence observed for solutions of a fluorescent material. 
Would you expect similar effects on phosphorescence ? Explain. 



chap 26 photochemistry 710 

26-2 Explain qualitatively how temperature could have an effect on the appear- 
ance of the absorption spectrum of a system such as shown in Figure 26-1, 
knowing that most molecules are usually in their lowest vibrational state 
at room temperature. 

26-3 Make diagrams of at least five different singlet states and three different 
triplet states of the system shown in Figure 26-2. 

26-4 What visible color would you expect the substance to have whose spectrum 
is shown in Figure 26-3 ? 

26-5 The 77 -> 77* absorption spectra of trans,trans-, trans,cis-, and cis,cis-l,4- 
diphenylbutadiene show maxima and e values (in parentheses) at about 
3300 A (5.5 X 10 4 ), 3100 A (3 x 10*), and 3000 A (3 x 10 4 ), respectively. 
What is the difference in energy between the transitions of these isomers in 
kilocalories per mole? Why should the trans,trans isomer have a different 
A max than the other isomers? (It may be helpful to make scale drawings or 
models.) 

26-6 How would you expect the spectra of compounds [11] and [12] to compare 
with each other and with the spectra of cis- and ?ra«^-l,2-diphenylethene 
(stilbene) ? Explain. 




26-7 Why must the resonance forms [1], [2], [3], etc., for butadiene correspond 
to a singlet state ? Formulate the hybrid structure of a triplet state of buta- 
diene in terms of appropriate contributing resonance structures. 



26-8 a. p-Nitrodimethylaniline gives a yellow solution in water which fades 
to colorless when made acidic. Explain. 
b. p-Dimethylaminoazobenzene (Section 22-9B) is bright yellow in aqueous 
solution (A max 4200A) but turns intense red (A max 5300 A) if dilute acid 
is added. If the solution is then made very strongly acid, the red color 
changes to a different yellow (A raax 4300 A) than the starting solution. 
Show how one proton could be added to ^-dimethylaminoazobenzene 
to cause the absorption to shift to longer wavelengths and how addition 
of a second proton could shift the absorption back to shorter wavelengths. 



26-9 The well-known indicator and laxative, phenolphthalein, undergoes the 
following changes as a neutral solution is made successively more basic : 













exercises 


711 


, H( v^ 


il f 


> 


.OH 


HO - 


i rV 




V 


\X 


J 




k. 


KkJ 




c 


ii 








- c 


II 
o 




H( S^ 








e °r^ 






-U 1 


>< 


=\ 


_ r\ 


, kl 


>=o 




d 


— \J 


n 




U 1 


"co 2 e 






k> 


v co 2 e 






Some of these forms are colorless, some intensely colored. Which would you 
expect to absorb at sufficiently long wavelengths to be visibly colored ? Give 
your reasoning. 



26-10 A steroid molecule, only part of whose structure is shown below, contained 
one atom of deuterium in a position cis to the hydroxyl group. On irradiation 
of a mixture of this compound, oxygen, and an aryl ketone sensitizer, one 
of the following hydroperoxides was obtained in high yield and the other in 
very small yield. Identify the major product. Give your reasoning. 



HO' 




o=0 



£s 



D 



HO 




0-OH 



xrs 



D 



HO 




O-OD 



H 



chap 27 cyclization reactions 715 

Ring formation can take place if functional groups in the same molecule react 
intramolecularly. Cyclic structures can also be formed by the addition re- 
action of two unsaturated molecules with one another. In this chapter we 
shall first examine intramolecular ring closure, especially as it involves car- 
bonyl compounds, and then examine the addition reactions of alkenes and 
polyenes that lead to cyclic structures. The first group of reactions enables us 
to review some familiar chemistry; the second introduces new reactions and 
new concepts and takes us deeper into molecular orbital theory than we have 
hitherto gone. Those whose interests are less theoretical may not wish to 
pursue this subject beyond Section 27 • 4. 



27" I cyclization reactions of carhonyl compounds 

Carboxylic acids react with alcohols and amines to form esters and amides. 

RC0 2 H + HOR ► RC0 2 R + H 2 

RC0 2 H + H 2 NR > RCONHR + H 2 

Hydroxy acids and amino acids undergo the same reactions to give cyclic 
structures (lactones and lactams, Sections 13-10B and 17-2), provided that 
the functional groups are favorably situated with respect to one another. 

O 

o // 

f H 2 C-C 

HOCH 2 CH 2 CH 2 C ► / \ + H 2 



OH 



CH 2 

y-butyrolactone 



O 

o // 

// H 2 C-C 

H 2 NCH 2 CH,CH,C * I \ + H,0 

2 2 2 2 \ H,C ,NH 



OH 



CH, 

y-butyrolactam 



Indeed, these reactions occur much more readily than their intermolecular 
counterparts; 7- and <5-hydroxyacids cyclize spontaneously to give lactones. 
The same pattern applies to anhydride formation; heating acetic acid has 
little effect but warming succinic acid or phthalic acid produces the an- 
hydrides. 

/° 



-C H 2 C-C 

OH A I \ 



<a \ / 

H,C-C^ H 2 C-C 



o 



OH 



V 




o 



chap 27 cyclization reactions 716 

When more than three carbon atoms intervene between the carboxyl groups 
of a dicarboxylic acid, pyrolysis of the acid (or its thorium salt) produces 
ketones (Section 13 ■ 1 IB). 

^C0 2 H ^^ 

(CH 2 )„ — ^-» (CH 2 ), C = + C0 2 + H 2 

Vo 2 H ^-^ w>3 

Spontaneous cyclization occurs with y- and cS-hydroxyaldehydes and 
ketones to give hemiacetals (Section 1 1 ■ 4B). 

OH 

P ' 

/ H 2 C-C-H 

HOCH 2 CH 2 CH 2 C ► I \ 

\ H 2 C O 

H \H 2 



The best known examples are provided by carbohydrates (Section 15-3). 

The Claisen condensation of esters (Section 13-9B) has its intramolecular 
counterpart in the Dieckmann reaction. 

O O O O 

// // NaOC 2 H 5 II II 

CH 3 -C + CH 3 -C — — -—* CH 3 -C-CH 2 -C-OC 2 H 5 + C 2 H 5 OH 

\ " \ C 2 H 5 OH J i 2 5 2 5 



o 

_ / o 

h CH 2 C^ qc ^ ^ H 2/ C-^ 

I O ► H 2 C I + C 2 H 5 OH 

S CH,-C 2 \ 

2 ^OC 2 H 5 C0 2 C 2 H 5 

The course of this cyclization reaction is the same as that of the Claisen 
reaction. Ethoxide ion abstracts a proton from one of the two activated 
methylene groups in the diester to give an anion. 

O o 

II II 

HC /CH 2 -C-OC 2 H 5 

2| + C 2 H 5 O e 



H 2 C. H,C^ e 

^CH 2 -C-OC 2 H 5 2 ^CH-C-OC 2 H 5 

II II 

o o 

The highly nucleophilic anion thus formed attacks a carbonyl group and, 
particularly in dilute solution, this is likely to be the group in the same 
molecule. Loss of ethoxide ion gives the product. 



sec 27.1 cyclization reactions of carbonyl compounds 717 



H 2 C 

I 
H,C. 



O 

II 

,CH,— C-OC 2 H s 



V CH— C— OC 2 H 5 

II 

o 



- H 2 c( 



C-OC 2 H 5 

yi^CH CO2C2H5 

H 2 



H,C 



H 



.0 



e 



\ ^CHC0 2 C 2 H 5 + C 2 H 5 
H 2 



Another very useful ring closure reaction that bears a superficial resem- 
blance to the above reaction is the acyloin condensation. The reactant is again 
a dicarboxylic ester but the reagent is sodium dispersed in a hydrocarbon 
solvent instead of sodium ethoxide dissolved in ethanol. This seemingly small 
difference in the reagent changes the course of the reaction completely. 
Sodium is unable to function as a base in a hydrocarbon solvent and instead 
it reacts as a one-electron reducing reagent to produce a radical anion at each 
carbonyl group. Ring closure followed by elimination of ethoxide ions and 
further reduction produces the dianion [1] which on addition of water yields 
the hydroxy ketone [2] (an acyloin). 



CH 

I 
CH 




I 
CH 2 — C— OC 2 H 5 

CH 2 



-C — OC 2 H, 



I 

o 6 



H 2 O fc 



H 2 C /C ^ ' 



C-OC,H, 



(-2 C 2 H 5 Oe) 



H 2 C 

I 
H,C. 



H, O e 






2Na- 



H 2 
H 2 C^ ^<r 



H, 



H,C 



-OH 



1 2 , -\ / ^/^'\ 



H 2 



OH 




[1] 



.O fc 



^O fc 



chap 27 cyclization reactions 718 

This reaction, which works very well for large rings, has been used in an 
ingenious synthesis of a catenane, a compound whose two rings are not joined 
by bonds but are held together like links in a chain. A large-ring compound 
[5] was prepared by the acyloin condensation followed by Clemmensen re- 
duction (Section 1 1 ■ 4F) using deuterated reagents. This produced a large 
carbocyclic ring containing some deuterium label. (Partial exchange of the 
a hydrogens occurred and. produced a compound containing on the average 
five deuterium atoms per molecule.) 



DCl 




Zn-Hg 




The labeled compound was then dissolved in xylene, more [3] added, and 
the acyloin condensation repeated. If the long chain of [3] happens to be 
threaded through a molecule of [5] when ring closure occurs the catenane [6] 
will be produced. 





The product was purified by chromatography and after all traces of [5] 
were removed, it was found that some deuterium label remained, suggesting 
the presence of about 2% [6] mixed with [4]. Cleavage of the hydroxy ketone 
ring by oxidation produced a dicarboxylic acid containing no deuterium and 
the large-ring deuterium-containing hydrocarbon [5]. 



27-2 cjcloaddition reactions of carbon-carbon 
multiple bonds 

The tendency of an alkene such as ethene to undergo a cycloaddition reaction 
with another unsaturated molecule depends on two important factors: 
whether the other molecule contains isolated or conjugated double bonds, 
and whether the system is activated by heat or by light. For example, most 
alkenes show little tendency to dimerize to cyclobutanes thermally (A), but 
the reaction can usually be brought about readily by the action of light (hv). 

II + II -^ □ 

On the other hand, substituted alkenes usually react readily on being 
warmed with conjugated dienes to give cyclohexenes. Light (hv) is not re- 
quired for this reaction. 



<r 



sec 27.2 cycloaddition reactions of carbon-carbon multiple bonds 719 

The reaction of a conjugated diene with an alkene is known as the Diels- 
Alder reaction and is described in some detail in the next section. 



A. D1ELS-ALDER REACTION 

Although the Diels-Alder reaction can be conducted with ethene as the alkene 
(often referred to as the dienophile), addition occurs much more readily if the 

O 

II 
alkene contains electron-withdrawing groups such as — C— , — C=N, or 

— N0 2 . 

One of the reasons that the reaction has proved of value, especially in the 
synthesis of natural products, is that it is highly stereospecific. First, and 
most obvious, the diene reacts in the s-cis 1 conformation of its double bonds 
because the double bond in the product (a six-membered ring) necessarily has 
the cis configuration. 



CH 2 h 



I 



CH, 



^H 



s-cis 
conformation 



CH 2 =CH 2 

H 
X 

I II 

2 ^(y ^H 



H 2 C^ ^C^ 



H 2 

stable cis 
double bond 



CH 2 H 

I 

H /C% CH 2 

s-trans 
conformation 

CH 2 =CH 2 

CH 2 H 

V CH 2 / 

c==c 
/ ch\ 

H X CH 2 

highly strained 
trans double bond 



Cyclic dienes with five- and six-membered rings usually react readily because 
they are fixed in s-cis configurations. 



C(CN) 2 
+ II 

C(CN) 2 



25° 




(CN) 2 
(CN), 



Second, the configurations of the diene and the dienophile are retained in 
the adduct. This means that the reactants (or addends) come together to give 
cis addition. Two illustrative examples follow which are drawn to emphasize 
how cis addition occurs. In the first example, dimethyl maleate, which has cis 



1 The designation s-cis means that the double bonds lie in a plane on the same side 
(cis) of the single bond connnecting them. The opposite and usually somewhat more stable 
conformation is called s-trans. 



chap 27 cyclization reactions 720 

ester (C0 2 CH 3 ) groups, adds to 1,3-butadiene to give a cw-substituted 
cyclohexene. 







ocH 3 

.OCH, 



V 



(shows retention of 
configuration in the 
dienophile) 



In the second example, cis addition of a dienophile to trans,trans-2,4-hexa.- 
diene is seen to yield the product with the two methyl groups on the same side 
of the cyclohexene ring. 



CH, 



H,C 





CH, 



(shows retention of 
configuration of the 
diene methyl sub- 
stituents) 



The Diels-Alder reaction is believed to occur by a one-step synchronous 
process in which bonds form simultaneously between each end of the diene 
and the dienophile. 

The difference between thermal activation and photoactivation of dienes is 
shown in Figure 27-1. 

The thermal process is a Diels-Alder reaction between two molecules of 
1,3-butadiene, one of which acts as the diene and the other as the dienophile. 
The photochemical reaction is analogous to a reaction between two alkenes. 
The absorption of a photon by 1,3-butadiene is expected to activate the 



Figure 27-1 Dimerization of 1,3-butadiene by thermal activation and photo- 
activation. 



2CH,=CH-CH=CH, 




1,4 addition 



(trans) (cis) 



n£>. 



2 addition 




sec 27.2 cycloaddition reactions of carbon-carbon multiple bonds 721 

molecule, but it is less obvious why the excited state prefers to react by the 
1,2-addition route rather than by the usual 1,4-path. There is no reason, how- 
ever, to expect electronically excited states, particularly those with different 
arrangements of electron spin, to undergo the same reactions as ground states. 
A rationale for the different routes has been developed recently based on 
orbital symmetry considerations and is discussed in Section 27 • 5. 



B. CYCLIZATION REACTIONS OF ALKYNES 

Alkynes can act as dienophiles in the Diels-Alder reaction to produce non- 
conjugated cyclohexadienes. 




Acetylene can be readily polymerized to cyclooctatetraene by the action of 
nickel cyanide. 

Ni(CN) 2 
4HC = CH — ~^> 
50° 

80 90° 

With this reaction, cyclooctatetraene could be manufactured easily on a 
large scale; however, profitable commercial uses of the substance have yet 
to be developed. 

It is easy to become confused about the reactions of alkynes with transition- 
metal ions. Acetylene, for example, is hydrated under the influence of Hg" 
(Section 5-4); it forms salts with Ag 1 (Section 5-5); it is dimerized to vinyl- 
acetylene by Cu 1 (Section 5-4); it is polymerized to cyclooctatetraene by Ni u 
(above) ; and we shall see later in this chapter that it and other alkynes can be 
oxidatively coupled by the action of Cu 11 . 

C. 1,3-DIPOLAR ADDITIONS 

Alkenes and some other compounds with multiple bonds undergo 1,3-cyclo- 

addition with a variety of substances which can be formulated as 1,3-dipolar 

© e 

molecules of the type X — Y — Z. 

C e e — C \ 

II + X-Y-Z ► I V 

The 1,3-dipolar compounds seldom carry full formal charges on the terminal 
atoms and, indeed, in the list of these reagents shown in Figure 27-2, the 
1,3-dipolar form is not the one we would regard as the most important of the 
forms contributing to the resonance hybrid. (Each of the 1,3-dipolar struc- 
tures shown in Figure 27 • 2 has an atom with an incomplete octet whereas in 



chap 27 cyclization reactions 722 



a e 
ozone ®0— 0— e < ► 0=0— O 

® e a e e & 

organic azides R-N~N=N < ► R— N=N=N < ► R— N-N=N 

nitrile oxides R-C=N-0 < > R— C=N-0 

diazoalkanes R 2 C— N=N < ► R 2 C=N=N 





Figure 27-2 Some 1,3-dipolar reagents. 



the 1,2-dipolar structures each atom has a rilled octet.) Nonetheless, alkenes 
add 1,3 to these substances to give five-membered rings. 

A simple example is the addition of phenyl azide to norbornene. Here the 
azide is written to correspond to the resonance form that appropriately 
accounts for the occurrence of the addition. 



/C 6 H 5 

•N 
\ 

norbornene phenyl azide 

In all of these reactions heterocyclic compounds are formed, although the 
ozone adduct undergoes further reaction (Section 4-4G). The basis for the 
names shown was given earlier (Chapter 25). 



II + O -* I o a trioxole 

' ^ O 

R 

^C \ 

II + N ► [ N a triazole 

^ N 



— c 

1 

-C- 

1 


\ 
o 




R 

I 


1 
— c- 

1 

— C- 

1 


N 



R 

I 

C« 



R 

I 



c \ -C'\\ 

II + N > | N an oxazole 

R ^r< R I V R 

|| + n ► I N a diazole 

/C v e// -C^S 



sec 27.3 fluxional systems 723 

These reactions, which are believed to be one-step synchronous processes 
like the Diels-Alder reaction, also take place with alkynes. 



+ X-Y-Z ► [I Y 

Z 



27-3 fluxional systems 



In earlier chapters we frequently encountered the phenomenon of tautom- 
erism, the rapid equilibration of structural isomers. 

O O O OH 

2,4-pentanedione and || || || | 

its enol form (Section CH 3 -C-CH 2 -C-CH 3 > CH 3 -C-CH=C-CH 3 

12-6) 



2-hydroxypyridine and 
oc-pyridone (Section 
18 -IE) 



H 



OH O 

barbituric acid, keto -**\ ^\ 

and enol forms (Section ^ |] - > H ¥ | 



25-7) 



HO ^NT X)H Or ^N' X> 

H 



All of the above examples involve proton shifts but there is nothing in the 
definition that restricts us to this kind of process. For example, a rapid 
equilibrium exists at 100° between 1,3,5-cyclooctatriene and the bicyclic com- 
pound shown here, which is also an example of tautomerism. (The term 
"valence tautomerism" or "valence isomerism" is sometimes used to de- 
scribe such equilibria and the reactions are sometimes called Cope rearrange- 
ments.) 




155 



The nmr spectra of systems such as this clearly reveal the position of 
equilibrium and even the rate at which the forward and reverse reactions 
occur. (See Section 7-6D for a discussion of how nmr can be used to measure 
the rates of conformational change.) 

The situation with cyclooctatraene is similar. Although the eight-membered 
ring is the major form in the equilibrium mixture some of its reactions are 
those of the bicyclic tautomer. 



chap 27 cyclization reactions 724 




CU 



\maleic anhydride 



XI 

-ci 




Tautomerism, which involves the relocation of atoms, should not be con- 
fused with resonance, which is formulated as a dispersal of electrons over 
several nuclei and which can be represented by drawing two or more valence 
structures in which the atomic locations are the same. Tautomers are thus 
distinctly different chemical entities with different atomic locations. It is also 
possible to have equilibria between molecules with the same formal struc- 
ture — for example, the rapid shifting of a hydrogen atom between the two 
oxygens of the carboxyl group in acetic acid or the slow equilibration of the 
methylene groups in 1,5-hexadiene. 



CH, 



O* 



OH 



OH 

J 

\ 
O 



50% 



50% 



50% 



50% 



This situation is very different from resonance because the atomic locations 
are different, and it is not strictly tautomerism because the molecules are not 
isomers. The term fluxional molecules has been coined to describe the partic- 
ipants in such equilibria. In the absence of some sort of label it is not possible 
to distinguish the two forms; nonetheless, a pathway between the fluxional 
molecules must exist and this has become a matter of interest to many 
chemists. 

A facile pathway for the 1,5-hexadiene equilibration exists in which bond 
rupture and bond formation occur at the same time and which has the cyclic 
transition state shown. 



■4? 



transition state 



A curious fluxional molecule, called barbaralane [7], is shown by its nmr 
spectrum to exist in two equivalent forms that are in rapid equilibrium at 
room temperature. 



sec 27.4 annulenes 725 




[7] 




If the CH 2 group is replaced by a CH=CH group the number of fluxional 
forms rises dramatically (see Exercise 27 • 4). The latter compound has been 
given the name bullvalene. (The classical languages have heretofore provided 
the basis for naming new compounds but this approach is now in some danger 
of being replaced by whimsy.) 

The ease with which the two forms of barbaralane interconvert reflects the 
importance of synchronous bond breaking and bond making in lowering the 
energy of a transition state. An important part of this is the fact that the re- 
acting atoms are held in favorable positions by the ring structure. 

The similarity between the equilibrium in 1,5-hexadiene and in barbaralane 
can be seen if we draw the latter so that the hexadiene portion of the molecule 
is clearly displayed. 





27-4 annulenes 



There has been considerable interest for many years in the synthesis of conju- 
gated cyclic polyalkenes with a large enough number of carbons in the ring to 
permit attainment of a strainless planar structure. Inspection of models shows 
that a strainless structure can only be achieved with two or more of the double 
bonds in trans configurations, and then only with a large enough ring that the 
"inside" hydrogens do not interfere with one another. 

In discussing compounds of this type, it will be convenient to use the name 
[njannulene to designate the simple conjugated cyclic polyalkenes, with n re- 
ferring to the number of carbons in the ring — benzene being [6]annulene. The 
simplest conjugated cyclic polyolefin that could have a strainless planar ring 
containing trans double bonds, except for interferences between the inside 
hydrogens, is [10]annulene. Inside-hydrogen interferences are likely to be of 
at least some importance in all annulenes up to [30]annulene. 




chap 27 cyclization reactions 726 

Several annulenes have been synthesized and found to be reasonably stable 
— at least much more so than could possibly be expected for the corresponding 
open-chain conjugated polyenes. An elegant synthesis of [18]annulene pro- 
vides an excellent illustration of some of the more useful steps for preparation 
of annulenes. The key reaction is oxidative coupling of alkynes by cupric 
acetate in pyridine solution. 

Cu" 

2 RC=CH > RC=C— C=CR 

C5H5N 

This type of oxidative coupling with 1,5-hexadiyne gives a 6% yield of the 
cyclic trimer [8], which rearranges in the presence of potassium ?-butoxide 
to the brown, fully conjugated 1,2,7,8,13, 14-tridehydro[18]annulene [9]. 



3 HC=C-CH,-CH 2 -C=CH 



C 5 H 5 N 





[9] [18]annulene 

Hydrogenation of [9] over a lead-poisoned palladium on calcium carbonate 
catalyst (the Lindlar catalyst, of general utility for hydrogenation of alkynes 
to alkenes) gives [18]annulene as a brown-red crystalline solid, reasonably 
stable in the presence of oxygen and light. 



27-5 orbital symmetry and cycloaddition 

We pointed out earlier that an alkene cyclodimerization 2 usually requires 
photoactivation whereas the Diels- Alder reaction between an alkene and a 
conjugated diene occurs thermally. A rationale for this difference and for the 
stereochemistry of a large number of cyclization and ring-opening reactions 
has been developed recently by several theorists, including Longuet-Higgins, 
Fukui, Woodward, and Hoffmann, using molecular orbital theory. In 
Chapter 6 we described the four n molecular orbitals in 1,3-butadiene that 
result from interaction of four p orbitals, one on each carbon atom. The 

2 A few alkenes such as tetrafluoroethene, CF 2 = CF 2 , cyclodimerize thermally but these 
reactions are known to go by a radical, not a concerted, pathway. 



sec 27.5 orbital symmetry and cycloaddition 727 

ff-bonded skeleton of the molecule is ignored in this treatment and only the 
7i electrons are considered. 



-c— c 



The wave functions that describe the energies of an electron in each of the 
four p orbitals of butadiene can be combined algebraically to give us an 
approximation set of four molecular orbitals with different energies. (These 
molecular orbitals are linear combinations of atomic orbitals and the pro- 
cedure is thus known as the LCAO approach.) Two of these molecular orbitals 
are bonding (energy lower than that of the isolated p orbitals) and two are 
antibonding (energy higher than that of the isolated orbitals). In a crude 
analogy the molecular energy levels can be compared with the energies of 
the standing waves of a vibrating string. The energy of a standing wave with 
a given amplitude increases with the number of nodes as shown on the left 
side of Figure 27-3. 

In a molecular orbital made up of a linear combination of p orbitals, the 
coefficients of the p-orbital functions can have positive or negative values. 
If the signs of the coefficients of two adjacent p orbitals overlapping in the % 
manner are the same, then the positive parts of the atomic orbital lobes are 
pointed in the same direction and we say that there is no node between the 
atoms and that the molecular orbital is bonding. If the signs of the coefficients 
for the two adjacent orbitals are different, the arrangement has a node and is 
antibonding between these atoms. Figure 27 ■ 4 shows schematically the bond- 
ing and antibonding arrangements for the two p-7t orbitals in ethene. 

A schematic representation showing the nodes for the simple LCAO molec- 
ular orbitals of butadiene is given on the right side of Figure 27-3. Here, the 
relative sizes of the orbitals are drawn to reflect the values obtained by nu- 
merical calculations. The first two molecular orbitals are bonding (zero and 
one node) and the second two are antibonding (two and three nodes) (see 
Section 6 ■ 7). In the ground state of butadiene, the first two orbitals are doubly 
occupied whereas in the excited state an electron is raised from molecular 
orbital ^ 2 t0 ^3 ( see Figure 26-3). 

A rule for determining whether or not cycloadditions are allowed can be 
stated as follows : The orbitals that overlap in the transition state between the 
highest occupied level of one reactant and the lowest unoccupied level of the 
other reactant must be of the same symmetry (sign). The % orbitals of ethene 
and 1,3-butadiene are shown in Figure 27-5. 

This rule predicts that concerted combination will not occur between two 
molecules of ethene in their ground states because the lobes of the lowest 
unoccupied level in one molecule and the highest occupied level in the other 
do not have corresponding signs. However, if one of the ethene molecules is 
raised to its first excited state by light absorption the situation is altered. The 
lobes of the orbitals of the excited state and those of the lowest unoccupied 
level of the second ethene molecule now correspond and combination can 
occur (see Figure 27 -6). 



chap 27 cyclization reactions 728 




vibrating string 



+ \/ ., \/ 



nodes" 



"if ^4 



"J fc 



nodes 



+'\ Is'-, 



LCAO molecular orbitals 



Figure 27-3 Nodes in a vibrating string in comparison with nodes in the 
LCAO it molecular orbitals of butadiene. 



Figure 27-4 Interaction of p orbitals in ethene. 



fi 


— 








■■A*' : A 


9: 


"\ /"It 

V w 


■-■'■: 


no 


node, 


bonding 


.!?.. 


node, anti-bonding 



sec 27.5 orbital symmetry and cycloaddition 729 









"tfe 


--i-V 


;+; 


::^v: : 


,->. 


-■'i 


■\ 






••*; 


; '--: 


:>i.* 


3S 


,-;; 


;+>; 


antibonding 


■;.- 


Ys 










-"'-■v' 


'■■ + \. 


C : +-v 


: „^; 


■■.'r-: 


: '*: : 








V 




"J 


v 


ij^ 


f*i' 




II 











+ 


© 


f 


.';■: 


:'':, 




.:■"'- 


bonding 




•+A 


; -'"-? : 


^ r -' 


# 


# 


>| 


■':■;•■ 


<&: 


a*s 


:.'*! 


=:■.-■■.:■" 








■^y : 


§§ 


n 


~ 




CH 2 = 


= CH 2 


CH 2 = 


=CH- 


-CH 


= CH 2 



Figure 27-5 The molecular orbitals of ethene and 1,3-butadiene showing their 
p-orbital precursors. 



The Diels-Alder reaction is quite different. When both 1,3-butadiene and 
ethene are in their ground states, the 1,4 lobes of the highest occupied level 
of the diene match in sign those of the lowest unoccupied level of the alkene 
and thermal reaction is allowed (see Figure 27-5). (This conclusion is also 
reached if the lowest unoccupied level of the diene and highest occupied level 
of the alkene are considered.) Photoexcitation of the reactants, however, de- 



Figure 27-6 Orbital symmetries and cycloaddition pathways for ethene. 



hv 









highest occupied lowest unoccupied 

thermal addition 









mi 



highest occupied lowest unoccupied 

photochemical addition 



chap 27 cyclization reactions 730 

stroys this correspondence and 1,4-addition does not occur. Instead, 1,2 
addition takes place. 




2 ^^ 



^^ 



The 1,2 addition is allowed because the 1,2 lobes of the highest occupied 
level in the excited state of the diene (the first antibonding level) match in 
sign the lobes of the lowest unoccupied level of the alkene (the antibonding 
level, also). 

Many other cycloaddition reactions have been examined recently from the 
point of view of orbital symmetry principles and these principles have been 
found to predict not only the effect of thermal activation and photoactivation 
but also the correct stereochemistry of the addition compounds. Furthermore, 
many cyclization reactions, such as the Diels-AIder reaction, are reversible 
and the same considerations are found to apply to the reverse reactions also. 



summary 

Intramolecular ring closure of bifunctional compounds occurs readily with 
hydroxy or amino acids, with hydroxy aldehydes or ketones, with dicarboxylic 
acids, and with diesters, provided the reacting groups are favorably situated 
with respect to each other. 



HO(CH 2 )„C0 2 H 



H 2 N(CH 2 )„C0 2 H 



HO(CH 2 )„CHO 



H0 2 C(CH 2 )„C0 2 H 




(CH,). C ~ H 




c=o 



summary 731 



R0 2 C(CH 2 )„C0 2 R 




2/.-1 , 

CHCO,R 



(CH 2 )„ <y 
V CHOH 



Formation of five- and six-membered rings is favored, but in the case of 
the conversion of diesters to hydroxy ketones (acyloin condensation) very 
large rings can also be produced. This reaction has enabled a chain-link 
compound called a catenane to be synthesized. 

Alkenes dimerize to cyclobutanes under the influence of light, whereas the 
reaction of an alkene and a conjugated diene to produce a cyclohexene (the 
Diels-Alder reaction) occurs thermally. 





The reaction has the following characteristics: electron-withdrawing sub- 
stituents Z in the alkene (the dienophile) increase the reaction rate; the con- 
figurations of the diene and dienophile are retained in the product showing 
that cis addition occurs; and the reaction occurs via a one-step concerted 
mechanism. 

Photoactivation of a conjugated diene results in preferential formation of 
a four-mernbered ring rather than the six-membered Diels-Alder product. 



hv 



□ 



2 i^^- 




^- 



Alkenes react with 1 ,3-dipolar compounds to form five-membered hetero- 
cyclic rings. 



\ 
+ Y 

o / 



The 1,3-dipolar compounds include ozone (0 3 ), organic azides (RN 3 ), nitrite 
oxides (RCNO), and diazoalkanes (R 2 CN 2 ). 

Fluxional molecules are those that have identical structures and configu- 
rations but that can be distinguished either by means of isotopic labeling or 
by the effect of their interconversion on their nmr spectrum. A number of 



chap 27 cyclization reactions 732 

these systems involve the rearrangement of 1,5-dienes via six-membered 
cyclic transition states. 



This type of reaction is greatly accelerated if the reacting atoms are held in 
favorable positions by a network of bonds such as are present in barbaralane 
[1] or bullvalene [2]. 






[1] PI 

Oxidative coupling of alkynes, followed by base-catalyzed prototropic re- 
arrangement and partial hydrogenation, gives large-ring conjugated poly- 
alkenes called annulenes. 

CH 

n ■ III ► ► ' 

CH 

Alkynes can act as dienophiles in the Diels-Alder reaction and acetylene, 
itself, can be polymerized to give cyclooctatetraene. 

Whether a cycloaddition will be subject to thermal activation or photo- 
activation can be predicted using rules based on orbital symmetry arguments. 
The 7i molecular orbitals of alkenes and alkadienes can be formulated as 
combinations of p orbitals, the lobes of which have either a positive or nega- 
tive sign. Reaction is allowed only between atoms whose lobes have the same 
sign when the highest occupied level of one of the reactants and the lowest 
unoccupied level of the other reactant are considered. 

The allowed reactions are the following: 

(1) Thermal alkene-alkadiene cyclization (Diels-Alder reaction). 



highest occupied lowest unoccupied level 

level of ethene of 1 ,3-butadiene 

(or lowest unoccupied level of ethene + 
highest occupied level of 1 ,3-butadiene) 

(2) Alkene photodimerization. 



■:+:i 



highest occupied lowest unoccupied 

level of the excited level of the ground 

state of ethene state of ethene 

Reactions that are not allowed include thermally activated concerted al- 
kene cyclodimerization and the photoactivated Diels-Alder reaction. When 
the latter reaction is attempted, cyclobutanes are formed. 



exercises 733 

exercises 

27-1 What products would you expect from the Diels- Alder addition of tetra- 
cyanoethene to cis,trans-2,4-hexa.diene and cw,cw-2,4-hexadiene? Explain. 

27-2 Write structures for the products of the following reactions : 

N 
C 6 H sN V, 150° C 6 H 5 CHO 

I C— C 6 H 5 



"V 



\ (C 2 H 5 ) 3 N CS 2 

b. 2 C=N-NHC 6 H 5 _ Ha » 

CI 



X C=N-OH (C2Hs)3N C6HsCN - 



/ -HC1 

CI 

27-3 The rate of the Diels-Alder addition between cyclooctatetraene and tetra- 
cyanoethene is proportional to the tetracyanoethene concentration 
[C 2 (CN) J at low concentrations of the addends but becomes independent 
of [C 2 (CN) 4 ] at high concentrations. Write a mechanism which accounts for 
this behavior. 

27-4 The compound bullvalene, Ci H 10 , has the following structure: 




a. Write several fluxional forms of this molecule (the total number is 
greater than 10 6 ). 

b. At 100° equilibration is rapid and at 0° it is slow. How many different 
kinds of proton would you expect to be seen in a well-resolved nmr 
spectrum at each of these temperatures ? 

27-5 Work out a synthesis of [20]annulene from the coupling product of allyl- 
magnesium bromide with l,4-dibromo-2-butene, which is reported to be 
1,5,9-decatriene. 



chapter 28 J 
polymjers 



chap 28 polymers 737 

Polymers are substances made up of recurring structural units, each of which 
can be regarded as derived from a specific compound called a monomer. The 
number of monomeric units is usually large and variable, a given polymer 
sample being characteristically a mixture of molecules with different molecular 
weights. The range of molecular weights encountered may be either small or 
very large. 

The properties of a polymer, both physical and chemical, are in many ways 
as sensitive to changes in the structure of the monomer as are the properties 
of the monomer itself. This means that to a very considerable degree the 
properties of a polymer can be tailored to particular practical applications. 
We have already considered several methods of synthesis of monomers and 
polymers and mechanisms of polymerization reactions in earlier chapters, 
and much of the emphasis in this chapter will be on how the properties of 
polymers can be related to their structures. 

The thermal polymerization of cyclopentadiene by way of the Diels-Alder 
reaction provides a simple concrete example of how a monomer and a polymer 
are related. 




tetramer polymer 

The first step in this polymerization is formation of the dimer, which in- 
volves cyclopentadiene' s acting as both diene and dienophile. This step occurs 
readily on heating but slowly at room temperature. In subsequent steps, cyclo- 
pentadiene adds to the relatively strained double bonds of the first-formed 
polymer. These additions require higher temperatures (180° to 200°). If 
cyclopentadiene is heated to 200° until substantially no further reaction oc- 
curs, the product is a waxy solid having a degree of polymerization n ranging 
from two to greater than six. 

Polycyclopentadiene molecules have double bonds for end groups and a com- 
plicated backbone of saturated fused rings. The polymerization is reversible, 
and on strong heating the polymer reverts to cyclopentadiene. 



28-1 types of polymers 



Polymers can be classified several different ways : according to their structures, 
the types of reactions by which they are prepared, their physical properties, 



chap 28 polymers 738 




polymer chains as 
in elastomers and 
thermoplastics 



cross links 



Figure 28-1 Schematic representation of a polymer with few cross links. 

or their technological uses. However, these classifications are not all mutually 
exclusive. 

From the standpoint of general physical properties, we recognize three 
types of solid polymers: elastomers (rubbers or rubberlike elastic substances), 
thermoplastic polymers, and thermosetting polymers. These categories overlap 
considerably but are nonetheless helpful in defining general areas of use and 
types of structures. Elastomers (uncured) and thermoplastics typically have 
long polymer chains with few, if any, chemical bonds acting as cross links 
between the chains. This is shown schematically in Figure 28 ■ 1 . 

Such polymers, when heated, normally become soft and more or less fluid 
and can then be molded into useful shapes. The main difference between an 
elastomer and a thermoplastic polymer is in the degree of attractive forces 
between the polymer chains, as discussed in the next section. Thus, although 
elastomers, which are not cross-linked, are normally thermoplastic, not all 
thermoplastics are elastomers. 

Cross links are extremely important in determining physical properties be- 
cause they increase the molecular weight and limit the motion of the chains 
with respect to one another. Only two cross links per polymer chain are re- 
quired to connect together all the polymer molecules in a given sample to 
produce one gigantic molecule. As a result, introduction of only a few cross 
links acts to greatly reduce solubility and tends to produce a gel polymer, 
which, although insoluble, will usually absorb (be swelled by) solvents in 
which the uncross-linked polymer is soluble. The tendency to absorb solvents 
decreases as the degree of cross linking is increased. 

Thermosetting polymers are normally made from relatively low molecular 
weight, usually semifluid substances which, when heated in a mold, become 
highly cross linked, thereby forming hard, infusible, and insoluble products 
having a three-dimensional space network of bonds interconnecting the 
polymer chains (Figure 28-2). 

Figure 28-2 Schematic representation of the conversion of an uncross-linked 
polymer to a highly cross-linked polymer. 




heat 




uncross-linked polymer highly cross-linked polymer 

(heavy lines represent cross links) 



sec 28.2 forces between polymer chains 739 

physical properties of polymers 

28-2 Jorces between polymer chains 

Polymers are produced on an industrial scale primarily, although not exclu- 
sively, for use as structural materials. Their physical properties are particularly 
important in determining their usefulness, be it as rubber tires, sidings for 
buildings, or solid rocket fuels. 

Polymers that are not highly cross-linked have properties that depend upon 
the degree and kind of forces that act between the chains. By way of example, 
consider a polymer such as Polythene (polyethene) which, in a normal commer- 
cial sample, will be made up of molecules having 1000 to 2000 CH 2 groups 
in continuous chains. Since the material is a mixture of different molecules, 
it is not expected to crystallize in a conventional way. 1 Nonetheless, X-ray 
diffraction shows polyethene to have very considerable crystalline character, 
there being regions as large as several hundred angstrom units in length, which 
have ordered, zigzag chains of CH 2 groups oriented with respect to one 
another like the chains in crystalline low-molecular-weight hydrocarbons. 
These crystalline regions are often called crystallites. Between the crystallites 
of polyethene are amorphous, noncrystalline regions in which the polymer 
chains are more randomly ordered with respect to one another (Figure 28 ■ 3). 
These regions essentially constitute crystal defects. 

The forces between the chains in the crystallites of polyethene are the so- 
called van der Waals or dispersion forces, which are the same forces acting 
between hydrocarbon molecules in the liquid and solid states and, to a lesser 
extent, in the vapor state. These forces are relatively weak and arise through 
synchronization of the motions of the electrons in the separate atoms as they 
approach one another. The attractive force that results is rapidly overcome 
by repulsive forces when the atoms get very close to one another. 

In other kinds of polymers, much stronger intermolecular forces can be 
produced by hydrogen bonding. This is especially important in the poly- 
amides, such as the nylons, of which nylon (66) or polyhexamethyleneadip- 
amide is most widely used. 



O H O H 

II I II I 

-N-C-eCH 2 )rC-N-tCH 2 feN-C-eCH 2 )3-C-N-tCH 2 ^ 
H 9 H O 

6 H 6 H 

II I II I 

-N-C-fCH^C-N-fCH^N-C-fCHAC-N-fCH^ 

H p H O 

possible hydrogen-bonded structure for 
crystallites of nylon (66), polyhexamethyleneadipamide 

1 Quite good platelike crystals have been formed from dilute solutions of certain polymers 
such as polyethene. In the crystals, the polymer chains seem to run back and forth in 
folds between the large surfaces of the plates. 



chap 28 polymers 740 

The effect of temperature on the physical properties of polymers is very 
important to their practical uses. At low temperatures, polymers become hard 
and glasslike because the motion of the polymer chains in relation to each 
other is slow. The approximate temperature below which glasslike behavior 
is apparent is called the glass temperature and is symbolized by T g . When a 
polymer containing crystallites is heated, the crystallites ultimately melt; 
this temperature is usually called the melting temperature, symbolized as T m . 
Usually, the molding temperature will be above T m and the mechanical strength 
of the polymer will diminish rapidly as the temperature approaches T m . 

Obviously, another temperature of great importance in the practical use of 
polymers is the temperature near which thermal breakdown of the polymer 
chains occurs. Decomposition temperatures will obviously be sensitive to im- 
purities, such as oxygen, and will be influenced strongly by the presence of 
inhibitors, antioxidants, and so on. Nonetheless, there will be a temperature 
(usually rather high, 200° to 400°) at which uncatalyzed scission of the bonds 
in a chain will take place at an appreciable rate, and, in general, you cannot 
expect to prevent this type of reaction from causing degradation of the poly- 
mer. Clearly, if this degradation temperature is comparable to T m , as it is for 
polyacrylonitrile, difficulties are to be expected in simple thermal molding of 
the plastic. This difficulty is overcome in making polyacrylonitrile (Orion) 
fibers by dissolving the polymer in N,N-dimethylformamide and forcing the 
solution through fine holes into a heated air space where the solvent evaporates. 

Physical properties such as tensile strength, X-ray diffraction pattern, re- 
sistance to plastic flow, softening point, and elasticity of most polymers can 
be understood in a general way in terms of crystallites, amorphous regions, 
the degree of flexibility of the chains, and the strength of the forces acting 
between the chains (dispersion forces, hydrogen bonding, etc.). One way to 
approach the problem is to make a rough classification of properties of solid 
polymers according to the way the chains are disposed in relation to each 
other. 

1 . An amorphous polymer is one with no crystallites. If the forces between 
the chains are weak and if the motions of the chains are not in some way 
severely restricted, such a polymer would be expected to have low tensile 
strength and be subject to plastic flow in which the chains slip by one another. 

2. An unoriented crystalline polymer is one which is considerably crystal- 
lized but has the crystallites essentially randomly oriented with respect to one 
another as in Figure 28-3. Such polymers, when heated, often show rather 
sharp T m points, which correspond to the melting of the crystallites. Above 
T m , these polymers are amorphous and undergo plastic flow, which permits 
them to be molded. Other things being the same, we expect T m to be higher 
for the polymers with stiff chains (high barriers to internal rotation). 

3. An oriented crystalline polymer is one in which the crystallites are ori- 
ented with respect to one another, usually as the result of a cold-drawing 
process. Consider a polymer such as nylon, which has strong intermolecular 
forces and is in an unoriented state like the one represented by Figure 28-3. 
If the material is subjected to strong stress, say along the horizontal axis, at 
some temperature (most easily above T g ) where at least some plastic flow can 



sec 28.2 forces between polymer chains 741 

occur, elongation will take place and the crystallites will be drawn together 
and oriented along the direction of the applied stress (Figure 28-4). 

An oriented crystalline polymer usually has a much higher tensile strength 
than the unoriented polymer. Cold drawing is an important step in the pro- 
duction of synthetic fibers. 

4. Elastomers are intermediate in character between amorphous and crys- 
talline polymers. The key to elastic behavior is to have a polymer that has 
either sufficiently weak forces between the chains or a sufficiently irregular 
structure to be very largely amorphous. The tendency for the chains to orient 
can often be considerably reduced by random introduction of methyl groups 
which, by steric hindrance, inhibit ordering of the chains. An elastomer needs 
to have some crystalline (or cross-linked) regions to prevent plastic flow and, 
in addition, should have rather flexible chains (which means T g should be low). 
The structure of a polymer of this kind is shown schematically in Figure 28 • 5. 
The important difference between this elastomer and the crystalline polymer 
of Figure 28 • 3 is the size of the amorphous regions. When tension is applied 
and the material elongates, the chains in the amorphous regions straighten 



Figure 28-3 Schematic diagram of crystallites (enclosed by dotted lines) in a 
largely crystalline polymer. 




chap 28 polymers 742 




Figure 28-4 Schematic representation of an oriented crystalline polymer 
produced by drawing in the horizontal direction. The crystalline regions are 
enclosed with dotted lines. 



out and become more nearly parallel. At the elastic limit, a semicrystalline 
state is reached, which is different from the one produced by cold drawing a 
crystalline polymer in that it is stable only while under tension. The forces be- 
tween the chains are too weak in the absence of tension to maintain the crys- 
talline state. Thus, when tension is released, contraction occurs and the 
original, nearly amorphous, polymer is produced. 

Probably the best known elastomer is natural rubber, which is a polymer 
of 2-methyl-l,3-butadiene (isoprene) with virtually all the double bonds in 
the cis configuration. 



CH, 



CH 2 =C-CH=CH 2 

isoprene 




natural rubber (c«-polyisoprene) 



Figure 28'S Schematic representation of an elastomer in relaxed and stretched 
configurations. The crystalline regions are enclosed by dotted lines. 




stretch 



relax 




largely amorphous 
polymer 



largely crystalline 
polymer 



sec 28.3 correlation of polymer properties with structure 743 

When pure isoprene is polymerized in the laboratory with the help of 
radical initiators the product is hard and brittle, quite unlike natural rubber. 
This material is similar to a naturally occurring substance known as gutta 
percha that is used to make covers for golf balls. It owes its properties to 
its trans arrangement of double bonds which allow the chains to lie alongside 
one another in a semicrystalline array. 




When the double bonds are cis, as in natural rubber, steric hindrance keeps 
the chains from assuming a similar ordered structure and the bulk of the 
material exists in an amorphous state with randomly oriented chains. When 
the cis polymer is stretched, the chains are straightened and tend to become 
oriented; but since this is an unfavorable state, the material snaps back to the 
amorphous state when released. Although radical polymerization of isoprene 
produces gutta percha, Ziegler polymerization (Section 28-5A) gives a cis 
polymer that is identical with natural rubber. The correlation between poly- 
mer structure and properties are examined further in the next section. 

A good elastomer should not undergo plastic flow in either the stretched or 
relaxed state, and when stretched should have a "memory" of its relaxed 
state. These conditions are best achieved with natural rubber (cw-polyiso- 
prene) by curing (vulcanizing) with sulfur. Natural rubber is tacky and under- 
goes plastic flow rather readily, but when heated with 1 to 8 % by weight of 
elemental sulfur in the presence of an accelerator, sulfur cross links are intro- 
duced between the chains. These cross links reduce plastic flow and provide 
a sort of reference framework for the stretched polymer to return to when it is 
allowed to relax. Too much sulfur completely destroys the elastic properties 
and gives hard rubber of the kind used in cases for storage batteries. 

28-3 correlation of polymer properties 
with structure 

With the aid of the concepts developed in the previous section, it is possible 
to correlate the properties of many of the technically important thermoplastic 
and elastic polymers with their chemical structures. We can understand why 
the simple linear polymers such as Polythene (polyethene), polyformaldehyde, 
and Teflon (polytetrafluoroethene) are crystalline polymers with rather high 
melting points. Polyvinyl chloride, polyvinyl fluoride, and polystyrene as 
usually prepared are much less crystalline and have lower melting points; 
with these polymers, the stereochemical configuration is very important in 



chap 28 polymers 744 

determining the physical properties. Polystyrene, made by radical polymeri- 
zation in solution, is atactic. This means that, if we orient the carbons in the 
polymer chain in the form of a regular zigzag, the phenyl groups will be ran- 
domly distributed on one side or the other when we look along the chain, as 
shown in Figure 28-6. Polymerization of styrene with Ziegler catalysts 
(Sections 4-4H and 28 -5 A) produces isotactic polystyrene, which is different 
from the atactic polymer in that all of the phenyl groups are located on one 
side of the chain. The difference in properties between the atactic and isotactic 
materials is considerable. The atactic polymer can be molded at much lower 
temperatures and is much more soluble in most solvents than is the isotactic 
product. There are many other possible types of stereoregular polymers, one 
of which is called syndiotactic and has the side-chain groups oriented alter- 
nately on one side and then the other, as shown in Figure 28 ■ 6. 

Polypropene, made by polymerization of propene with Ziegler catalysts, 
appears to be isotactic and highly crystalline with a melting point of 175°. It 
can be drawn into fibers that resemble nylon fibers although, as might be 
expected, they do not match the 270° melting point of nylon and are much 
more difficult to dye. 



Figure 28-6 Configurations of atactic, isotactic, and syndiotactic poly- 
styrene. The conformations are drawn to show the stereochemical relations of 
the substituent groups and are not meant to represent necessarily the stable 
conformations of the polymer chains. 






O^p?" 0^^ H h^=o 

0-^i H C^a^ h 0^d^ H 





Table 28-1 Representative synthetic thermoplastic and elastic polymers and their uses" 



monomer(s) 


formula 


type of 
polymerization 


physical 
type 


°c 


Tm, 

°c 


trade 
names 


uses 


ethene 


CH 2 =CH 2 


radical 
(high pressure) 


semi- 
crystalline 


^0 


110 


Polythene 
Alathon 


film, containers, 
piping, etc. 






Ziegler 


crystalline 


-120 


130 






vinyl chloride 


CH 2 =CHC1 


radical 


atactic, 
semi- 
crystalline 


80 


180 


polyvinyl 
chloride, 
Geon 


film, insulation, 
piping, etc. 


vinyl fluoride 


CH 2 =CHF 


radical 


atactic, 
semi- 
crystalline 


45 




Tedlar 


coatings 6 


vinyl chloride 
vinylidene chloride 


CH 2 =CHC1 
CH 2 =CC1 2 


radical 


crystalline 


variable 




Saran 


tubing, fibers, film, 
structural materials 


chlorotrifluoroethene 


CF 2 =CFC1 


radical 


atactic, 
semi- 
crystalline 


<o 


210 


Kel-F 


gaskets, insulation" 


tetrafluoroethene 


CF 2 =CF 2 


radical 


crystalline 


<-100 


330 


Teflon 


gaskets, valves, insu- 
lation, filter felts, 
coatings'" 


propene 


CH 2 =CHCH 3 


Ziegler 


isotactic, 
crystalline 


-20 


175 




fibers, molded 
articles 


hexafluoropropene 
vinylidene fluoride 


CF 2 =CFCF 3 
CH 2 =CF 2 


radical 


amorphous 


-23 




Viton 


rubber articles 



Table 284 Representative synthetic thermoplastic and elastic polymers and their uses" {continued) 



monomer(s) 


formula 


type of 
polymerization 


physical 
type 


°C 


Tm, 

°C 


trade 
names 


uses 


2-methylpropene 


CH 2 =C(CH 3 ) 2 


cationic 


amorphous 


-70 




Vistanex, 
Oppanol 


pressure-sensitive 
adhesives 


2-methylpropene 


CH 2 =C(CH 3 ) 2 


cationic 


amorphous 






butyl 
rubber 


inner tubes 


chloroprene 


CH 2 =C(C1)CH=CH 2 


radical 


amorphous 


-40 




Neoprene 


rubber articles 6 


isoprene 


CH 2 =C(CH 3 )CH=CH 2 


Ziegler, Li 


amorphous 
(cw-1,4) 


-70 


28 


natural 
rubber, 
Ameripol, 
Coral 
rubber 


rubber articles, 


styrene 


CH 2 =CHC 6 H 5 


radical 


atactic, 
semi- 
crystalline 


85 


<200 


Styron, 
Lustron 


molded articles,- 
foam 


styrene 


C-H.2 ^^CriCetls 


Ziegler 


isotactic 


100 


230 






vinyl acetate 


CH 2 =CH0 2 CCH 3 


radical 


amorphous 


40 




polyvinyl 
acetate 


adhesives 


vinyl alcohol 


(CH 2 =CHOHK 


hydrolysis of 
polyvinyl 
acetate 


crystalline 




dec. 


polyvinyl 
alcohol 


water-soluble adhe- 
sives, paper sizing 


vinyl butyral 


X C3H 7 ' 


polyvinyl alcohol 
and butanal 


amorphous 






polyvinyl 
butyral 


safety-glass laminate 



Table 28-1 (continued) 



monomer(s) 


formula 


type of 
polymerization 


physical 
type 


°c 


T m , 

°C 


trade 
names 


uses 


formaldehyde 


CH 2 =0 


anionic 


crystalline 




179 


Delrin 


molded articles 


acrylonitrile 


CH 2 =CHCN 


radical 


crystalline 


100 9 


>200 


Orion 


fiber 


methyl methacrylate 


CH 2 =C(CH 3 )C0 2 CH 3 


radical 


atactic, 
amorphous 


105 




Lucite, 
Plexiglas 


coatings, molded 
articles 






anionic 


isotactic, 
crystalline 


115 


200 








/=\ 


anionic 


syndiotactic, 
crystalline 


45 


160 






ethylene glycol tereph- 
thalate 


H0 2 C^ ^>C0 2 C 2 H 4 OH 


ester interchange 
between dimethyl 
terephthalate and 
ethylene glycol 


crystalline 


56 


260 


Dacron, 

Mylar, 

Cronar, 

Terylene 


fiber, film 


hexamethylenediamine 
and hexanedioic acid 
(adipic acid) 


NH 2 (CH 2 ) 6 NH 2 
H0 2 C(CH 2 ) 4 C0 2 H 


anionic 
condensation 


crystalline 


50 


270 


nylon, 

Zytel 


fibers, molded 
articles 



" Much useful information on these and related polymers is given by F. W. Billmeyer, Jr., A Textbook of Polymer Chemistry, Interscience, New York, 1957; J. K. Stille, 
Introduction to Polymer Chemistry, Wiley, New York, 1962; F. Buecne, Physical Properties of Polymers, Interscience, New York, 1962; and W. R. Sorenson and T. W. 
Campbell, Preparative Methods of Polymer Chemistry, Interscience, New York, 1961. 

6 Exceptional outdoor durability. 

c Used where chemical resistance is important. 

d Excellent self-lubricating and electrical properties. 

e Used particularly where ozone resistance is important. 

f These monomers are not the starting materials used to make the polymers, which are actually synthesized from polyvinyl acetate. 

9 T g is 60° when water is present. 



•0 


B 



o 

•0 
a 



chap 28 polymers 748 

Although both linear polyethene and isotactic polypropene are crystalline 
polymers, ethene-propene copolymers prepared with the aid of Ziegler cata- 
lysts are excellent elastomers. Apparently, a more or less random introduction 
of methyl groups along a polyethene chain reduces the crystallinity sufficiently 
drastically to lead to a largely amorphous polymer. 

Polyvinyl chloride, as usually prepared, is atactic and not very crystalline. 
It is relatively brittle and glassy. The properties of polyvinyl chloride can be 
improved by copolymerization, as with vinyl acetate, which produces a softer 
polymer (Vinylite) with better molding properties. Polyvinyl chloride can 
also be plasticized by blending it with substances of low volatility such as 
tricresyl phosphate and di-«-butyl phthalate, which, when dissolved in the 
polymer, tend to break down its glasslike structure. Plasticized polyvinyl 
chloride is reasonably elastic and is widely used as electrical insulation, plastic 
sheeting, and so on. 

Table 28 • 1 contains information about a number of representative impor- 
tant polymers and their uses. 



preparation of synthetic polymers 

A prevalent but erroneous notion has it that useful polymers, such as those 
given in Table 28-1, can be, and are, made by slap-dash procedures applied 
to impure starting materials. In actual fact the monomers used in most large- 
scale polymerizations are among the purest known organic substances. 
Furthermore, to obtain uniform, commercially useful products, extraordinary 
care must be used in controlling the polymerization reactions. The reasons are 
simple — namely, that formation of a high-molecular-weight polymer (high 
polymer) requires a reaction that proceeds in very high yields, and purification 
of the product by distillation, crystallization, and so on, is difficult, if not 
impossible. Even a minute contribution of any side reaction that stops poly- 
mer chains from growing will seriously affect the yield of high polymer. 

In this section, we shall discuss some of the more useful procedures for the 
preparation of high polymers, starting with examples involving condensation 
reactions. 



28-4 condensation polymers 



There is a very wide variety of condensation reactions 2 that, in principle, 
can be used to form high polymers. However, as explained above, high poly- 
mers can only be obtained in high-yield reactions, and this limitation severely 
restricts the number of condensation reactions having any practical import- 
ance. A specific example of an impractical reaction is the formation of poly- 
tetramethylene glycol by reaction of tetramethylene bromide with the sodium 
salt of the glycol. 

2 A condensation reaction is usually taken to mean one in which two molecules react to 
split out water or some other simple molecule. 



sec 28.4 condensation polymers 749 



NaO-fCH^ONa + Br(CH 2 ) 4 Br 



-fO-f CH^OJr + Na Br 



It is unlikely that this reaction would give useful yields of any very high 
polymer because E2 elimination, involving the dibromide, would give a 
double bond end group and prevent the chain from growing. 



A. POLYESTERS 



A variety of polyester-condensation polymers are made commercially. Ester 
interchange (Section 13-8) appears to be the most useful reaction for prepa- 
ration of linear polymers (see Figure 28-7). 



Figure 28-7 Reactions used to prepare the polymers Dacron and Lexan. 



CH,0 2 C 



a y-C0 2 CH 3 + HOCH 2 CH 2 OH 
dimethyl terephthalate ethylene glycol 



-200° 



O 



metal oxide catalyst 



O-C-4 >-C-0-CH 2 -CH 2 + + CH 3 OH 

polyethylene glycol terephthalate 
(Dacron) 



O 

/=\ II 



HO 



CH, 



CH 3 

bisphenol A 



OH + (C 6 H<0),CO 



diphenyl 
carbonate 



300° 



CH 3 O 

°~\ //~°\ \ / °~ C+ " + CfiH5 ° H 
CH 3 

polybisphenol A carbonate 
(Lexan) 



chap 28 polymers 7S0 



-0-CH 2 . CH 2 - 

CrK 

1 


-o- 


o 

II II 
■ C M C "°- 


CH 2 . CH /CH 2 - 

1 


-0- 




II II 

-C C- 


1 
O 

| 




\J 


o 

1 




O 


^f C=0 




r 


^c=o 






^x=o 




L 


\^C=0 






1 



1 

CH 
-O-CHr ^CH 2 - 


-o- 




-c c-o 

II II 


1 


CH 

CH2 CH2"" 


-0- 


Q 

-C c— 

II II 







o o 







Figure 28-8 Glyptal resin. 



Thermosetting space-network polymers are often prepared through the 
reaction of polybasic acid anhydrides with polyhydric alcohols. A linear poly- 
mer is obtained with a bifunctional anhydride and a bifunctional alcohol, but 
if either reactant has three or more reactive sites, then formation of a three- 
dimensional polymer is possible. For example, two moles of glycerol can 
react with three moles of phthalic anhydride to give a highly cross-linked resin, 
which is usually called a glyptal (Figure 28 • 8). 

B. NYLONS 

A variety of polyamides can be made by heating diamines with dicarboxylic 
acids. The most generally useful of these is nylon (66), the designation (66) 
arising from the fact that it is made from the six-carbon diamine hexamethyl- 
enediamine, and the six-carbon dicarboxylic acid, hexanedioic acid (adipic 
acid). 



H0 2 C(CH 2 ) 4 C0 2 H + NH 2 (CH 2 ) 6 NH 2 



280° 



o o 

II II H H 

C(CH 2 ) 4 C-N-fCH 2 )s-NH- + H,0 



The polymer can be converted into fibers by extruding it above its melting 
point through spinnerettes, then cooling and drawing the resulting filaments. 
It is also used to make molded articles. Nylon (66) is exceptionally strong and 
abrasion resistant. 

The starting materials for nylon (66) manufacture can be made in many 
ways. Apparently, the best route to adipic acid is by air oxidation of cyclo- 
hexane by way of cyclohexanone. 



sec 28.4 condensation polymers 75 1 



O 

H 2 II 

XL ,C^ C0 2 H 

HX" "CH 2 ^^ H 2 C- -CH 2 ^o^ / 2 

I I — — 77* I I » (<-H 2 ) 4 

H 2 C^ c /CH 2 - H ^° H 2 C^ C /CH 2 \ CQ2H 



Hexamethylenediamine is prepared from the addition product of chlorine to 
butadiene (Section 6 • 2) by the following steps : 



Ci 2 CH 2 -CH=CH-CH 2 + CH 2 =CH-CH-CH 2 

CH 2 =CHCH=CH 2 <• I I -II 

CI CI CI CI 

2 NaCN H 2 

► NCCH,CH=CHCH 2 CN ► H 2 N(CH 2 ) 6 NH 2 

—2 Nad " metal 

catalyst 



Both the 1,2- and 1,4-chlorine addition products give the same dinitrile with 
sodium cyanide. 
Nylon (6) is obtained by the polymerization of £-caprolactam. 

.C-c ° ° 

H 2 C \ A II II 

2 | NH ^— -CH 2 -C-NH-CH 2 -CH 2 -CH 2 -CH 2 -CH 2 -C-NH- 

h >%Jh 2 

H 2 



Note that here the intramolecular interaction between amino and carboxyl 
groups (lactam formation) is replaced by intermolecular interaction (poly- 
amide formation). e-Caprolactam is prepared by the Beckmann rearrange- 
ment (Section 16- 1E2) of cyclohexanone oxime, which can be made, in turn, 
from cyclohexanone. 




C. PHENOL-FORMALDEHYDE (BAKELITE) RESINS 

One of the oldest-known thermosetting synthetic polymers is made by con- 
densation of phenol with formaldehyde using basic catalysts. The resins that 
are formed are known as Bakelites. The initial stage in the base-induced re- 
action of phenol and formaldehyde yields a hydroxybenzyl alcohol. This part 



chap 28 polymers 752 

of the reaction closely resembles an aldol addition and can take place at either 
an ortho or the para position. 




0=0" 

\=/CH 2 -O e 



e / \ 

O-f VCH,OH 



The next step in the condensation is formation of a dihydroxydiphenyl- 
methane derivative which for convenience is here taken to be the 4,4' isomer. 



ho-*;, ^-ch 2 oh + ^ /)~ OH H2Q> HO 



\ /r CH2 ~x / 



OH 



This reaction is likely to be an addition to a base-induced dehydration 
product of the hydroxybenzyl alcohol. 



^)l CH ^ H ~ ° 



°J<q> 



CH 2 + i ^O 



H«= 



-> o- 



CH 2 

CH, V = / 



* HO-4 ,y-CH 2 -l >~OH 



Continuation of these reactions to all of the available ortho and para posi- 
tions of the phenol leads to a cross-linked three-dimensional polymer 
(Figure 28-9). 



Figure 28-9 Phenol-formaldehyde resin. 




sec 28.5 addition polymers 753 

28-5 addition polymers 

We have already discussed the synthesis and properties of a considerable 
number of addition polymers in this and earlier chapters. Our primary con- 
cern here will be with some aspects of the mechanism of addition polymeriza- 
tion that influence the character of the polymers formed. 

A. VINYL POLYMERIZATION 

The most important type of addition polymerization is that of the simple 
vinyl monomers such as ethene, propene, styrene, and so on. In general, we 
now recognize four basic kinds of polymerization of vinyl monomers — radical, 
cationic, anionic, and coordination. The elements of the mechanisms of the 
first three of these have been outlined earlier (Section 4-4H). The possibility, 
in fact the reality, of a fourth mechanism is forced on us by the discovery of 
the Ziegler and other (mostly heterogeneous) catalysts, which apparently do 
not involve "free" radicals, cations or anions, and which can and usually 
do lead to highly stereoregular polymers. Although a great deal of work has 
been done on the mechanism of coordination polymerization, the details of 
how each unit of monomer is added to the growing chains is mostly conjecture. 
With titanium-aluminum catalysts, the growing chain probably has a C — Ti 
bond; further monomer units are then added to the growing chain by co- 
ordination with titanium, followed by an intramolecular rearrangement to 
give a new growing-chain end and a new vacant site on titanium where a new 
molecule of monomer can coordinate. 



X -^^ /C H 2 

— CH 2 — CH 2 — CH 2 — CH 2 '' ' HjCX"'' 

— CH 2 -CH 2 — CH 2 — CH 2 

► — CH 2 -CH 2 -CH 2 -CH 2 -CH 2 -CH 2 

Ti'' 



In the coordination of the monomer with the titanium, the metal is probably 
behaving as an electrophilic agent and the growing-chain end may well be 
transferred to the monomer as an anion. Since this mechanism gives no ex- 
plicit role to the aluminum, it is surely a considerable oversimplification. 
Ziegler catalysts polymerize most monomers of the type RCH=CH 2 , pro- 
vided the R group is one that does not react with the organometallic com- 
pounds present in the catalyst. 

B. RADICAL POLYMERIZATION 

In contrast to coordination polymerization, formation of vinyl polymers by 
radical chain mechanisms is reasonably well understood — at least for the 
kinds of procedures used on a laboratory scale. The first step in the reaction 
is the production of radicals; this can be achieved in a number of different 



chap 28 polymers 754 

ways, the most common being the thermal decomposition of an initiator, 
usually a peroxide or an azo compound. 

N — ' o— o ^-^ o 

benzoyl peroxide 

CHi CHi CHi 

| | | 

CH 3 -C— N=N— C— CH 3 4 °°' 80 °> 2 CH 3 -C- + N 2 

CN CN CN 

2,2'-azobis(2-methylpropanonitri!e) 

Many polymerizations are carried out on aqueous emulsions of monomers. 
For these, water-soluble inorganic peroxides, such as persulfuric acid, are 
often employed. 

Addition of the initiator radicals to monomer produces a growing-chain 
radical which combines with successive molecules of monomer until, in some 
way, the chain is terminated. It will be seen that addition to an unsymmetrical 
monomer, such as styrene, can occur in two ways. 



-* 2X- initiation 

>-CH-CH 2 X 



/ — > V_/ 



X ' + \ /-CH=CH, 



\ /-CH-CH 2 



All evidence on the addition of radicals to styrene indicates that the process 
by which X- adds to the CH 2 end of the double bond is greatly favored over 
addition at the CH end. This direction of addition is in accord with the con- 
siderable stabilization of benzyl-type radicals relative to alkyl-type radicals 
(see Section 24-2). Polymerization will then result in the addition of styrene 
units to give phenyl groups only on alternate carbons ("head-to-tail" 
addition). 

CsHU CeHs C 6 Hs 

I II 

X-CH 2 -CH- ► X-CH 2 -CH-CH 2 -CH- 



> X-CH 2 -CH-CH 2 -CH-CH 2 -CH- 

I I I 

C6H5 CeHs CeHs 

In general, we predict that the direction of addition of an unsymmetrical 
monomer will be such as to give always the most stable growing-chain radical. 

The process of addition of monomer units to the growing chain can be 
interrupted in different ways. One is chain termination by combination or dis- 
proportionation of radicals. Explicitly, two growing-chain radicals can com- 



sec 28. S addition polymers 755 

bine with formation of a carbon-carbon bond, or disproportionation can 
occur with a hydrogen atom being transferred from one chain to the other. 



X-\-CH 2 -CH^CH 2 -CH- + -CH-CH^CH-CHj^X 

/ C,,H 5 \ C 6 H 5 C„H 3 /C 6 H 5 \ 

4 r CH 2 -CH-/CH 2 -CH-CH-CH 2 -\CH-CH 27 [x 



combination 
. > X-r-CH 



C„H 5 \ C„H 5 C 6 H 5 /c b H 5 



H 5 \ C„H 5 C 6 H 5 /■ 

I-£cH = CH + CH 2 -CH 2 -\ 



-b"5 

-* X-VCH 2 -CH-/CH = CH + CH 2 -CH 2 -VCH-CH 2 ^X 



disproportionation 

The disproportionation reaction is the radical equivalent of the E2 reaction. 



: C>Vc— + ;C— ► c = c + h — c- 



Which mode of termination occurs can be determined by measuring the 
number of initiator fragments per polymer molecule. If there are two initiator 
fragments in each molecule, termination must have occurred by combination. 
One initiator fragment per molecule indicates disproportionation. Apparently 
styrene terminates by combination; but, with methyl methacrylate, both 
reactions take place, disproportionation being favored. 

C. CATIONIC AND ANIONIC POLYMERIZATION 

Polymerization of alkenes by the cationic mechanism is most important for 
2-methylpropene and a-methylstyrene, which do not polymerize well by other 
methods, and was discussed earlier in considerable detail (Section 4-4H). 

In general, we expect that anionic polymerization (Section 4 -4H) will occur 
when the monomer carries substituents that will tend to stabilize the anion 
formed when a basic initiator, such as amide ion, adds to the double bond of 
the monomer. Cyano and carbalkoxy groups are favorable in this respect and 

R R 

e / P./ 

H 2 N: + CH 2 =C ► H 2 N— CH 2 — C 

X H H 

it is reported that acrylonitrile and methyl methacrylate can be polymerized 
with sodium amide in liquid ammonia. Styrene and isoprene undergo anionic 
polymerization under the influence of powerful bases such as butyllithium 
and phenylsodium. 

Ethylene oxide reacts readily with aqueous hydroxide ion to give either 
ethylene glycol or polymers of various chain length, depending on the quantity 
of water present. 



chap 28 polymers 756 



H 2 



HOCH 2 CH 2 OH 

ethylene glycol 



CH 2 

I \)^ 
I / 
CH 2 

deficient 
—^5 > HOCH 2 (CH 2 OCH 2 )„CH 2 OH 

polyethylene glycol 

Polyethylene glycol polymers can be viscous liquids or waxy solids (Carbo- 
wax, Section 10-11) depending on the molecular weight, and all are water 
soluble. This property makes them valuable for preserving archeological 
relics. Water-logged wooden objects that would warp or disintegrate on being 
dried can be repeatedly saturated with polyethylene glycol, thus removing the 
water and adding a permanent filler simultaneously. The Swedish warship 
Vasa recently raised from the bottom of Stockholm harbor, where it lay for 
over three centuries, is being preserved in this way. 



D. COPOLYMERS 

When polymerization occurs in a mixture of monomers, there will be some 
competition between the different kinds of monomers to add to the growing 
chain and produce a copolymer. Such a polymer will be expected to have 
quite different physical properties than a mixture of the separate homopoly- 
mers. Many copolymers, such as butadiene-styrene, ethene-propene, Viton 
rubbers, and vinyl chloride-vinyl acetate plastics are of considerable commer- 
cial importance. 



28-6 naturally occurring polymers 

There are a number of naturally occurring polymeric substances that have a 
high degree of technical or biological importance. Some of these, such as 
natural rubber, cellulose, and starch have regular structures and can be re- 
garded as being made up of single monomer units. Others such as wool, silk, 
and deoxyribonucleic acid are copolymers. We have considered the chemistry 
of most of these substances in some detail earlier and we shall confine our 
attention here to silk, wool, and collagen. 



A. SILK 

Silk fibroin is a relatively simple polypeptide, the composition of which varies 
according to the larva by which it is produced. The commercial product, 
obtained from the cocoons of mulberry silk moths, contains glycine, L-ala- 
nine, L-serine, and L-tyrosine as its principal amino acids. Silk fibroin has 
an oriented-crystalline structure. The polypeptide chains occur in sheets, each 
chain with an extended configuration parallel to the fiber axis (not the a helix, 
Figure 17-7), and hydrogen bonded to two others in which the directions of 
the peptide chain are reversed (Figure 28-10). 



sec 28.6 naturally occurring polymers 757 



/ \ / \ 

"0=C N-H -0=C N— H— 

CHR RHC CHR RH^ 

/ \ / \ 

-H— N C=0-H-N C = 0- 

\ / \ / 

C=0-H-N C=0~-H-N 

RHC CHR RHC 7 X CHR 

\ / \ / 

N-H— 0=C N-H— 0=C 

/ \ / \ 

-0=C N-H-0=C N-H- 

CHR RHC N CHR RHC X 

/ \ / \ 

--H-N C=0-H— N C=0- 

\ / \ / 

^ C =0-H-N C =o-H-N 

RHC CHR RHC X X CHR 

\ / \ / 

N— H— 0=C N-H — 0=C 

/ \ / \ 



Figure 28-10 Hydrogen-bonded structure of silk fibroin. Note that the 
peptides run in different directions in alternate chains. 



B. WOOL 

The structure of wool is more complicated than that of silk fibroin, because 
wool, like insulin (Section 17-4), contains a considerable quantity of cystine 
(Table 17 • 1), which provides disulfide cross links between the peptide chains. 
These disulfide linkages play an important part in determining the mechanical 
properties of wool fibers because, if the disulfide linkages are reduced, as with 
ammonium thioglycolate solution, the fibers become much more pliable. 

RCH 2 -S-S-CH 2 R + 2 HSCH 2 CC- 2 NH 4 ► RCH 2 SH + HSCH 2 R 

(wool disulfide cross link) 

+ NH 4 2 CCH 2 -S-S-CH 2 C0 2 NH 4 

Advantage is taken of this in the curling of hair, thioglycolate reduction 
being followed by restoration of the disulfide linkages through treatment with 
a mild oxidizing agent while the hair is held in the desired, curled position. 



C. COLLAGEN 

The principal protein of skin and connective tissue is called collagen and is 
primarily constituted of glycine, proline, and hydroxyproline. Collagen mole- 
cules are very long and thin (14 A x 2900 A), and each appears to be made up 
of three twisted polypeptide strands. When collagen is boiled with water, the 
strands come apart and the product is ordinary cooking gelatin. Connective 
tissue and skin are made up of fibrils, 200 to 1000 A wide, which are indicated 
by X-ray diffraction photographs to be composed of collagen molecules 
running parallel to the long axis. Electron micrographs show regular bands, 
about 700 A apart, across the fibrils. It is believed that these correspond to 



chap 28 polymers 758 



700 A 
| -*-l h«- 1 !— — 2900 A H 























Figure 28-1 1 Schematic diagram of collagen molecules in a fibril so arranged 
as to give the 700-A spacing visible in electron micrographs. 



collagen molecules, all heading in the same direction but regularly staggered 
by about a fourth of their length (Figure 28 • 1 1). 

The conversion of collagen fibrils to leather presumably involves formation 
of cross links between the collagen molecules. Various substances can be used 
for the purpose, but chromium salts act particularly rapidly. 



28-7 dyeing of Jibrous polymers 



Many fibrous polymers from synthetic and natural sources are used in the 
manufacture of fabrics, and the need for a variety of methods and a variety 
of dyes should be clear from the chemical diversity of these polymers. At the 
nonpolar extreme are substances such as polypropene, a long-chain hydro- 
carbon; in the middle is cotton, a polyglucoside with ether and hydroxyl 
linkages; at the polar end is wool, a polypeptide structure, cross-linked by 
cystine and containing free acid and amino groups. 

In virtually all dyeing processes the dye must do more than color the surface 
of the fiber. It must also penetrate the fiber and not be removed during wash- 
ing and cleaning operations. Thus, a water-soluble color applied directly to 
medium- or non-polar fibers normally is poorly wash-fast, and some stratagem 
has to be developed to keep it in the fiber. Some of the methods of producing 
wash-fast dyes follow. 



A. DYES WITH POLAR GROUPS 



Substitution of polar groups such as amino and sulfonic acid groups into 
colored molecules often improves wash-fastness by enabling the dye to com- 
bine with polar sites in the fiber. This is a particularly useful technique with 
wool and silk, both of which are polypeptides and contain many strongly 



sec 28.7 dyeing of fibrous polymers 759 

polar groups. Martius Yellow (Section 26-2), which is strongly acidic, is a 
simple direct dye for wool and silk. For cotton, linen, and rayon, which are 
cellulose fibers, it is more difficult to achieve wash-fast colors by direct dyeing. 
Congo Red was the first reasonably satisfactory direct dye for cotton. It has 
polar amine and sulfonate groups which, in the fiber, can form hydrogen 
bonds to the cellulose ether and hydroxyl groups and to other dye molecules, 
thus reducing its tendency to be leached out in washing. 




Congo Red 
B. DISPERSE DYES 

The use of water-insoluble, fiber-soluble (" disperse ") dyes is helpful for many 
of the medium- and less-polar fibers. Such dyes usually give true solutions in 
the fiber — the absorption of the dye not being dependent on combination 
with a limited number of polar sites. Disperse dyes are usually applied in the 
form of a dispersion of finely divided dye in a soap solution in the presence 
of some solubilizing agent such as phenol, cresol, or benzoic acid. The process 
suffers from the fact that usually the absorption of dye in the fiber is slow and 
is best carried out at elevated temperatures in pressure vessels. 

l-Amino-4-hydroxyanthraquinone is a typical dye which can be used in 
dispersed form to color Dacron (polyethylene glycol terephthalate). Absorp- 
tion of this dye is a solution process, as indicated by the fact that, even up to 
high dye concentrations in the fiber, the amount of dye in the fiber is directly 
proportional to the equilibrium concentration of dye in the solution. 




OH 



l-arnino-4-hydroxyanthraquinone 
(red-violet solid, used to dye fabrics pink) 



C. MORDANT DYES 

One of the oldest known methods of producing wash-fast colors is with the 
aid of metallic hydroxides to form a link between the fabric and the dye. The 
production of cloth dyed with "Turkey red," the coloring material of the 
root of the madder plant, using aluminum hydroxide as a binder or " mordant," 
has been carried out for many centuries. The principal organic ingredient of 



chap 28 polymers 760 

Turkey red has been shown to be 1,2-dihydroxyanthraquinone ("alizarin") 
and this substance is now prepared synthetically from anthraquinone. 




1 ,2-dihydroxyanthraquinone 
(alizarin) 

Mordant dyes are useful on cotton, wool, or silk, and are applied in a 
rather complicated sequence of operations whereby the cloth is treated with 
a solution of a metallic salt in the presence of mild alkali and a wetting agent 
for the purpose of forming a complex of the fiber with the metal cation. The 
dye is then introduced and an insoluble complex salt (often called a "lake") 
is formed in the fiber. With alizarin and aluminum hydroxide, it is probable 
that the binding to the dye involves salt formation at the 1-hydroxyl and 
coordination to aluminum at the adjacent carbonyl group. 

fiber 



X)H 



Apparently, this chelated type of structure is important in contributing to 
the excellent light-fastness of most mordant dyes. 

A variety of metals can be used as mordants, but aluminum, iron, and 
chromium are most commonly used. Mordant dyes normally have reasonably 
acidic phenolic groups and some kind of an adjacent complexing group which 
fills the function of the carbonyl group in alizarin. 



D. VAT DYES 

Another and very effective way of making fast colors is to introduce the dye 
in a soluble form (which may itself be colorless) and then generate the dye in 
an insoluble form within the fiber. Most commonly, the soluble form of the 
dye is a reduced form, the dye being produced by oxidation. The overall 
process is known as "vat dyeing," the name arising from the vats used in the 
reduction step. 

The famous dyes of the ancients, indigo and Tyrian purple (royal purple), 
can be applied this way; reduced, soluble forms of these dyes occur naturally. 
In the case of indigo, this is a glycoside, indican, which occurs in the indigo 



sec 28.7 dyeing of fibrous polymers 761 

plant. Enzymic or acid hydrolysis of indican gives 3-hydroxyindole (" in- 
doxyl"), which exists in equilibrium with the corresponding keto form. 




indoxyl 




CH, 



Air oxidation of indoxyl produces indigo probably by a radical mech- 
anism (Figure 28-12). 



Figure 28-12 Preparation of indigo from indoxyl. 



leucoindigo (indigo white) 




indigo 




chap 28 polymers 762 

The last stage of this reaction, the oxidation of leucoindigo, will be seen to 
resemble conversion of hydroquinone to quinone. X-Ray studies have shown 
that indigo has the trans configuration of the double bond. Indigo is very 
insoluble in water and most organic solvents. It absorbs strongly at 5900 A. 

In the ordinary dyeing process, indigo is reduced to the colorless leuco form 
which, as an enol, is soluble in alkaline solution and is applied to fabric in 
this form. That alkaline solutions are required for solubilization of the leuco 
form of most vat dyes restricts the use of such dyes to fabrics such as cotton 
and rayon which, unlike wool and silk, are reasonably stable under alkaline 
conditions. 

Oxidation of the leuco form to the dye in the fiber can be achieved simply 
with oxygen of the air; but this is slow, and it is more common to regenerate 
the dye by passing the fabric, which has absorbed the leuco form of the dye, 
into a solution containing chromic acid or perboric acid. 



summary 

Polymers can be classified on the basis of their physical properties as elasto- 
mers (elastic substances), thermoplastic polymers (substances that flow when 
heated), or thermosetting polymers (rigid, insoluble, amorphous substances 
possessing cross links between the polymer chains). Polymers that are not 
cross-linked may be partly crystalline by virtue of van der Waals forces or 
hydrogen bonds causing ordering of the chains. These substances become 
glasslike at low temperatures (T g ) and begin to liquefy when heated above the 
melting temperature (T m ). 

Atactic polymers have a random orientation of groups along a chain 
whereas isotactic polymers have the groups oriented in the same direction. 
Syndiotactic polymers have a regular alternating orientation of the groups. 

Synthetic polymers can be prepared by condensation reactions between: 

(a) esters or anhydrides and alcohols 



R0 2 C-Z-C0 2 R + HO-Y-OH 



O + HO-Y— OH 




O 



O 



O 

II 



-C-O-Y-O-C-Z-C-O- 



O 
II 
-Y-O-C-Z- 



(b) carboxylic acids and amines 
H0 2 C-Z-C0 2 H + H 2 N-Y-NH 2 



o o o 

II II II 

-> -C-NH-Y-NH-C-Z-C-NH-Y- 



(c) phenols and formaldehyde 



exercises 763 



OH 





.CH,- 



Addition polymerization of a wide variety of vinyl compounds is brought 
about by radical initiators (peroxides or azo compounds). Chain growth is 

CH 2 =CHZ '—* -CH 2 -CH-CH 2 -CH-CH 2 -CH- 

I I I 

z z z 

terminated by radical disproportionation or by radical combination. 

Cationic and anionic polymerization is also possible with certain monomers. 

Naturally occurring polymers include cellulose and starch (polymers of 
glucose), rubber (a cis polymer of isoprene), nucleic acids (copolymers of 
substituted pentoses and phosphoric acid), wool, silk, and collagen and pro- 
teins in general (polypeptide copolymers). 

Four general procedures for dyeing fibrous polymers are (a) direct dyeing — 
useful with silk and wool, which are proteins and contain highly polar groups, 
and sometimes with cellulose fibers such as cotton, linen, and rayon; (b) dis- 
perse dyeing, direct solution of the dye in the fiber — useful with Dacron; 
(c) mordant dyes, formation of a complex of a metal salt, dye, and fiber — 
useful with cotton, wool, or silk; (d) vat dyeing, oxidation of a soluble form 
of a dye to give an insoluble form within the fiber — useful with cotton and 
rayon. 



exercises 

28-1 Write a reasonable mechanism for the thermal depolymerization of cyclo- 
pentadiene tetramer. How could you chemically alter the tetramer to make 
thermal breakdown more difficult ? Explain. 

28-2 Suppose a bottle of cyclopentadiene were held at a temperature at which 
polymerization is rapid, but depolymerization is insignificant. Would the 
polymerization result in conversion of all of the cyclopentadiene into essen- 
tially one gigantic molecule ? Why or why not ? How would you carry on the 
polymerization so as to favor formation of polymer molecules with high 
molecular weights ? 

28-3 Show how each of the following polymer structures might be obtained from 
suitable monomers by either addition or condensation. More than one step 
may be involved. 

a. -CH 2 -CH 2 -CH 2 -CH 2 -CH 2 -CH 2 -CH 2 - 

b. -N-CH 2 -CH 2 -N-CH 2 -CH 2 -N-CH 2 -CH 2 - 

I I I 

CH3 CH3 CH3 



chap 28 polymers 764 

c. — CH— CH— CH— CH— CH— CH— 

I II I I I 

CH 3 CH 3 CH 3 CH 3 CH 3 CH 3 

OOO 

II II II 

d. -0-CH 2 -CH 2 -C-0-CH 2 -CH 2 -C-0-CH 2 -CH 2 -C- 

e. -CH 2 -CH — CH 2 — CH — CH 2 — CH- 

II I 

0-C-CH 3 O-C-CH3 O-C-CH3 

II II II 

o o o 



/. 


-CH 2 -CH-CH 2 -CH-CH 2 — CH- 
1 1 1 




OH OH OH 


9- 


-ch 2 ^^ch 2 -ch 2 h^~^ch 2 -ch 2 h^~^ch 2 




1 




h. -0-CH 2 -CH-CH 2 -0-C-0-CH 2 -CH-CH 2 -0- 

O 

I 

c-o 

I 

o 

1 II 

-0-CH 2 -CH-CH 2 -0-C-0-CH 2 -CH-CH 2 -0- 

O 



28 -4 High-pressure polyethene (p. 1 02) differs from polyethene made with the aid of 
Ziegler catalysts (p. 103) in having a lower density and lower T,„ . It has been 
suggested that this is due to branches in the chains of the high-pressure mater- 
ial. Explain how such branches might arise in the polymerization process 
and how they would affect the density and T m . 

28-5 Radical-induced chlorination of polyethene in the presence of sulfur 
dioxide produces a polymer with many chlorine and a few sulfonyl chloride 
(— S0 2 C1) groups, substituted more or less randomly along the chains. 
Write suitable mechanisms for these substitution reactions. What kind of 
physical properties would you expect the chlorosulfonated polymer to have 
if substitution is carried to the point of having one substituent group to 
every 25 to 100 CH 2 groups? How might this polymer be cross linked? (A 
useful product of this general type is marketed under the name of Hypalon.) 

28-6 When polyethene (and other polymers) are irradiated with X rays, cross 
links are formed between the chains. What changes in physical properties 
would you expect to accompany such cross linking? Would the polyethene 
become more elastic? Explain. 

Suppose polyethene were cross-linked by irradiation above T m ; what 
would happen if it were then cooled ? 




exercises 765 

28-7 Answer the following questions in as much detail as you can, showing your 
reasoning : 

a. Why is atactic polymethyl methacrylate not an elastomer? 

b. How might one make a polyamide which is an elastomer? 

e. What kind of physical properties are to be expected for atactic poly- 
propene ? 

d. What would you expect to happen if a piece of high-molecular-weight 

polyacrylic acid -f-CH,, — CH -)- were placed in a solution of sodium 

hydroxide? ' 

COjH 

e. What kind of properties would you expect for high-molecular-weight 
poly-p-phenylene ? 



H poly-p-phenylene 



/ Are the properties, listed in Table 28-1, of polychloroprene as pro- 
duced by radical polymerization of chloroprene (2-chlorobutadiene) 
such as to make it likely that trans 1 ,4 addition occurs exclusively ? 

28-8 The material popularly known as Silly Putty is a polymer having an 
— O — Si(R) 2 — O — Si(R) 2 — O — backbone. It is elastic in that it bounces and 
snaps back when given a quick jerk but rapidly loses any shape it is given 
when allowed to stand. Which of the polymers listed in Table 28-1 is likely to 
be the best candidate to have anything like comparable properties ? Explain. 
What changes would you expect to take place in the properties of Silly Putty 
as a function of time if it were irradiated with X rays (see Exercise 28-6)? 

28-9 What kind of a polymer would you expect to be formed if />-cresol were 
used in place of phenol in the Bakelite process ? 

28-10 Polymerization of methyl methacrylate with benzoyl peroxide labeled with 
14 C in the aromatic ring gives a polymer from which only 57% of the 14 C 
can be removed by vigorous alkaline hydrolysis. Correlation of the 14 C 
content of the original polymer with its molecular weight shows that, on the 
average, there are 1.27 initiator fragments per polymer molecule. Write 
mechanism(s) for this polymerization that are in accord with the experi- 
mental data, and calculate the ratios of the different initiation and termina- 
tion reactions. 

28-11 The radical polymerization of styrene gives atactic polymer. Explain 
what this means in terms of the mode of addition of monomer units to the 
growing-chain radical. 

28-12 Polyvinyl alcohol prepared by hydrolysis of vinyl acetate (Table 28-1) does 
not consume measurable amounts of periodic acid or lead tetraacetate 
(Section 1 1 -3). However, the molecular weight of a typical sample of the 
polymer decreases from 25,000 to 5000. Explain what these results mean in 
terms of the structure of polyvinyl alcohol and of polyvinyl acetate. 



chap 28 polymers 766 

28-13 Ozonizations of natural rubber and gutta percha, which are both polyiso- 
prenes, give high yields of levulinic aldehyde (CH 3 COCH 2 CH 2 CHO) and 
no 2,5-hexanedione (CH 3 COCH 2 CH 2 COCH 3 ). What are the structures of 
these polymers ? 

28-14 Devise a synthesis of polyvinylamine, remembering that vinylamine itself 
is unstable. 

28-15 How will the side chains on the L-amino acids of silk fibroin be oriented with 
respect to the fiber sheets? 

28-16 Apparently the economically important chain reaction wool + moths -► 
holes + more moths has, as a key step, scission of the disulfide linkages of 
cystine in the polypeptide chains by the digestive enzymes of the moth larva. 
Devise a method of mothproofing wool which would involve chemically 
altering the disulfide linkages (review Chapter 19). 



wmwm 












33a 






k'* 3 ^'^ 



&5 VH 



[■y^h'*!''* *"' 



;>?■ *g' S-**¥»P- ,i t ?r.T5SSaS«S ^fcSs^S'v £'F> '-*? ■'.'*- iWJJ 



•.:,-.'-.'. ! .4-S-.:--V ft! 



£?■'.¥ W-*\$ -Sir; "&& ££#• ftSi S^^-'V.KSi 



Faisi?.-- va> 



'■■■■\yi 



>-mmMM$ 




#1 






'^Ji P»s S?:j 



nttm* 



m'i^^M^]t0i$$§^ i 



chap 29 some aspects of the chemistry of natural products 769 

The area of organic chemistry that deals primarily with the structures and 
chemistry of the compounds which are synthesized by living organisms is 
extremely large and highly variegated. Many types of natural products, includ- 
ing the carbohydrates, amino acids, proteins and peptides, and alkaloids 
(discussed in earlier chapters), have been investigated in such detail that whole 
volumes or series of volumes have been, or could be, devoted to their occur- 
rence, isolation, analysis, structure proof, chemical reactions, synthesis, 
biological function, and the biogenetic reactions by which they are produced. 

The chemistry of many classes of natural products is of general interest 
quite apart from their biochemical importance. Thus, the chemistry of the 
bicyclic terpenes contributed much to the interesting and unusual chemistry 
of such ring compounds long before satisfactory syntheses were available 
by the Diels-Alder reaction (Chapter 27). Similarly, studies of the chemistry 
of the steroids has added as much or more to our knowledge of conformations 
in cyclohexane rings as studies of cyclohexane derivatives themselves. Many 
other equally cogent examples could be cited. 

Our plan in this chapter is to first consider in some detail how the structures 
of natural products are established, both by classical procedures and by 
modern instrumental methods. We shall then consider in an illustrative way 
two rather closely related classes of natural products, terpenes and steroids. 
Finally, we discuss some of the aspects and uses of biogenetic schemes for the 
syntheses carried on by living systems. Throughout, we attempt to show how 
much of the material covered earlier in this book is pertinent to the study of 
natural products. 

29-1 civetone 

The active principle of civet, a substance collected from the scent gland of the 
African civet cat, is called civetone. This compound and one of similar nature 
called muscone, isolated from a scent gland of the Tibetan musk deer, are 
used in preparation of perfumes. Although civetone and muscone do not 
themselves have pleasant odors, they have the property of markedly enhancing 
and increasing the persistence of the flower essences. 

The structure of civetone was established in 1926 by the Swiss chemist 
Ruzicka. The starting material for his work was commercial civet imported 
from Abyssinia packed in buffalo horns — an inhomogeneous, yellow-brown 
unctuous substance, containing 10 to 15% water, intermixed with civet-cat 
hairs, and possessing a less-than-pleasant odor. Of several methods of 
isolation of the active principle, the most useful involved destruction of the 
glycerides present by hydrolysis with alcoholic potassium hydroxide, frac- 
tional distillation of the unsaponifiable neutral material under reduced 
pressure, and treatment of the distillate of bp 140° to 180° (3 mm) with 
semicarbazide hydrochloride in the presence of acetate ion (Section 11-4D). 
The crystalline product formed was the semicarbazone of civetone, and the 
yield indicated that the starting material contained 1 to 1 5 % of the active 
principle. Decomposition of the purified semicarbazone with boiling oxalic 
acid solution gave, after reduced-pressure distillation, crystalline civetone 
of mp 31°. The pure substance showed no optical activity. 



chap 29 some aspects of the chemistry of natural products 770 

Civetone is a ketone (an aldehyde would hardly have survived the alkaline 
isolation procedure) which was shown by its elemental analysis and molec- 
ular weight to be C 17 H 30 O. Saturated open-chain ketones have the general 
formula C„H 2 „0 and civetone has four hydrogens less, which means that it 
must have a triple bond, or two double bonds, or one double bond and one 
ring, or two rings. Civetone reacts with permanganate, gives a dibromide, and 
absorbs one mole of hydrogen in the presence of palladium. The presence of 
one double bond and one ring is therefore indicated, and a partial structure 
can be written as follows : 



/ 
o=c x 



C14H3 



;c 



Oxidation of civetone with cold potassium permanganate solution gave a 
dibasic keto acid which was at first thought to be C 16 H 28 5 (loss of a carbon) 
but later was shown to be C 17 H 30 O 5 . The formation of this acid confirms the 
presence of a ring and shows that each of the double-bonded carbons carries 
a hydrogen, because otherwise a dibasic acid with the same number of carbons 
could not be formed. 



/ f ] CH [O] / -co 2 H 

O-C \ C 14 H 2 J II ►0 = C^C 14 H 28 [ 



A key step in the determination of the structure of civetone was to find out 
how many carbon atoms separate the carbonyl group and the double bond. 
This was done by oxidation of civetone under conditions such as to lead to 
cleavage both at the double bond and at the carbonyl group. Different oxida- 
tion procedures gave somewhat different results but in all cases mixtures of 
dibasic acids were formed, the mildest conditions leading to formation of 
pimelic acid, H0 2 C(CH 2 ) 5 C0 2 H; suberic acid, H0 2 C(CH 2 ) 6 C0 2 H; and 
azelaic acid, H0 2 C(CH 2 ) 7 C0 2 H. The formation of azelaic acid indicates that 
there is at least one continuous chain of seven CH 2 groups forming a bridge 
between the carbonyl group and double bond. 

(CH 2 ) 7 
o=c . -i- ~^— 

- A {c 7 H 14 r CH H °^ u ,/ c ° 2H 

Since all the dicarboxylic acids isolated from the oxidation had continuous 
chains, Ruzicka inferred that the other seven carbons were also linked up 
in a continuous chain and that civetone is actually 9-cycIoheptadecenone. 

/ CH ^CH 
= C II 

Vh 2 ) 7 - CH 

9-cycloheptadecen- 1 -one 

This was an exciting conclusion at a time when the largest known mono- 
cyclic compounds were cyclooctane derivatives, and, along with the demon- 



sec 29.1 civetone 771 

stration in 1925 of the existence of cis- and trans-decalin (Section 29-4), 
provided decisive evidence against the Baeyer theory of angle strain in large 
carbocyclic rings (Section 3-4C). 

The postulated presence of a cycloheptadecene ring in civetone was sup- 
ported by oxidation of dihydrocivetone by chromic acid to heptadecanedioic 
acid. 



oo 3 



/ V H 2 ,Pd / CH 2 

° =C \ Jh^^ ° =C \ ^h ch 3 co 2 h,h 20 

(CH 2 ) 7 (CH 2 ) 7 " 2 

dihydrocivetone 



H0 2 C(CH 2 ) 15 C0 2 H 



heptadecanedioic 
acid 



Further evidence was obtained in confirmation of the structure of civetone, 
one particularly interesting series of transformations being as follows: 



0=C 



(CH 2 ) 7 / CH ^^CH 

_/ f H Zn(Hg) / CH 

(CH 2 ) 7 (CH 2 ) 7 



2H : 



civetane 



Pt 



H (CH 2 ) 7 _ 

\/ 
C 

HO / (CH 2 ) 7 ' 

dihydrocivetol 



CH 2 
I 
CH, 



KHSO 4> 180° 
(-H 2 0) 



/ 



HC 



,(CH 2 ) 7 __ C 

I 
\ ^CH 2 

CH-fCH 2 ) 6 

civetane 



The interest in these reactions is the demonstration that the symmetry of the 
civetone ring is such that civetane (cycloheptadecene) produced by the Clem- 
menson reduction (Section 11-4F) of civetone is identical with civetane 
obtained by reduction of civetone to dihydrocivetol and dehydration over 
an acidic catalyst. 

In some quarters, the structure of a natural product is not regarded as 
really confirmed until a synthesis is achieved by an unambiguous route, and 
research aimed at such syntheses has been a fascinating and popular part of 
organic chemistry for many years. Syntheses of naturally occurring sub- 
stances often yield very considerable benefits in the development of new 
synthetic reactions and furthermore may offer the possibility of preparing 
modified forms of the natural products which are of biochemical or medical 
interest. 

In the case of civetone, a synthesis was not achieved until long after the 
structure was established, but the finding of the cycloheptadecene ring in 
civetone led Ruzicka to develop a method for the synthesis of large-ring 
compounds which, although now largely superseded by other procedures, 
gave the complete series of cyclic ketones from C 9 to C 21 and a number of 
higher examples as well. 



(CH 2 )„ 



heat 
^ThbT 




(CH 2 )„ 



O + CO, + H,0 



chap 29 some aspects of the chemistry of natural products 772 

Ruzicka showed that pyrolyzing the appropriate dicarboxylic acid with 
thorium oxide gave 20 % of cyclooctanone and 1 to 5 % yields of the higher 
ketones. The yields are in the 1 % range from C 9 to C 12 , where conformational 
difficulties are to be expected during ring formation (Section 3-4C). 

Musk-type odors are found to be associated with the C 14 to C 17 cycloalka- 
nones, being particularly strong with cyclopentadecanone, which is available 
commercially under the name Exaltone. Interestingly, the odors of civetone 
and dihydrocivetone are the same. Further evidence for the presence of 
the 17-membered ring in civetone is supplied by the identity of synthetic 
cycloheptadecanone with dihydrocivetone. The synthesis of civetone was 
reported by Stoll and co-workers in 1948. 



29-2 spectroscopic methods in the determination 
of the structures of natural products 

The use of the types of spectroscopic methods described in Chapter 7 has 
greatly reduced the difficulty in determining the structures of the natural 
products of medium and low molecular weights. We have given many illustra- 
tions of the kind of information which is obtained from ultraviolet, infrared, 
and nmr spectroscopy in earlier chapters. Had these methods been available, 
their application to the problem of determining the structure of civetone 
would have been very helpful, but probably not decisive, for the reason that 
civetone is mostly saturated hydrocarbon and distinction between some of the 
possible isomers would be difficult if not impossible by spectroscopic methods. 
Considerable difficulty can be expected in structure determinations of cyclic 
compounds which have several saturated rings with no functional groups to 
permit degradation by selective oxidation. A typical case is that of que- 
brachamine, an indole alkaloid with a complex polycyclic ringsystem. 




'\ J C2H? 
H 

quebrachamine 

The ultraviolet spectrum of quebrachamine is typical of an indole and the nmr 
spectrum shows that the indole system is substituted at the 2 and 3 positions. 
Oxidation fails to open the saturated ring system, and the only very useful 
degradative reaction found so far is distillation with zinc dust at 400°, which 
yields a complex mixture of nitrogen compounds, including several pyridine 
derivatives. 

Fortunately, mass spectrometry is showing great promise in handling 
structural problems of just this variety and, rather than try to review the 
application of the other forms of spectroscopy to natural products, we shall 



sec 29.2 spectroscopic methods and structures of natural products 773 

concentrate here on mass spectrometry, which has hardly been mentioned 
since Chapter 7. 

One important use of mass spectrometry in the quebrachamine problem 
was in the identification of the components of the mixture of pyridine bases 
formed in the zinc-dust distillation. The mixture was separated by gas chro- 
matography (Section 7T) and the fractions identified by their mass spectra 
(Section 7-2B). 

The procedure for identification of compounds by mass spectrometry is 
first to determine mje for the intense peak of highest mass number. In most 
cases this peak corresponds to the positive ion M® (a radical cation) formed 
by removal of just one electron from the molecule M being bombarded, and 
the mje value of M® is the molecular weight. 

Removal of a nonbonding electron generally occurs more easily than 
removal of an electron from a chemical bond. In the case of compounds such 
as amines, ketones, and alcohols, all of which contain nonbonding electrons, 
the radical cation M® can be expected to have the following structures: 

:0 
M RNH 2 R-C-R R-O-R 

:6 ffi 
M® RNH 2 R-C-R R-O-R 

Incorrect molecular weights are obtained if the positive ion M® becomes 
fragmented before it reaches the collector, or if two fragments combine to 
give a fragment heavier than M®. The peak of M® is especially weak with 
alcohols and branched-chain hydrocarbons, which readily undergo frag- 
mentation by loss of water or side-chain groups. With such compounds the 
peak corresponding to M® may be 0.1 % or less of the total ion intensity. 

The pressure of the sample in the ion source of a mass spectrometer is 
usually about 1CT 5 mm, and, under these conditions, buildup of fragments to 
give significant peaks with mje greater than M® is rare. The only exception to 
this is the formation of (M + 1) peaks resulting from transfer of a hydrogen 
atom from M to M®. The relative intensities of such (M + 1) peaks are 
usually sensitive to the sample pressure and may be identified this way. 

With the molecular weight available from the M® peak with reasonable 
certainty, the next step is to study the cracking pattern to determine whether 
the mje values of the fragments give any clue to the structure. In the mixture 
of pyridine derivatives obtained by zinc-dust distillation of quebrachamine, 
the principal substance present showed M® at 107 and strong peaks at 106 and 
92 (Figure 29T). Loss of one mje unit has to be loss of hydrogen, while loss of 
fifteen corresponds to NH or CH 3 . Fragmentation of NH from a pyridine 
derivative seems to be a drastic change, but loss of a methyl radical, CH 3 , 
is reasonable, particularly if it leads to a stabilized positive fragment. The mje 
value of 107 corresponds to pyridine with one ethyl or two methyl groups. 
Using 3-ethylpyridine and 3,5-dimethylpyridine as specific examples, we 
could then have fragmentation reactions as follows : 



chap 29 some aspects of the chemistry of natural products 774 



100 



50 I 




100 



50 




100 



50 



j 

, I, , 


1 

III. .1,1. ...1 


I AM', 
1 I 


1, 


1 

', ,1 


1 

it.. 1 


ll. M, Jl 


1 1 


1 ;J 


- 
1, 


1 


1" ' " 

ll 1 


Li 


i, Jl 


M 

llpllfl 


i 

1 

15 
I 


M " 



40 



60 



100 rn/e 



m.w. 107 



a 



GH 2 CH 3 



107 



CrloCHa 



m.w. 107 



Figure 29-1 Mass spectra of 2-, 3-, and 4-ethyIpyridines. The vertical scale 
is relative peak intensity. The spectrum of the C 7 H 9 N base from the zinc- 
dust distillation of quebrachamine is the same as that of 3-ethylpyridine. 
(By permission from K. Biemann, Mass Spectrometry, Organic Chemical Applications, 
McGraw-Hill, New York, f%2.) 



sec 29.2 spectroscopic methods and structures of natural products 775 



cr CH ' 



(M) 
107 




"TO" 



CH, 



(M) 
107 



M ffi 
107 



-CH 3 



CHCH, 



CH, 



H,C 



106 



H 3 C. 



CH, 



N 5 
92 



For both compounds, the fragment of mass 106 is a benzylic-type cation 
which is expected to be stabilized by electron derealization, which will 
distribute the positive charge over the ring. The fragment of mass 92 would be 
a stabilized benzylic cation in the one case and a high-energy phenyl-type 
cation in the other. The high intensity of the 92 peak suggests that the com- 
pound of mass 107 is 3-ethylpyridine and the identity of their mass spectra and 
other properties confirms this assignment. 

Demonstration that the mixture of pyridines from zinc-dust distillation of 
quebrachamine contains 75% of 3-ethylpyridine provided strong support 
for the presence of a 3-ethylpiperidine grouping in the alkaloid. 




3-ethylpiperidine 



Further evidence on the structure of quebrachamine was obtained by 
comparison of its mass spectrum with that of a transformation product [1] of 
a related alkaloid of known structure, aspidospermine. 



C 2 H 5 




CH,0 



[1] 



chap 29 some aspects of the chemistry of natural products 776 

It will be seen that formula [1] is actually that of a methoxyquebrachamine 
and, if the methoxy group could be replaced by hydrogen, a synthesis of 
quebrachamine would be achieved from aspidospermine, thus establishing the 
structure of quebrachamine. Comparison of the mass spectra of [1] and 
quebrachamine is much easier and no less definitive. The spectra (Figure 29-2) 
at first glance look rather different, but careful examination shows that they 
are actually very similar from m/e — 138 downward. Furthermore, virtually 
all of the peaks that appear in quebrachamine from 143 up to that of M® (282) 
have counterparts in the spectrum of the methoxy compound just 30 units 
higher. This difference of 30 mass units is just the OCH 2 by which the molec- 
ular weight of the methoxyquebrachamine exceeds the molecular weight of 
quebrachamine itself. 

Assuming the indole part of quebrachimine does not break up very easily, 
we would expect the smallest abundant fragments from that part of the mole- 
cule to have masses of 143 or 144. 



CH, 




C,H S 



mol. wt. = 282 




CH 2 + C 9 H 17 N 



139 



CH 3 + C 9 H 16 N- 



138 



It is significant that the largest fragment from the saturated part of que- 
brachamine would then have m/e 138 or 139. From this we can see why 
substitution of the methoxyl group affects all peaks of 143 and over, but not 
those below this number. 



CH, 




CH,0 



mol. wt. = 312 



CH,0 



H 


=CH 2 


+ C 


»H, 7 N 


73 






139 


CH, 








H 


-CH 3 


+ c 


»H 16 N 



174 



138 



It would be a serious error to imagine that in mass spectra nothing is 
observed but simple fragmentation of organic molecules on electron impact. 
Actually, even though electron impact produces highly unstable molecular 



110 



282 



125 



nucbracharuinc 



138 



143 
1/ 



96 



115 



.11.1. ■■■..■■11. 



1 1.. ■■iiill. ..1 1 1 . 



fe: 



ii 



157 



..iiil.lii iJllil U 



199 210 
■I'l'- l.l-l 



Ik 



253 



267 

JL 



90 110 m/e 130 



150 



170 



190 



210 



i 
230 



i i i i i I t 

250 270 290 



110 



125 



AM 



96 



LMI 



115 



'".1 1. 



138 



M2* 



I"'- -lll-'ll- .■»!' 



173 

V 

■ I, llllil 



187 








CHC ) 



229 240 



283 



297 



i 
90 



li i 

110 m/e 130 



1 
150 



170 



190 



210 



230 



250 



270 



290 



310 



Figure 29-2 Mass spectra of quebrachamine and a transformation product, formula [1], of aspidospermine. (By permission from K. 
Biemann and J. Am. Chem. Soc.) The peaks at m/e 141 in quebrachamine and 1S6 in [1] correspond to parent ions which bear a double 
positive charge, that is, M 2 ®, 



chap 29 some aspects of the chemistry of natural products 778 



ions, there is a strong tendency for breakdown to occur by chemically rea- 
sonable processes (as with the ethylpyridines), and this may involve re- 
arrangement of atoms from one part of the molecule to another. An excellent 
example of such a rearrangement is provided by the M ffi ion of ethyl 
butanoate, which breaks down to give ethene and a radical cation of the enol 
form of ethyl acetate. 



H 3 c :6 B 

H 2 c x. JL 
c c 

(M e ) 
116 



CH 2 
CH 2 



28 






H 2 CT OC 2 H 5 



The cyclic course of this fragmentation is revealed by studies of the mass 
spectra of 2-, 3-, and 4-deuterated ethyl butanoate. The 2,2-dideuterio 
compound gives the enol ion, now with mass 90; the 3,3-dideuterio isomer 
gives the enol ion of mass 88 ; while the 4,4,4-trideuterated ester produces an 
ion of mass 89. 



D 2 C 

I 



II 

2 

(M®) 
119 



-OC,H 5 



D 2 C' O 



~C' OC 2 H 5 
H 2 



CD 2 

II 
CH 2 

30 



\ .® 

I 

H 2 C^ OC 2 H 5 
89 



29-3 



terp< 



•enes 



The odor of a freshly crushed mint leaf, like many plant odors, is due to the 
presence in the plant of volatile C 10 and C 15 compounds, which are called 
rerpenes. Isolation of these substances from the various parts of plants, even 
from the wood in some cases, by steam distillation or ether extraction gives 
what are known as essential oils. These are widely used in perfumery, food 
flavorings, and medicines, or as solvents. Among the typical essential oils are 
those obtained from cloves, roses, lavender, citronella, eucalyptus, pep- 
permint, camphor, sandalwood, cedar, and turpentine. Such substances are of 
interest to us here because, as was pointed out by Wallach in 1887, the 
components of the essential oils can be regarded as derived from isoprene. 



sec 29.3 terpenes 779 



CH 3 

I 

H 2 cr c 

H 

isoprene head tail 

(2-methyl-l,3- 
butadiene, C 5 H 8 ) 



^K^ 



Myrcene (C 10 H 16 ), a typical terpene, occurs in the oil of the West Indian 
bay tree, whose leaves are used in the preparation of bay rum. The carbon 
skeleton is clearly divisible into two isoprene units. 



II ' 
H 2 CT ^CH 

-\- II 

H 2 C. CH 2 
CH 

II 

/ c \ 

H 3 C CH 3 

myrcene 

The isoprene units in myrcene (and in almost all other terpenes as well) are 
connected in a head-to-tail manner. Because it is time consuming to show all 
the carbon and hydrogen atoms of such substances, we shall represent the 
structures in a convenient short-hand notation in which the carbon-carbon 
bonds are represented by lines, carbon atoms being understood at the junc- 
tions or the ends of lines. By this notation, myrcene can be represented by 
formulas like the following. 




In the past the term terpene was reserved for C 10 hydrocarbons such as 
myrcene. Current practice is to designate as terpenes (or isoprenoids or 
terpenoids) all compounds that are multiples of the C 5 isoprene skeleton 
including hydrocarbons, alcohols, aldehydes, and so on. The C 10 compounds 
are monoterpenes, C 15 are sesquiterpenes, C 20 are diterpenes, C 30 are 
triterpenes, and so on. 

The empirical isoprene rule resulted from Ruzicka's observation that the 
majority of the terpene families could be considered as arising from head-to- 
tail combinations of isoprene units. Thus, the monoterpene (C 10 ) family 
represents two such units, the sesquiterpenes (C 15 ) three, the diterpenes 
(C 20 ) four, and so on. Some examples to illustrate this hypothesis are shown 
with the head-to-tail isoprene junctions indicated by dashed lines. In the 
cases of cyclic terpenes the thinner bonds indicate where subsequent cycli- 
zation has taken place. 



chap 29 some aspects of the chemistry of natural products 780 



Monoterpenes (C 10 ) 




CH,OH 




CH,OH 




geraniol 
(ginger grass) 



CHO 



citronellal 
(oil of citronella) 



nerol 
(rose) 



CH 2 OH 



citronellol 
(rose) 



limonene 
(lemon, orange) 






menthol 
(peppermint) 



ascaridole 
(chenopodium oil) 



camphor 
(camphor tree) 



The terpene alcohols shown above have floral odors and are important 
perfume ingredients. The aldehydes have much stronger, citruslike odors 
and occur in many essential oils such as oil of citronella and oil of lemon. 
Ascaridole is interesting in being a naturally occurring peroxide. Camphor 
is a volatile solid substance which for centuries was believed to have medicinal 
properties. It is now used chiefly as a plasticizer for cellulose nitrate (Section 
15-7) and is synthesized commercially from a-pinene (see Exercise 29-12). 




camphor 

Camphor has a very large molal freezing point depression constant which 
makes it useful for measuring molecular weights. 

Camphor also produces large changes in the surface tension of water and a 
chip of wood with a piece of camphor embedded in one end will be propelled 
by the difference in surface pressure produced between the ends as the 
camphor spreads on the surface of the water (camphor boat). This effect is 
used in a defense mechanism of a tiny water beetle (Stenus bipunctatus), which 
when threatened by birds, streaks across the surface of the water by expelling 
a mixture of surface-active terpenes from the tip of its abdomen. 

Sesquiterpenes (C 15 ) 



CH 2 OH 


. 


i 


rH 'v 


<>Np\ 


1^1^*1 


w 


Y^V 


1 o — 


farnesol 
(lily of the valley) 


/?-selinene 
(oil of celery) 


santonin 
(Artemisia) 



Farnesol, which occurs in lily of the valley and other plants, is a sex 
attractant for male insects. Santonin, extracted from the plant Artemisia, has 



sec 29.3 terpenes 781 

been used for centuries in India as a medicinal because of its anthelmintic 
property (ability to destroy intestinal worms). 

Diterpenes (C 20 ) 




CH,OH 




,CH 2 OH 



phytol 
(chlorophyll) 



vitamin A 




abietic acid 
(pine rosin) 



Whereas head-to-tail isoprene junctions can be readily picked out in phytol 
and vitamin A the situation with abietic acid and mariy of the higher cyclic 
isoprenoids is somewhat different because alkyl migrations have taken 
place during their formation. 

Phytol occurs as an ester of the propanoic side chain of chlorophyll 
(Figure 15-1) and as a side chain in vitamin K t (Section 23-7). Abietic acid 
is a major constituent of rosin, which is obtained as a nonvolatile residue in 
the manufacture of turpentine by steam distillation of pine oleoresin or 
shredded pine stumps. Abietic acid is the cheapest organic acid by the pound 
and is used extensively in varnishes and as its sodium salt in laundry soaps). 



Triterpenes (C 30 ) 




squalene 
(shark liver oil) 



lanosterol 
(wool fat) 



/?-amyrin 
(manila elemi) 



The cyclic structures of /?-amyrin and lanosterol and the folded conformation 
shown for squalene resembles the ring system of the steroids, an extremely 
important family of natural products to be discussed in the next section. The 



chap 29 some aspects of the chemistry of natural products 782 

resemblance has a fundamental basis, as we shall see when we examine the 
biosynthesis of terpenes and steroids in Section 29-5. 

Tetraterpenes (C 40 ) 




lycopene 
(plant pigment, tomatoes, etc.) 




'-carotene 
(plant pigment, carrots, etc.) 



The long conjugated chains in these compounds are responsible for their 
color (Section 7-5). 



29-4 steroids 

In the discussion of the isoprenoid compounds it was our intention to show 
how the occurrence, structures, and properties of a large and important class 
of natural products can be correlated. In keeping the discussion within 
reasonable bounds it was not possible to show how the various structures 
were established, or give any one compound particular attention. With 
steroids, we shall take the opposite approach of considering one member of 
the class, cholesterol, in some detail and then show only the structures of 
some other representative steroids. 

The term steroid is generally applied to compounds containing a hydro- 
genated cyclopentanophenanthrene carbon sketeton. Many of these com- 




cyclopentanophenanthrene 

pounds are alcohols, and sometimes the name sterol is used for the whole class. 
However, sterol is better reserved for the substances that are actually alcohols. 



A. CHOLESTEROL 



Cholesterol is an unsaturated alcohol of formula C 27 H 45 OH which has long 
been known to be the principal constituent of human gallstones. Cholesterol, 
either free or in the form of esters, is actually widely distributed in the body, 



sec 29.4 steroids 783 

particularly in nerve and brain tissue, of which it makes up about one sixth 
of the dry weight. The function of cholesterol in the body is not understood. 
Experiments with labeled cholesterol indicate that cholesterol in nerve and 
brain tissue is not rapidly equilibrated with cholesterol administered in the 
diet. Two things are clear: Cholesterol is synthesized in the body and its 
metabolism is regulated by a highly specific set of enzymes. The high specific- 
ity of these enzymes may be judged from the fact that the very closely 
related plant sterols, such as sitosterol, are not metabolized by the higher 
animals, even though they have the same stereochemical configuration of all 
groups in the ring and differ in structure only near the end of the side chain. 



HO 




HO' 




cholesterol 



sitosterol 



The cholesterol level in the blood generally rises with a person's age and body 
weight and is usually higher in populations whose diets are rich in animal fats. 
Atherosclerosis (hardening of the arteries) in man is often associated with 
high cholesterol levels in the blood and, indeed, it is possible to produce the 
disease in certain animals by feeding them diets high in cholesterol. 

Although cholesterol was recognized as an individual chemical substance 
in 1812, all aspects of its structure and stereochemical configuration were not 
settled until about 1955. The structural problem was a very difficult one, 
because most of cholesterol is saturated and not easily degraded. Fortunately, 
cholesterol is readily available, so that it was possible Jo use rather elaborate 
degradative sequences which would have been quite out of the question with 
some of the more difficultly obtainable natural products. 

The first step in the elucidation of the structure of cholesterol was the deter- 
mination of the molecular formula, first incorrectly as C 26 H 44 in 1859 and 
then correctly as C 27 H 46 in 1888. The precision required to distinguish be- 
tween these two formulas is quite high, since C 26 H 44 has 83.82% C and 
11.90% H, whereas C 27 H 46 has 83.87 %C and 11.99% H. Cholesterol was 
shown in 1859 to be an alcohol by formation of ester derivatives and in 1868 
to possess a double bond by formation of a dibromide. By 1 903 the alcohol 
function was indicated to be secondary by oxidation to a ketone rather than an 
aldehyde. The presence of the hydroxyl group and double bond when com- 
bined with the molecular formula showed the presence of four carbocyclic 
rings. Further progress was only possible by oxidative degradation. 

There is but one point of unsaturation in the cholesterol molecule and 
oxidative reactions are not expected to proceed very well. However, chromic 
acid has the property of attacking tertiary hydrogens, probably by removal of 
H : e and formation of a carbonium ion. Under these conditions, El elimination 
is expected, and this is likely to give the most highly substituted alkene which 
would then be cleaved by the chromic acid. With the side chain of cholesterol, 



chap 29 some aspects of the chemistry of natural products 784 

two points of cleavage might be expected. Both processes occur, although the 




CH 3 P 



* X) + 



cholesterol 
(partial structure) 



J~^y C0 2 H + (CH 3 ) 2 CO 



yields are poor. The observation that methyl isohexyl ketone was formed by 
cleavage of the side chain was important in that it gave the first identifiable 
fragment of known structure. The discovery of a second point of cleavage was 
even more significant, because it permitted correlation of cholesterol with 
another series of compounds, known as the bile acids. The principal bile 
acids are cholic acid [2] and desoxycholic acid [3]. 



CO,H 



HO' 




HO- 




CO,H 



desoxycholic acid 
[3] 



The presence of a number of substituents on the cycloalkane rings of 
molecules such as cholic acid and cholesterol gives rise to various stereo- 
chemical possibilities. The three hydroxyl groups in cholic acid are said to 
occupy a positions ; that is, they are directed away from the viewer when the 
skeleton of the molecule is oriented as shown above. On the other hand the 
hydroxyl group in cholesterol and the angular methyl groups in both choles- 
terol and cholic acid occupy fj positions. (See Section 15-3 for a similar use 
of a and /? in the carbohydrate series). 

Both cholic acid and desoxycholic acid occur in bile as sodium salts of 
N-acyl derivatives of glycine (RCONHCH 2 C0 2 e Na e ) and taurine, j8-amino- 
ethanesulfonic acid (RCONHCH 2 CH 2 S0 3 e Na®). The function of the salts in 
bile is to aid in the solubilization and assimilation of fats and hydrocarbons, 
such as carotene. The bile acids are obtained by alkaline hydrolysis of the 
peptide bonds. 

It will be seen that each of the six-membered rings in cholic acid carries a 



sec 29.4 steroids 78S 

hydroxyl group, and this is most important, because it provides an entry into 
the rings by various kinds of oxidative processes. Furthermore, the side chain 
is seen to be the same length as in one of the chromic acid- oxidation products 
of cholesterol. The general similarity of the structures of the bile acids and 
cholesterol, including the stereochemical relations of the rings, strongly 
suggests that one is the precursor of the other in the body or that the two have 
a common precursor. Tracer experiments have shown that cholic acid can, in 
fact, be manufactured from dietary cholesterol. 




HO ^ 



stereochemical configuration and 
numbering system of cholesterol 

[4] 



Proof that cholesterol and the bile acids have the same general ring system 
was achieved by reduction of cholesterol to two different hydrocarbons, 
cholestane and coprostane, which differ only in the stereochemistry of the 
junction between rings A and B. 





cholestane 



coprostane 



The differing stereochemistry at the ring junction is shown in the following 
conformational formulas in which the A and B rings are shown in the chair 
conformations. (The symbol, — ' — , indicates the ring junction with ring C in 
each case.) 



CH, 



CH 





cholestane 
[5] 



coprostane 
[6] 



The A and B rings of these compounds can be regarded as derivatives of the 
alicyclic compound decahydronaphthalene (decalin), C 10 H 18 , which exists in 
stable cis and trans forms. 



chap 29 some aspects of the chemistry of natural products 786 
H 





/ra«,s-decalin cw-decalin 

Oxidation of coprostane, but not cholestane, gave an acid which turned out 
to be identical with cholanic acid obtained by dehydration of cholic acid at 
300° followed by hydrogenation. 



CO,H 




cholanic acid 

Determination of the sizes of the rings in cholesterol and the bile acids was 
achieved in part by use of the so-called Blanc rule, which states that a six- 
carbon dicarboxylic acid on heating will give a ketone, whereas a five-carbon 
dicarboxylic acid will give an anhydride (Section 13-1 IB). Reduction of 
cholesterol to cholestanol followed by oxidation yielded a dibasic acid, which 
on heating formed a ketone. This indicated the ring on which the hydroxyl 
was located to be a six-membered ring. 



jS? 




H 

cholestanol 



H 3 C I^S 




-H 2 0, -C0 2 



-* o 




Application of the Blanc rule to dicarboxylic acids obtained by opening the 
B ring indicated this ring to be six-membered but gave the wrong answer on 
ring C, an anhydride being formed in place of a ketone. The correct ring size 
was obtained for ring D by removing the side chain from cholanic acid, 
opening the ring by oxidation, and showing that an anhydride was formed on 
pyrolysis. 

Location of the methyl groups was achieved by extended degradations — the 
methyl at C-10 being located by degradation of desoxycholic acid [7] to 
a-methyl-a-carboxyglutaric acid [8]. 



sec 29.4 steroids 787 




CO,H 



[7] 



H 3 C 
H0 2 C C0 2 H 

[8] 



With the wrong size for ring C, it was inevitable that at some stage incorrect 
structures would be proposed for cholesterol and the bile acids. Tentative 
structures [9] and [10] proposed in 1928 for desoxycholic acid and cholesterol, 
respectively, show a resemblance to the structures now known to be correct, 
but have a five-membered ring fused to ring A. 



HO 




HO 




[10] 



Shortly thereafter, X-ray diffraction measurements indicated sterols to be 
extended rather than compact molecules. This evidence combined with the 
formation of chrysene and methylcyclopentanophenanthrene [11] from 
selenium dehydrogenation of cholesterol led to postulation of the correct ring 
structure in 1932. 

Figure 29-3 Proton nmr spectrum of cholesterol at 100 MHz as a 10% 
solution in deuteriochloroform with reference to TMS at 0.00. At 60 MHz 
the chemical shifts are smaller and many of the features of the spectrum 
between 0.7 and 2.4 ppm are run together and less distinct. (Spectrum 
kindly furnished by Varian Associates.) 



!'il)0 



~T~ 

400 



.'in i 

CH, 
groups i 



il llz 



H 



II 



-( ) 



I 



_1 

4 



-OH 

I 



\ -* **■' I . : 



TMS | 



■V— | 



ppm 



chap 29 some aspects of the chemistry of natural products 788 

CH, 



cholesterol 



Se 



300° 





chrysene 



[11] 



The absolute configuration of the sterols and bile acids was established in 
1955. Cholesterol has eight asymmetric centers and therefore there are 256 
possible stereoisomers, but only cholesterol itself occurs naturally. 

The proton nmr spectrum of cholesterol at 100 MHz is shown in Figure 
29-3. Such spectra are obviously of considerable value in the determination of 
the structures of even quite complex natural products. With cholesterol, many 
of the protons at or near the functional groups stand out quite clearly. 



B. REPRESENTATIVE STEROIDS 

The structures and physiological functions of a number of important steroids 
are shown in Table 29 T . Total syntheses have been achieved for the important 
sterols, sex hormones, and adrenal cortical hormones. The need for large 
quantities of cortisone and related substances for therapeutic use in treatment 
of arthritis and similar metabolic diseases has led to intensive research on 
synthetic approaches for methods of producing steriods with oxygen functions 
at C-ll, which is not a particularly common point of substitution in steroids. 
The most efficient way of doing this is by microbiological oxidation, and 
cortisone can be manufactured on a relatively large scale from the saponin 
diosgenin, which is isolated from tubers of a Mexican yam of the genus 
Dioscorea. Diosgenin is converted to progesterone, then by a high-yield (80 to 
90 %) oxidation with the mold Rhizopus nigricans to 1 1-hydroxyprogesterone, 
and finally to cortisone. 



CO CH, 




COCH,OH 



cortisone 



sec 29.S biogenesis of terpenes and steroids 789 

Two synthetic steroids whose use has implications in the areas of phys- 
iology, economics, sociology, and religion are the compounds norethindrone 
[12] and mestranol [13]. These and a number of other compounds which are 
structurally related to the sex hormones (Table 29-1) can be used to control 
ovulation. Mixtures of such compounds are marketed as oral contraceptives 
under various names, Enovid, Ortho-Novum, and so on. 

H 3 C t"?c»CH 




CH 3 



[13] 



Vitamin D is of special interest as a photochemical transformation product 
of ergosterol. 




hv 



2820 A 




ergosterol 



HO 



vitamin D 2 

(X-ray diffraction studies indicate the 

transoid configuration of the 6,7 bond 

in the crystal) 



29 -S biogenesis of terpenes and steroids 

Inspection of the structures for the pentacyclic triterpene, jS-amyrin, and 
cholesterol shows such a striking resemblance between the carbon skeletons 
that it is not hard to imagine that the way in which these substances are 
synthesized, their biogenesis, may be closely related. 





amynn 



cholesterol 



chap 29 some aspects of the chemistry of natural products 790 










a. 







T3 


>- 

a. 


C 


*« 





u 




Ml 











u 





3 


tr 


u 


^ 


u X 





0. 



s 
§ 

"0 

S 





t> 




0. 











T3 


a 


fl 




« 


« 




u 


u 

s 


M) 




u 





3 




u 


>-. 


ox 





a. 



e 

-0 

a 





"o 


s 


S3 


> 




o 


C3 


a 



o 




J3 








O 






o 


1) 


>* 


e 


o 


o 




E 

u 
O 

J5 


C8 

3 


X 


so 




rt 






<y 


3 




Ml 



e 
o 

a 

o 



fl 

IS >> 

a x> 
oo 



c 

a 

3 
J3 



a a 

p 3 

S3 




Q 

'a 




3 

o 



> 



' fl 
u 

Q 



o 






™ 45 
. o 

g-a 

E r~ 

Mu 

8 .a 

o c 

S o 



J&-2 

a is 



o 

o 

a 

c 



p 



a 

3 

J3 




sec 29.5 biogenesis of terpenes and steroids 791 



(A 


SO 
CO 

•o 2 






coside at the 

gitalis plants, 

poison, used 

to regulate 




C/3 

o 

ft 

, (» 

T3 3 


!* 


O to 








c3 3 





a o 








^ d 


a. 


8 « 








> is 




s S 

!»> 


S » 
o a 






£■'-3 o S 
60 a .2 g 
K — "S "° 

ftK OB 

g g « 
o -o a S 




J3 cs 

■3& 






d 
.g 


aj w 
d cs 

8 M 


II 






^_, 


p q 






o >* o co 




ft M 


a 






J5 


2 8 


ft 


CS 


p 


~\ s 


t3 


X 


X 

Vx 




0^/\ 


o^> 


^~\ 




w-S? T > 










/ x 




\ 1 / 




o4 — ( 


S 


y-7 


' \° 




/ \ 




*•» / \ 


fO / 




.2 


/ \ 4) 


a 


K \ /~ 


o 


K \ 




'3 


\ /"A -S 


ca 




\ ^ 








\ / \ 'to 


1 

a 




/ CO 

/ o 
■v-K T3 




H7 — r* 
x \ / 


'So 


^ V // to 

\ // 4) 


x,/ \ o 

,"/ \ o 
\ / 


O ) 


T3 

s 


m / 

x \ 


\ <3 
/ S 




^ 















\ 


I 




b 




o 




Z^ 


t 




X 




X 






3 

5 












X 












o. 


CO 


8 g g 










^3 en -m 

S % 4 




"H 3 ;>, 






o jq « 




V 

j- 

0. 


^ a. 

* B 

§,2 
6.2 
3 s 


5? "0 P 
» £ ,9 

a * B 

8 o a " 

i> CS rt 
.. > 00 rt 

« o n J3 






o *3 > 
cs *g *t3 _>» 
S a o > 

rt C 5 c3 

o c»ja 

t; „ .gp 

gag* 




5 tj ft 

O « !fl 

' s |:I 

3 *2 d 
a mr 

° <o a 

•3 ' 3 & 
d o x 

o o<a 

ft ^ d 




sfl O a 

s 






g 3 VO £S 




g ao at 


ft 






J3 




C3 












X 














u, 










X 














o 












4) 

c 
o 


29 




o- 


r-Q 

v/\ X d 

t 7-0 a 


U / \ 


a 



x \ / 


u 

^\ to 

\ O 






o 5C 

■B t 

b _ 


\ / »> 


■* / \ s 

fn / \ +z 


U / 




s 

B 




/ co 
/ CD 


*\ 




o s < i v so 

o \ / \ = 
\ / \ ^3 


c 

-a 
s 


x( 


\ 


o 


u "7 — 1 






U ) V"X 


B 












3 




o 










-w 














U 














3 








o 




o o 


■4-t 












X X 


to 















i a 

« cr 

8 «•> 

<= S 1 

u o 
>-. H 

Si 'O 

St 

^ w 
o *-* 

1.S 

'"■ >, 

^^ 

M O 

eS 

■Ss 

o rS 
a> 

pa t« 



o 

■s 


| 






o 




<& 


1 




T3 






cd 


a 




75 


a 




<u 


& 










<u 


o 




o 


>j 




CJ 






fl 


& 






o 






l-H 




cd 


P. 
















nj 4D 










a 


a»i 




o 


ot 




a 


e 




P3 


Q> 




W 


J3 




C 












^H 


a! 




Q 


J3 




O 


ts 




■o 


c 




c 






M 


&n 




u 


T) 










m 


1> 






T3 




,.rl 








O 




%^ 






o 


>> 


CS 


>1 


ou 










6 




X 




M 




O 


c 


T3 


o 


c 


is 


I 


M 
O 


5 






X! 








HQ 


a 


a * 


* 



chap 29 some aspects of the chemistry of natural products 792 

This idea is heightened by the structure of lanosterol, the tetracylic triter- 
pene alcohol which occurs alongwith cholesterol in wool fat and has properties 
so typical of the sterols that it is better known as a sterol than a triterpene. 




CH 3 

lanosterol 

Positive evidence that terpene and steroid biogenesis are related has been 
provided by experiments with carbon-14 labels which show that acetic acid is 
their common biological precursor. The general biosynthetic pathway in- 
volving polymerization of C 5 units is shown in Figure 29-4. 

Acetate is converted to 3,5-dihydroxy-3-methylpentanoic acid (mevalonic 
acid) by coenzyme A. Further enzymic action leads to the diphosphate, 
isopentenyl pyrophospate (IPP), the phosphoryl donor being ATP. IPP is the 
biological equivalent of isoprene which is not itself found in nature. 



CH, 



o 



o 



CH, = C 



-CH 2 — CH,— O— P— O-P— O e = 

I I 

o o 

e e 

isopentenyl pyrophosphate (IPP) 



The formation of C 10 compounds occurs by the condensation of one 
molecule of IPP with one molecule of its structural isomer, 3,3-dimethyl- 
allyl pyrophosphate (3,3-D). This reaction involves a displacement of pyro- 
phosphate from 3,3-D by the methylene carbon of IPP and the loss of a proton 
from the C-2 position. The compound thus formed is a pyrophosphate 
derivative of geraniol and is the precursor of the whole set of monoterpenes. 




"o-p 2 o 6 3e + hp 2 o, 3£ 



geranyl pyrophosphate 

Further condensation with another molecule of IPP produces a C 15 
pyrophosphate, which turns out to be the precursor not only of the sesquiter- 
penes (C 15 ) but also squalene and lanosterol (C 30 ) and the steroids, as shown 
in Figure 29-4. 

The conversion of squalene to lanosterol is particularly interesting because, 



sec 29.5 biogenesis of terpenes and steroids 793 



3CH,CO,H 



CH 3 OH 

coenzyme A CH 2 CH 2 



C0 2 H CH 2 — OH 

mevalonic acid 



enzymic 
~„ „ 3G equilibration 
0-P 2 0e . 



IPP 



monoterpenes <- 
(C, ) 



Ci o-pyrophosphate 
(geranyl pyrophosphate) 



IPP 



sesquiterpenes <- 
(C 15 ) 



diterpenes 
(C 20 ) 



C ^-pyrophosphate 



IPP 



C 20 -pyrophosphate 



tetraterpenes 
(C 40 ) 



3,3-D 



squalene 
(C 30 ) 



lanosterol 
(C 30 ) 



cholesterol 
(C 27 ) 



Figure 29-4 General biogenetic route to terpenes and steroids starting from 
acetic acid. 



although squalene is divisible into isoprene units, lanosterol is not, a methyl 
being required at C-8 and not C-l 3. 



HO 




lanosterol 



chap 29 some aspects of the chemistry of natural products 794 

Some kind of rearrangement is therefore required to get from squalene to 
lanosterol. The nature of this rearrangement becomes clearer if we write the 
squalene formula in the shape of lanosterol. 




squalene 

When written in this form, we see that squalene is beautifully constructed 
for cyclization to the ring system of lanosterol and this conversion has 
recently been shown to occur in both plants and animals via [14], the 2,3- 
epoxide of squalene. Acid-catalyzed opening of the epoxide ring and ring 
closure in the remainder of the molecule are essentially synchronous and 




HO 




HO 




[16] 



produce a carbonium ion [15] which eliminates a proton to give [16], an 
isomer of lanosterol. Now if a sequence of carbonium ion-type rearrange- 
ments occurs we can see how [16] could be readily converted to lanosterol. 



summary 795 



HO 





HO 



ofH 9 ,CH 3 e ,CH 3 s 



HO 




ianosterol 



The evidence is strong that the biogenesis of Ianosterol actually proceeds 
by a route of this type. With squalene made from either methyl- or carboxyl- 
labeled acetate, all the carbons of Ianosterol and cholesterol are labeled as 
predicted. Furthermore, ingenious double-labeling experiments have shown 
that the methyl at £-13 of Ianosterol is the one that was originally located at 
C-14, whereas the one at C-14 is one that came from C-8. 

The conversion of Ianosterol to cholesterol involves removal of the three 
methyl groups at the 4, 4, and 14 positions, shift of the double bond at the 
B/C junction to between C-5 and C-6, and reduction of the C-24 to C-25 
double bond. The methyl groups are indicated by tracer experiments to be 
eliminated by oxidation to carbon dioxide. 



summary 

Elucidation of the structure of a natural product usually involves degradation 
to smaller fragments that can be identified with known compounds. Mass 
spectrometry can be used to advantage both for producing fragments and for 
identifying them, and for determining molecular weights. Fragmentation 
patterns in the mass spectrometer are best interpreted by postulating that 
stabilized cations will tend to be produced either directly or by rearrangement. 
Terpenes (isoprenoid compounds) contain carbon skeletons made by head- 
to-tail unions of C 5 isoprene units, for example. 



(isoprene) 



(myrcene) 



head 



tail 



chap 29 some aspects of the chemistry of natural products 796 



Many terpenes contain oxygen, for example, camphor and vitamin A. 
The junctions of the isoprene units in these compounds are shown here by 
dashed lines and ring-closure positions with arrows. 




(camphor) 




CH,OH 



(vitamin A) 



Steroids, which are derivatives of cyclopentanophenanthrene, include 
cholesterol (a sterol) and cholic acid (a bile acid). They have the same carbon 



HO 




HO- 




CO,H 



(cholesterol) 



H 

(cholic acid) 



skeleton in the rings but differ in configuration as shown. 

Studies using carbon- 14 labels have shown that the biogenesis of cholesterol 
follows the sequence: acetic acid, mevalonic acid, several isoprenoid pyro- 
phosphates, squalene, lanosterol, and cholesterol. 



exercises 

29-1 Muscone, the active principal of Tibetan musk, is an optically active ketone 
of formula Ci 6 H 3 oO. On oxidation it gives a mixture of dicarboxylic acids. 
At least two acids of formula C16H30O4 are formed, along with some 
dodecanedicarboxylic acid and suberic acid. 

Clemmenson reduction of muscone gives an optically inactive hydro- 
carbon shown by synthesis to be methylcyclopentadecane. Muscone is not 
racemized by strong acids or strong bases, although it does form a benzyli- 
dene (=CHC 6 H 5 ) derivative with benzaldehyde and sodium methoxide. 

What structure(s) for muscone are consistent with the above experimental 
evidence ? Give your reasoning. What additional evidence would be helpful ? 

29-2 Explain how use of ultraviolet, infrared, or nmr spectroscopy could be used 
to distinguish between the following possible structures for civetone. 

a. 9-cycloheptadecenone and 2-cycloheptadecenone 

b. m-9-cycloheptadecenone and fraw.s-9-cycloheptadecenone 

c. 8-methyl-8-cyclohexadecenone and 9-cycloheptadecenone 

d. 8-cycloheptadecenone and 9-cycloheptadecenone 



29-3 How could you use deuterium labeling to show that the fragmentation of 
3-ethylpyridine which occurs in the mass spectrometer results in loss of the 
CH 3 group and not an NH fragment ? Be as specific as possible. 



exercises 797 

29-4 The relative intensities of (M — 15) peaks for the 2- and 4-ethylpyridines are 
much less than for 3-ethylpyridine (see Figure 29-1). Suggest a reason for this. 
Can your explanation also account for the fact that intensity ratios (M — 15)/ 
(M — 1) for the different isomers fall in the ratio 3 > 4 > 2? Explain. 

29-5 a. Identify the fragments in the mass spectrum of quebrachamine with m/e 

values of 267, 253, 157, and 125. Show your reasoning. 

b. The very strong peak at 1 10 in the mass spectrum of quebrachamine has 

no counterpart at 172. How might a fragment of 110 mass units be 

reasonably formed by breakdown of the primary dissociation products ? 

29-6 The mass spectrum of 1-phenylpropane has a prominent peak at mass 92. 
With 3,3,3-trideuterio-l-phenylpropane, the peak shifts to 93. Write a 
likely mechanism for breakdown of 1-phenylpropane to give a fragment of 
mass 92. 

29-7 The mass spectra of alcohols usually show peaks of (M — 18), which 
correspond to loss of water. What kind of mechanisms can explain the 
formation of (M —18) peaks, and no (M — 19) peaks, from 1,1-dideuterio- 
ethanol and l,l,l,3,3-pentadeuterio-2-butanol? 

29-8 a. Write out all of the possible carbon skeletons for acyclic terpene and 
sesquiterpene hydrocarbons that follow the isoprene rule. Do not con- 
sider double-bond position isomers. 
b. Do the same for monocyclic terpene hydrocarbons with a six-membered 
ring. 

29-9 The terpene known as alloocimene (Ci Hi 6 ) shows A max at 2880 A and gives 
among other products 1 mole of acetone and 1 mole of acetaldehyde on 
ozonization. What is a likely structure for alloocimene ? Show your reasoning. 

29-10 Write structures for each of the optical and cis-trans isomers that are 
possible of the following isoprenoid compounds : 

a. myrcene e. /3-selinene 

b. farnesol /. oc-pinene 

c. limonene g. camphor 

d. phytol 

29-11 Nerol and geraniol cyclize under the influence of acid to yield a-terpineol. 
How could the relative ease of cylization of these alcohols, coupled with 
other reactions, be used to establish the configurations at the double bond of 
geraniol, nerol, and the corresponding aldehydes, geranial and neral ? Write 
a mechanism for the cyclizations. 



-OH 

a-terpineol 
29-12 Camphor can be made on an industrial scale from a-pinene (turpentine) by 



chap 29 some aspects of the chemistry of natural products 798 

the following reactions, some of which involve carbonium ion rearrange- 
ments of a type particularly prevalent in the bicyclic terpenes and the 
scourge of the earlier workers in the field trying to determine terpene 
structures. 



Ti0 2 -H 2 



CH 3 CQ 2 H 




180° k\ J H® 

a-pinene camphene isobornyl acetate 



H 2 



isoborneol 



[O] 




camphor 



Write mechanisms for the rearrangement reactions noting that hydrated 
titanium oxide is an acidic catalyst. 

29-13 How many optical isomers of cholic acid are possible? 

29-14 Assuming cholesterol has the stereochemical configuration shown below 
draw a similar configurational structure for cholic acid (including the 
hydroxyl groups). 



HO 




29-15 When the sodium salt of 12-ketocholanic acid is heated to 330°, 1 mole of 
water and 1 mole of carbon dioxide are evolved and a hydrocarbon 
" dehydronorcholene " is formed. Selenium dehydrogenation of this sub- 
stance gives methylcholanthrene. 




CH, 



methylcholanthrene 




12-ketocholanic acid 



What is a likely structure for "dehydronorcholene" and how does the 
formation of methylcholanthrene help establish the location of the sterol 
side chain on ring D ? 



exercises 799 

29-16 Using the stereochemical information shown in formulas [2], [4], and [5] 
decide whether the following substituents occupy equatorial or axial 
positions. 

a. the C-19 methyl in cholestane 

b. the C-5 hydrogen in cholestane 

c. the C-3 hydroxyl in cholesterol 

d. the C-3 hydroxyl in cholic acid 

e. the C-7 hydroxyl in cholic acid 



general index 



Abietic acid, 781 

Absolute configuration, of alanine, 383 

of glyceraldehyde, 382 

of lactic acid, 382 

of natural amino acids, 383 

of optical isomers, 381-384 

of tartaric acid, 382 

X-ray determination of, 382 
Acetaldehyde, acetal from, 288 

from ethyne, 279 

with hydrogen chloride, 289 

infrared spectrum of, 334-335 

with methanol, 284 

physical properties of, 277 
Acetaldol, from acetaldehyde, 306 

dehydration of, 306 
Acetals, from carbohydrates, 404 

formation of, 283-287 

hydrolysis of, 287 
Acetamide, physical properties of, 435 
Acetanilide, infrared spectrum of, 436 

nitration of, 606 
Acetate radicals, 618 
Acetic acid, 190 

acid dissociation constant of, 315, 
331 

in biogenesis, 792 

conjugate acid of, 339 

with ethanol, 254 

infrared spectrum of, 334-335 

with methanol, 253-254 

physical properties of, 331 
Acetoacetic acid, decarboxylation of, 

341 
Acetoacetic ester condensation (see 

Claisen condensation) 
Acetoacetic ester synthesis, 361 
Acetone, cyanohydrin of, 282-283, 285 

diacetone alcohol from, 307-308 



halogenation of, 303-304 
from isopropyl alcohol, 259 
photodissociation, 703 
physical properties of, 277 
from propyne, 121 
Acetonitrile, hydrolysis of, 344 
Acetophenone, formation, 565 
Acetoxy radicals, 618 
Acetyl chloride, with methanol, 252- 

253 
Acetyl pernitrite, in air pollution, 58 
Acetyl radicals, from acetone, 703 
Acetylacetone (see 2,4-Pentanedione) 
Acetylcholine, 506 
Acetylcholinesterase, 506 
Acetylene (see Ethyne) 
N-Acetyl-D-glucosamine, 412 
N-Acetylimidazole, 500 
Acid anhydrides (see Carboxylic acid 

anhydrides) 
Acid catalysis, in organic systems, 

497-498 
Acid chlorides (see Acyl chlorides) 
Acid dissociation constants, definition 

of, 13 
Acid halides (see Acyl halides) 
Acid strengths, and reactivity in S N 

reactions, 202 
Acrolein, with hydrogen chloride, 312 
Acrylic acid, hydration of, 355 

with hydrogen bromide, 355 
Acrylonitrile, polymer from, 747 

polymerization, 755 
Acyl cations, 565 

Acyl chlorides, amides from, 345, 431- 
432, 438 

with amines, 345, 431-432, 438 
from carboxylic acids, 340 
esters from, 345 



801 



general index 802 



formation of, 340 

hydrolysis of, 344-345 

with lithium aluminum hydride, 

349 
with organometallic compounds, 

232-233 
reduction of, 279 
from thionyl chloride, 280 
Acyl halides, in acylation of arenes, 
565 
with alcohols, 252-253 
Acylation, of amines, 431-432 
of arenes, 565-566 
of benzene, 651 
of heterocyclic compounds, 674- 

677 
of naphthalene, 575-576 
of pyridine, 679 
Acyloin condensation, 717 
1,4 Addition (see Conjugate addition) 
Addition polymers, 753-756 
Adenine, in DNA, 480 

prebiotic synthesis of, 486 
Adenine deoxyriboside, 480 
Adenosine, in ATP, 408 
in NAD®, 506 
structure of, 407-408 
Adenosine diphosphate (ADP), from 
ATP, 529 
ATP from, 509 
Adenosine triphosphate (ATP), in bio- 
genesis, 792 
hydrolysis of, 509, 529 
in oxidative phosphorylation, 509 
structure of, 408 
synthesis from ADP, 509 
Adipic acid, acid dissociation constant 
of, 357 
cyclopentanone from, 358 
nylon from, 747, 750 
physical properties of, 357 
preparation, 750 
Adrenal cortical hormones, 788 
Adrenaline, 658-659 
Aglycone groups, 407, 415 
AIBN (see 2,2'-Azobis [2-methylpro- 

panonitrile]) 
Alanine, formation of, 343 
physical properties of, 459 
prebiotic synthesis, 486 
Alanylcysteinylserine, 469 
Alcohols, acidic and basic properties, 



13, 251-255 
with acyl halides, 252-253 
from alkenes, 249 
from alkyl halides, 249 
from carbonyl compounds, 250 
with carboxylic acids, 252-255 
from cyclic ethers, 234 
dehydration of, 256-258 
as derivatives of water, 9 
by haloform reaction, 306 
hydration of, 249 
hydroboration of, 249 
hydrogen bonding in, 246-249 
with hydrogen halides, 255 
with Lucas reagent, 255 
nomenclature of, 187-190 
from organomagnesium com- 
pounds, 230-233, 250 
oxidation of, 259-260, 278, 306, 

336 
physical properties of, 245-247, 

331 
polyhydroxy, 260-262 
preparation of, 249-251 
reactions involving the C— O 

bond, 255-258 
reactions involving the O— H 

bond, 251-255 
reactions of, 251-262 
resolution of, 385 
spectroscopic properties of, 247- 

249 
with sulfuric acid, 256 
unsaturated, 262 
water solubility of, 246 
Alcohol :NAD oxidoreductase, 506 
Aldehydes, 282-284 

from alcohols, 259, 279 
with alcohols, 283-287 
aldol addition to, 306-309 
from alkenes, 279 
with amines, 288-289, 433 
aromatic, 656-657 
Cannizzaro reaction, 293 
from carboxylic acids, 279-280 
C— H stretching frequency, 294 
with Fehling's solution, 293 
by formylation, 652-653 
from 1,2-glycols, 279, 280-281 
halogenation of, 303-306 
hydration of, 285 
hydrogenation of, 290 



general index 803 



from nitriles, 280, 349 
nmr spectra of, 278, 294 
nomenclature of, 275-276 
with organomagnesium com- 
pounds, 232 
oxidation of, 292, 336 
physical properties of, 277 
polymerization of, 287-288 
preparation of, 278-281, 294 
purification of, 294 
reactions of, 281-294 
with sodium bisulfite, 294, 527 
spectroscopic properties of, 277- 

278 
tests for, 294-295 
a,/3-unsaturated, 311-312 
uv spectra of, 334 
Aldol addition, 306-309, 499 

to formaldehyde, 752 
Aldols, dehydration of, 309 
Alizarin (see 1 ,2-Dihydroxyanthra- 

quinone) 
Alkadienes, complexes with metals, 
235 
naming of, 82 
Alkaloids, 682-684 
Alkanes, 46-74 

from alkylboranes, 97 
bromination of, 119-120 
from carbonyl compounds, 291 
chemical reactions of, 56-61 
combustion of, 56-58 
isomers of, 48 
nitration of, 61, 442 
nomenclature of, 47 
in petroleum, 56 
physical properties of, 53-56 
structure of, 19 
substitution of, 59-61 
Alkenes, in alkylation of benzene, 564 
from ammonium hydroxides, 428 
chemical reactions of, 86-103 
cis-trans isomerism of, 84 
complexes with metals, 235 
configuration of, 84 
conjugated, bonding in, 127-147 
electrophilic addition to, 87-96 
hydration of, 497 
hydroboration of, 96-97 
hydroxy lation of, 286 
nomenclature of, 81-83 
oxidation of, 96-99, 279, 336, 562 



photorearrangement of cis-trans 
isomers, 707 

polymerization of, 99-103 

radical addition to, 94-96 

with singlet oxygen, 706 

structure of, 19 

from vinylboranes, 115 
Alkenyl groups, naming of, 82 
Alkoxide ions, 251 

oxidation of, 260 
Alkoxysilanes, 532 
Alkyl bromides, from alcohols, 255 

from carboxylic acids, 342 

from phosphorus tribromide, 256 
Alkyl cations (see Carbonium ions) 
Alkyl chlorides, physical properties of, 
331 

from thionyl chloride, 255-256 
Alkyl chlorosulfite, from thionyl- 

chloride, 256 
Alkyl groups, naming of, 49-51 

reactivity in El reactions, 208 

reactivity in E2 reactions, 206- 
207 

rearrangement of, 258, 280-281 

in S N reactions, 202 
Alkyl halides, 217-226 

from alcohols, 255-256 

with alkoxides, 252 

in alkylation of benzene, 564 

displacement reactions of, 193- 
194 

with enolate anions, 310 

hydrolyses of, 249 

with metals, 228 

nomenclature of, 187-190 

from organomagnesium com- 
pounds, 230-231 

physical properties of, 217 

preparation of, 218-219 

reactions of, 219 

solubility of with silver, 203 

spectra of, 217-218 

uv spectra of, 217-218 
Alkyl hydrogen sulfates, alkenes from, 
256 

displacement of, 256 

formation of, 256 
Alkyl sulfates, with alkoxides, 252 
Alkyl sulfides (see Thioethers) 
Alkylation, of amines, 428 

of arenes, 564-565 



general index 804 



of benzene, 651 
of carbonyl compounds, 315 
of esters, 354 
of ketones, 499 
of nitrites, 440 
of phenols, 631-632 
of pyridine, 679 
Alkylbenzenes, oxidation, 649 
radical halogenation, 650 
synthesis, 564-566 
Alkylidenephosphoranes, formation, 
530 
reactions, 531 
Alkynes, 111-121 

acidity of, 116-117 

addition reactions of, 1 1 3-1 1 5 

hydration of, 114, 279 

hydroboration of, 115 

hydrogenation of, 726 

naming of, 111 

nucleophilic addition to, 115 

oxidation of, 114 

oxidative coupling, 726 

physical properties of, 112 

with silver ammonia solution, 

116-117 
structure of, 19 
Alkynide salts, 116-117 
Alkynyl groups, naming of, 111 
Allene, 83 

Allenes, optical isomerism of, 379 
Ally] alcohol, reactivity of, 262 
Allyl anion, resonance structures of, 

139 
Allyl bromide (see 3-Bromopropene) 
Allyl cation, resonance structures of, 

138 
Allyl chloride 189, 220-221 
Allyl halides, preparation of, 220 

reactivity in S N reactions, 221-222 
o-Allylphenol, by Claisen rearrange- 
ment, 632 
from phenol, 631 
Aluminum chloride, in alkylation of 

arenes, 564 
Aluminum oxide, in alcohol dehydra- 
tion, 257 
American Chemical Society, 52 
Amide group, resonance in, 434 
Amides, 434-440 

acidic properties of, 437 
from acyl chlorides, 345 



amines from, 439 

from anhydrides, 345 

basic properties of, 438 

dehydration of, 280 

from esters, 345 

hydrogen bonding in, 435 

hydrolysis of, 344-345, 497 

infrared spectra of, 435 

with lithium aluminum hydride, 
349 

nmr spectra of, 435 

with organomagnesium com- 
pounds, 232 

physical properties of, 435 

preparation of, 438-439, 468 

protonation of, 438 

reactions of, 439 

reduction of, 430, 439 

spectral properties of, 435 

uv spectra of, 437 
Amination, of pyridine, 679 
Amine oxides, formation of, 433, 523 

optical activity of, 434 

rearrangement of, 434 
Amines, 421-434 

with acid derivatives, 438 

with acids, 431 

as acids and bases, 13, 426-427, 

614-615 
acylation of, 431^432 
aromatic, 614-620 
basic properties, 614-615 
in cancer therapy, 429 
carcinogenic properties, 620 
as derivatives of ammonia, 9 
diazonium salts from, 591 
halogenation of, 432 
hydrogen bonding in, 423-425 
infrared spectra of, 424-425 
naming of, 190 
from nitriles, 280 
with nitrous acid, 432, 616-617 
nmr spectra of, 425 
nomenclature of, 421-423 
oxidation of, 432-433 
physical properties of, 423-425, 

614 
preparation of, 428—43 1 
reactions of, 431-434 
by reduction, 430 
spectroscopic properties of, 423- 
425 



general index 80S 



stereochemistry of, 425-426 
Amino acid sequence, in peptides, 470 
Amino acids, 457-467 

acid-base properties of, 458-460 
acidic and basic types, 457 
analysis of, 463-467 
configuration of, 457 
essential, 458 
lactams from, 715 
ninhydrin test, 463-464 
with nitrous acid, 463 
synthesis of, 458 
Amino sugars, 401-402 
m-Aminoanisole, formation, 595 
4-Aminobiphenyl, carcinogenic pro- 
perties, 620 
,/3-Aminoethanesulfonic acid, 784 
2-Aminoethanol, 421 
l-Amino-4-hydroxyanthraquinone,759 
Aminomalorionitrile, from hydrogen 

cyanide, 486 
2-Aminonaphthalene, carcinogenic 

properties, 620 
p-Aminophenol, acid dissociation con- 
stant, 629 
diazotization, 618 
uv spectrum, 629 
2-Aminopyridine, from pyridine, 678 
Ammonia, acid dissociation constant 
of, 13, 251 
base dissociation constant of, 13, 

423 
bond angles in, 8 
derivatives of, 9 
physical properties of, 10, 423 
Ammonium cyanate, 3 
Ammonium hydroxides, formation of, 
428 
thermal decomposition of, 428 
Ammonium salts, nomenclature of, 
421-423 
in S N reactions, 194 
Ammonium sulfide, in reduction of 

nitro compounds, 609 
Ammonium thioglycolate, 747 
Amygdalin, structure and occurrence, 

657 
Amylopectin, 412 
Amylose, 412 
^-Amyrin, 781, 789 
Analysis, of amino acids, 463-467 
of peptides, 469^*70 



Anchimeric assistance, 502 
Androsterone, 791 
Angle strain, in cycloalkanes, 67-68 
Anhydrides (see Carboxylic acid an- 
hydrides) 
Aniline, acid dissociation constant, 614 
base dissociation constant of, 423, 

427, 614 
bromination, 616 
from bromobenzene, 595 
derealization in, 427 
from nitrobenzene, 609 
physical properties of, 423 
resonance hybrid of, 615 
stabilization energy of, 605 
tautomer of, 427, 605 
uv spectrum, 556, 614 
Anilinium ion, resonance hybrid of, 

614 
Anisaldehyde, in benzoin condensa- 
tion, 656 
Anisole, cleavage of, 632 
formation of, 630 
physical properties of, 263 
[10] Annulene, structure of, 725 
[18] Annulene, synthesis of, 726 
Annulenes, 725-726 
Anomeric effect, 407 
Anomers, of glucose, 404-405 
Anthocyanidins, 687 
Anthocyanins, 687 
Anthracene, bond lengths in, 574 
monosubstitution products, 551 
physical properties, 553 
reactions, 577 
resonance hybrid of, 575 
uv spectrum of, 557 
Antibiotics, 4, 527 

streptomycin, 402 
Antigen, composition of, 413 
Antimony pentafluoride, in super 

acids, 13 
Antipellagra factor, 684 
Antipernicious anemia factor, 681-682 
L-Arabinose, structure of, 400-401 
Arenes, 549-578 

acylation of, 565-566 
alkylation of, 564-565 
complexes with halogens, 562-563 
deuteration of, 567 
halogenation of, 562-563, 
nitration of, 561-562 



general index 806 



nmr spectra of, 558 
nomenclature of, 549-552 
physical properties of, 553 
reactions of, 559-577 
spectroscopic properties of, 554- 

558 
sulfonation of, 566 
Arginine, physical properties, 461 
Aromatic amines {see Arylamines) 
Aromatic halides {see Aryl halides) 
Aromatic hydrocarbons {see Arenes) 
Aromatic side-chain derivatives, 648- 

665 
Aromatic substitution {see Electro- 
philic aromatic substitution, Nucleo- 
philic aromatic substitution, etc.) 
Aryl cations, from diazonium com- 
pounds, 619 
Aryl groups, naming of, 550 
Aryl halides, 589-598 

from diazonium salts, 591 
infrared spectra, 590 
physical properties of, 590 
preparation of, 590-592, 618-619 
reactions of, 592-596 
reactivity of, 589 
Aryl halogen compounds {see Aryl 

halides) 
Aryl nitrogen compounds, 605-620 
Aryl oxygen compounds, 627-643 
Arylamines, from nitro compounds, 

592 
Arylhydrazines, from diazonium salts, 

619 
Arylmethyl halides, physical proper- 
ties, 652 
Ascaridole, 780 
Ascorbic acid, pK H A of, 418 

structure of, 413 
Asparagine, physical properties, 461 
Aspartic acid, physical properties, 

461 
Asphalt, from petroleum, 58 
Aspidospermine, mass spectrum, 776 

structure of, 775 
Aspirin, 658 

Asymmetric induction, 386-387 
Asymmetric synthesis, 386-387 

in biochemical systems, 387 
Asymmetry, and optical activity, 372- 

374 
Atherosclerosis, 783 



Atomic orbital models, of 1,3-buta- 
diene, 142 

of pyrrole, 672-673 

of trimethylenemethyl, 142 
ATP {see Adenosine triphosphate) 
Atropa belladonna, atropine from, 684 
Atropine, 684 
Autoxidation, of ethers, 265 

nitroarene catalysis in, 613 
Avogadro, A., 3 
Axial positions and substituents {see 

Cyclohexane) 
Azelaic acid, from civetone, 770 
Azides, as 1,3-dipoles, 721-723, 731 

reduction of, 430 
Azines, formation of, 289 
Azo compounds, from diazonium 
compounds, 443 

from hydrazines, 443 
Azobenzene, 443 

from nitrobenzene, 61 1 
2,2'-Azobis(2-methylpropanonitrile) 
(AIBN), 443 

initiator in radical polymerization, 
754 
Azomethane, 422, 443 
Azoxybenzene, 610 
Azulene, rearrangement, 578 

resonance hybrid of, 578 



Bacteriophage DNA, 484 
Bakelites, 751-752 
Barbaralane, valence tautomers, 724 
Barbituric acid, 685 

tautomers of, 723 
Barbituric acids, synthesis of, 686 
Base catalysis, in organic systems, 499 
Base ionization constant, definition of, 

13 
Beckmann rearrangement, 429-430,439 
Benzal chloride, 651 

physical properties, 653 

radical chlorination, 650 
Benzaldehyde, acetals from, 287 

in benzoin condensation, 656 

benzoin from, 656 

phenylalanine from, 458 

uv spectrum, 556 
Benzedrine, 659 
Benzene, acylation, 565-566 

alkylation, 564 



general index 807 



bonding in, 129-133 

bromination, 128, 563 

deuteration, 567 

excited singlet state, 701 

heat of combustion of, 129 

nitration, 592 

physical properties, 553 

resonance in, 549 

resonance structures of, 132 

shape of, 127 

stabilization energy of, 128-129 

structure, 549 

substitution reactions of, 560 

sulfonation, 566 

thiophene in, 678 

uv spectrum, 556-557 
Benzene-de , formation of, 567 
Benzenediazocyanide, 618 
Benzenediazonium chloride, 617 

diazo coupling of, 620 
Benzenediazonium cyanide, 618 

isomers of, 618 
Benzenesulfonic acid, formation of, 566 

phenol from, 627 
Benzhydrol, in photoreduction, 705 
Benzhydrol radicals, formation and 

dimerization, 704 
Benzhydryl chloride, physical proper- 
ties, 653 
Benzidine, basicity of, 614 

carcinogenic properties, 620 

from hydrazobenzene, 611 

uv spectrum, 614 
Benzidine rearrangement, 61 1 
Benzilic acid, from benzil, 313 

rearrangement of, 313, 321 
Benzocaine, 659 
Benzofuran, 679 

Benzoic acid, acid dissociation con- 
stant of, 331 

formation of, 649 

physical properties of, 331 
Benzoic acids, acid dissociation con- 
stants of, 661 
Benzoin condensation, 656-657 
Benzoin, formation of, 656-657 
Benzophenone, photoreduction of, 
704-705 

as photosensitizer, 706 

triplet state of, 705 
Benzopinacol, by photoreduction, 
704-705 



/>-Benzoquinone, additions to, 639-640 

stabilization energy of, 636 
Benzoquinuclidine, basicity of, 622 
Benzotrichloride, physical properties 
of, 653 

reactions of, 651 

from toluene, 650 
Benzotrifluoride, physical properties 

of, 653 
Benzoyl peroxide, initiator in radical 

polymerization, 754 
Benzyl bromide, 589 

physical properties of, 653 
Benzyl cation, resonance stabilization 
of, 653 

resonance structures of, 589 
Benzyl chloride, 

nitration of, 568 

physical properties of, 653 

radical chlorination of, 650 

reactions of, 651 
Benzyl fluoride, physical properties of, 

653 
Benzyl halides, reactivity of, 

589 
Benzyl iodide, physical properties of, 

653 
Benzyl radical, resonance hybrid of, 

650 
Benzyloxycarbonyl group, 471 
Benzyne, 596, 627 
Betaines, definition of, 531 
Biacetyl, formation of, 703 
Bicarbonate ion, resonance structures 

of, 138 
Bile acids, 784 
Bile pigments, 681 
Biogenesis, of cholesterol, 792-795 

of lanosterol, 792-795 

of squalene, 792-795 

of terpenes and steroids, 789-794 
Biphenyl, 550 

formylation of, 652 

uv spectrum of, 556 
Biphenyls, optical isomerism of, 380 

structure of, 380 
Blanc rule, 786 

Boat form, of cyclohexane, 64 
Bond angles, in dimethyl ether, 10 

in methanol, 10 

strain in cycloalkanes, 67-68 

in water, 7-8 



general index 808 



Bond energies, conjugation effects on, 
24 

relative order of C— H, 60 

of silicon compounds, 532 

table of, 24-25 
Bond lengths, in anthracene, 574 

in benzene, 549 

boron-nitrogen, 537 

in ethane, 35 

in ethene, 35 

in ethyne, 35 

in naphthalene, 574 
Bonding, in boron hydrides, 537 

in conjugated unsaturated sys- 
tems, 127-147 

the covalent bond, 5 

d orbitals in, 517-520 

delocalized, 131 

in double bonds, 1 30 

electron-deficient, 92 

in ethene, 131 

in ferrocene, 235 

the ionic bond, 5 

in organic compounds, 5 

in organometallic compounds, 
227, 230 

polarity of, 6 

three-center, 538 
Bonds, ir-type, 34 
Bonds, cr-type, 34 
Borazines, structure of, 537 
Boric acid, 536 
Boron, organic compounds of, 536- 

540 
Boron hydrides, addition to alkenes, 
96-97 

bonding in, 538 
Boron trifiuoride, 536 

diborane from, 97 

etherate of, 264 
Bromination, of anthracene, 577 

of benzene, 128, 563 

of phenanthrene, 577 
Bromine, addition to alkynes, 113, 114 

addition to cyclohexene, 89 

addition to 4-methyl-2-hexene, 86 

addition to multiple bonds, 37 

with benzene, 128, 563 

with 1,5-hexadiene, 127 

with 2,4-hexadiene, 127 

reaction with alkanes, 60-61 
Bromobenzene, 589 



from benzene, 128 
chlorination of, 590 
nitration of, 568 
physical properties of, 590 
with potassium amide, 595 
o-Bromochlorobenzene, preparation 

of, 590 
/»-Bromochlorobenzene, preparation 

of, 590 
Bromocyclohexane, conformations of, 

67 
2-Bromocyclohexanone, formation of, 

303 
Bromoethane (see Ethyl bromide) 
Bromoform, from methyl ketones, 

305-306 
1-Bromohexane, with ethylene oxide, 

265 
Bromomethane (see Methyl bromide) 
l-Bromo-3-methylbutane, from 3- 

methyl-1-butyne, 119 
2-Bromo-2-methylbutane, from 2- 

methylbutane, 61 
solvolysis of, 208 
2-Bromo-3-methyl-l-butene, from 3- 

methyl-1-butyne, 119 
l-Bromo-2-methylpropane (see Iso- 

butyl bromide) 
4-Bromo- 1 -naphthylamine, diazotiza- 

tion of, 591 
p-Bromonitrobenzene, with methoxide, 

594 
Bromonium ions, 91, 134 
l-Bromo-2-phenyl-l-propene, by E2 

elimination, 707 

photoisomerization of, 707 
1-Bromopropane (see Propyl bromide) 
2-Bromopropane (see Isopropyl bro- 
mide) 
2-Bromopropanoic acid, substitution 

of, 343-344 
3-Bromopropanoic acid, from acrylic 

acid, 355 
3-Bromopropene, 589 

with sodium phenoxide, 631 
Bromotoluenes, 550 
Brucine, 683 

as resolving agent, 385 
Bullvalene, valence tautomers of, 725 
1,2-Butadiene, 82 
1,3-Butadiene, 82 

conformations of, 133, 719 



general index 809 



conjugate addition to, 134, 221, 

' 719 

cycloaddition to, 729 

with diethyl maleate, 720 

dimerization of, 720 

7r-electron systems of, 141-142 

excited state of, 1 68, 701 

hexamethylenediamine from, 751 

with hydrogen chloride, 221 

molecular orbitals of, 698, 727- 
729 

nylon from, 75 1 

resonance structures of, 136 
Butadiyne, 111 

with methanol, 115 
Butane, from ethyne, 119 

isomers of, 48 

in natural gas, 57 

physical properties of, 55, 63, 83 
1,4-Butanediol, physical properties of, 

261 
2,3-Butanediol, optical isomers of, 392 
2,3-Butanedione, 313, 703 
Butanethiol, occurrence, 522 
Butanoic acid, 190 

acid dissociation constant of, 331 

physical properties of, 331 

synthesis of, 354 
1-Butanol (see w-Butyl alcohol) 
2-Butanol (see i-Butyl alcohol) 
2-Butanone (see Methyl ethyl ketone) 
1,2,3-Butatriene, 82 
2-Butenal, 306 
1-Butene, infrared spectrum of, 163 

naming of, 81 

physical properties of, 84 
2-Butene, geometric isomers of, 84 
c/.y-2-Butene, heat of combustion of, 85 

physical properties of, 84-85 
trans-2-Butene, heat of combustion of, 
85 

physical properties of, 84-85 
2-Butene ozonide, 97 
cw-Butenedioic acid (see Maleic acid) 
/raay-Butenedioic acid (see Fumaric 

acid) 
2-Butenoic acid, from 3-hydroxybu- 

tanoic acid, 356 
3-Butenoic acid, acid dissociation con- 
stant of, 331 

physical properties of, 331 

reduction of, 340 



3-Buten-l-ol, formation of, 340 
Butenyne, 111 

from ethyne, 115 
hydrogenation of, 119 
f-Butoxy carbonyl group, 472 
n-Butyl alcohol, physical properties of, 

247 
.y-Butyl alcohol, asymmetry of, 372 
from 2-butanone, 373 
physical properties of, 247 
projection formulas of, 374 
/-Butyl alcohol, in acetal formation, 
287 
from /-butyl chloride, 197, 199, 

206-207 
manufacture of, 251 
from 2-methylpropene, 88 
physical properties of, 247, 535 
Butyl alcohols, vapor-phase chromato- 

gram of, 155 
/-Butyl bromide, equilibration with 
isobutyl bromide, 94 
from propene, 93 
Butyl carbitol, physical properties of, 

264 
n-Butyl chloride, 53 
/-Butyl chloride, with aqueous base, 

197 
/-Butyl compounds (see Isobutyl com- 
pounds) 
/-Butylamine, base dissociation con- 
stant of, 423 
oxidation of, 433 
physical properties of, 423 
/-Butylbenzene, nitration of, 568 

physical properties of, 553 
/-Butylcarboxaldehyde, acetals from, 

287 
r-Butylcyclohexane, formation of, 635 
4-/-Butylcyclohexyl chloride, confor- 
mations of, 71 

cis-trans isomers of, 71 
4-/-Butyl-4-isopropyldecane, naming 

of, 52 
/-Butylphenol, hydrogenation of, 635 
2-Butylpyridine, from pyridine, 678 
2-Butyne, naming of, 111 
Butyric acid (see Butanoic acid) 
y-Butyrolactam, 467 

from ethyl 4-aminobutanoate, 468 
formation of, 715 
y-Butyrolactone, 467 



general index 810 



formation of, 715 

from 3-hydroxybutanoic acid, 356 



Cahn-Ingold-Prelog system, for con- 
figuration, 382 
Calcium carbide, as a salt of ethyne, 1 1 6 
Camphene, from a-pinene, 798 
Camphor, 780 

in molecular-weight determina- 
tion, 157, 780 
from a-pinene, 798 
as plasticizer, 780 
synthesis of, 780 
Camphor- 10-sulfonic acid, as resolving 

agent, 385 
Cancer therapy, drugs in, 429 
Cannizzaro, S., 3 
Cannizzaro reaction, 294 

acids and alcohols from, 650 
ofglyoxal, 313, 321 
Capell, L. T., 552 

e-Caprolactam, polymerization of, 751 
Carbanions, in autoxidation, 613 
methyl, 32 

in nucleophilic addition, 115 
in polymerization, 100 
stable triarylmethyl, 654 
Carbene, from chloroform, 223 
from diazomethane, 224 
insertion reactions of, 224 
from iodomethylzinc iodide, 224 
Carbenes, methylene, 32-33 
Carbitols, definition of, 264 

from ethylene oxide, 266 
Carbohydrates, 399-415 

classification of, 400^02 
with immunological specificity, 
413^114 
Carbon- 14, in biogenesis experiments, 

792 
Carbon dioxide, with organomag- 

nesium compounds, 232 
Carbon monoxide, air pollution from, 
58 
in methanol formation, 251 
poisoning, 681 
Carbon tetrachloride, dipole moment 
of, 8 
as a fire extinguishing fluid, 222 
fluorination of, 224 
as an infrared solvent, 217 



infrared spectrum of, 217-218 
from methane chlorination, 28, 

60, 222 
phosgene from, 222 
physical properties of, 222, 535 
reactivity in S N reactions, 223 
with sodium, 222 
toxicity of, 222 
Carbonium ions, in alkylation of 
arenes, 564 
methyl, 32 

in polymerization, 101 
rearrangements of, 208, 258, 280- 

281, 565 
relative stability of, 94 
stable triarylmethyl, 653-654 
Carbonyl compounds, from alkenes 
by ozonization, 98 
with amines, 288-289 
with organometallic compounds, 

231-233, 250 
oxidation of, 293 
reduction of, 250, 290-292 
structures of, 252 
unsaturated, 310-313 
Carbonyl groups, bond energy of, 275 
effect of chemical shift, 278 
electronic transitions in, 277-278 
equilibrium in additions to, 275, 

277, 282 
with hydrogen cyanide, 282-283 
infrared stretching frequencies for, 

278 
polarity of, 277, 278 
reactivity of, 275-277, 281 
Carbowax (see Polyethylene glycol) 
Carboxyl group, inductive effect of, 

356-357 
Carboxylate radicals, decarboxylation 
of, 342 
from diacyl peroxides, 342 
in Hunsdieker reaction, 342 
in Kolbe electrolysis, 342 
Carboxylate salts, solubility of, 332 
Carboxylic acid anhydrides, in acyla- 
tion of arenes, 565 
amides from, 345 
with amines, 431-432, 438 
esterification of, 345 
hydrolysis of, 344-345 
with lithium aluminum hydride, 
349 



general index 811 



Carboxylic acids (see also Dicarboxylic 
acids), 329-343 

acid ionization constants of, 329, 
336-337 

acidity of, 336-339 

from alcohols, 259, 336 

from aldehydes, 292, 336 

from alkenes, 336 

from alkylmalonic esters, 354 

basic properties of, 338-339 

decarboxylation of, 341-342, 354 

derivatives of, 344 

from 1,2-glycols, 336 

by haloform reaction, 306 

halogenation of, 342-343 

from hydrocarbons, 336 

hydrogen bonding in, 330 

hydroxy, 356 

by malonic ester synthesis, 336 

from methyl ketones, 292 

from nitriles, 336 

nmr spectra of, 334 

nomenclature of, 190-191, 329 

odors of, 333 

from organomagnesium com- 
pounds, 232-233, 336 

with phosphorus halides, 340 

physical properties of, 330-334 

preparation of, 336, 649 

reactions of, 339-343 

reduction of, 279-280, 340 

solubility of, 331-332 

spectra of, 334-335 

with thionyl chloride, 340 

unsaturated, 355-356 

uv spectra of, 334 
Carboxylic ester formation, 252-255 

acid catalysts in, 253-254 

equilibrium in, 254-255 

mechanism of, 252-255 

steric hindrance in, 255 
Carboxylic ester interchange, 345 

acid-catalyzed, 348-349 

base-catalyzed, 348 

in condensation polymerization, 
749 
Carboxylic esters, from acid chlorides, 
345 

acidic properties of, 350-351 

from alcohols, 252-255 

alkylation of, 354 

amides from, 345, 438 



with amines, 345, 438 
from anhydrides, 345 
anions from, 350-351 
condensation of, 352-354 
formation and hydrolysis of, 497- 

498 
hydrolysis of, 344-345, 497, 500 
with lithium aluminum hydride, 

349 
naming of, 191 

with organomagnesium com- 
pounds, 232-233 
reduction of, 349 

Carcinogenic compounds, aromatic 
amines, 620 
polynucleaf hydrocarbons, 553 

j3-Carotene, 699-700, 782 

Catalase, properties and function of, 
478 

Catalysis, in acetal formation, 286 
in alcohol dehydration, 257 
in alkyl chloride formation, 255 
in condensation reactions, 288 
in cyanohydrin formation, 283 
definition of, 495 
in dehydration of alcohols, 258 
in ester formation, 253-254 
in ester interchange, 345 
general acid and base, 500-501 
in halogenation of carbonyl com- 
pounds, 303-304 
in hemiacetal formation, 284 
heterogeneous and homogeneous, 

36, 497 
in hydrogenation, 290 
intramolecular, 502 
in methanol formation, 251 
micelles in, 333 
nucleophilic, 499-500 
in organic systems, 495-503 
poisoning of, 279 
in reduction of acyl chlorides, 279 
specific-acid, 501 

Catechol, acid dissociation constant of, 
629 
structure of, 635 
uv spectrum of, 629 

Catenane, 718 

Celcon, 288 

Cellobiose, hydrolysis of, 410 
structure of, 408-409 

Cellosolves, definition of, 264 



general index 812 



from ethylene oxide, 266 
Cellulose, fibers from, 410 

hydrolysis of, 408 

structure of, 410 
Cellulose acetate, 410 
Cellulose nitrate, 410 
Cellulose xanthate, 410 
Cetane, heat of combustion, 510 
Chain reaction, in methane chlorina- 

tion, 31 
Chain termination, in radical poly- 
merization, 754-755 
Chair form, of cyclohexane, 64-67 
Charge-transfer complexes, of halo- 
gens with arenes, 562-563 

of polynitro compounds, 612-613 

quinhydrones, 638 
Chelate ring, in phenols, 634 
Chemical Abstracts, 52 
Chemical evolution, 486-487 
Chemical shift {see also Nuclear mag- 
netic resonance spectroscopy) 

of enol-OH groups, 315 
Chirality, 382 

definition of, 372 
Chitin, 412 
Chloral hydrate, 261 
Chlorambucil, 429 
Chlorination, of chloromethane, 28-30 

of methane, 28-32 

photochemical, 28-32, 703 
Chlorine, absorption of light by, 30 

with alkanes, 59-61 

with methane, 28 
a-Chloro ethers, from aldehydes, 289 
Chloroacetic acid, acid dissociation 
constant of, 331, 338 

glycine from, 458 

physical properties of, 331 
Chloroacetone, formation of, 303 
o-Chloroanisole, amination, 595 
Chlorobenzene, nitration of, 568-569, 
571, 606 

phenol from, 595, 627 

physical properties of, 590 
2-Chloro- 1,3 -butadiene, polymer from, 

746 
1-Chlorobutane {see w-Butyl chloride) 
2-Chlorobutanoic acid, acid dissocia- 
tion constant of, 331 

physical properties of, 331 
3-Chlorobutanoic acid, acid dissocia- 



tion constant of, 331 
physical properties of, 331 
4-Chlorobutanoic acid, acid dissocia- 
tion constant of, 331 
physical properties of, 331 
l-Chloro-2-butene, from 1,3-buta- 

diene, 221 
3-Chloro-l-butene, from 1,3-buta- 

diene, 221 
Chlorocyclohexane, conformations of, 

67 
rra/K-2-ChlorocyclohexanoI, from 

cyclohexane, 88 
Chlorocyclopentane, 62 
l-Chloro-2,4-dinitrobenzene, with di- 

ethylamine, 593 
4-Chloro-2-ethyl-2-buten-l-ol, 187 
l-Chloro-2-fluoro-l , 1 ,2,2-tetrabromo- 

ethane, nmr spectrum of, 177-178 
Chloroform, 60 

as an anesthetic, 222 

ethanol in, 222 

fluorination of, 225 

with hydroxide ion, 223 

as an infrared solvent, 217 

infrared spectrum of, 217-218 

from methane chlorination, 60, 

222 
from methyl ketones, 305-306 
phosgene from, 222 
physical properties of, 222, 535 
solvent properties of, 222 
Chloromethane {see Methyl chloride) 
Chloromethyl cation, 652 
Chloromethylation, of benzene, 652 
of heterocyclic compounds, 674— 
677 
l-Chloro-2-methylbutane, from 2- 

methylbutane, 61 
l-Chloro-3-methylbutane, from 2- 

methylbutane, 61 
2-Chloro-2-methylbutane, from 2- 
methylbutane, 61 
solvolysis of, 208 
3-Chloro-2-methylbutane, from 2- 

methylbutane, 61 
4-Chloro-2-methylbutane-2-sulfonic 

acid, preparation of, 527 
4-Chloro-2-methyl-2-butanethiol, oxi- 
dation of, 527 
cw-3-Chloro-l-methylcyclopentane, 
with hydroxide ion, 201 



general index 813 



2-Chloro-2-methylpropane, from 2- 

methylpropene, 93 
1-Chloronaphthalene, physical proper- 
ties of, 590 
2-Chloronaphthalene, physical proper- 
ties of, 590 
3-Chloro-l-nitrobutane, 53 
Chloronium ions, 92 
2-Chloropentanoic acid, 190 
5-Chloropentanoic acid, acid dissocia- 
tion constant of, 331 
physical properties of, 331 
Chlorophyll, in photosynthesis, 399 
phytol in, 781 
as pyrrole derivative, 681 
structure of, 399 
Chloroprene (see 2-Chloro-l,3-buta- 

diene) 
1-Chloropropane (see Propyl chloride) 
2-Chloropropane (see Isopropyl chlo- 
ride) 
2-Chloropropanoic acid, from pro- 
panoic acid, 343 
3-Chloropropanoic acid, from pro- 
panoic acid, 343 
2-Chloro-l-propene, from propyne, 

121 
/3-Chloropropionaldehyde, from acro- 
lein, 286 
3-Chloro-l -propyne, in S N reactions, 

221 
m-Chlorotoluene, physical properties 

of, 590 
o-Chlorotoluene, formation of, 591 

physical properties of, 590 
p-Chlorotoluene, hydrolysis of, 595 

physical properties of, 590 
Chlorotrifiuoroethene, polymer from, 

745 
Cholanic acid, 786 
Cholestane, 785 
Cholesterol, 782-788 
Cholesterol, absolute configuration, 
788, 798 
dehydrogenation of, 787 
from lanosterol, 795 
metabolism of, 783 
methyl groups, in 786-787 
molecular formula of, 783 
nmr spectrum of, 787 
numbering system of, 785 
oxidative degradation, 783-784 



reduction of, 785 
ring sizes of, 786 
X-ray diffraction of, 787 
Cholic acid, 784 
Cholic acids (see Bile acids) 
Chromatography, in amino acid anal- 
ysis, 465-466 
gas-liquid, 153, 155-156 
liquid-liquid, 153 
liquid-solid, 153-154 
Chromic acid, in oxidation of alcohols, 

259 
Chromic oxide, in oxidation of al- 
cohols, 259 
Chromone, 687 

Chromosomes, composition of, 477 
Chymotrypsin, in ester hydrolysis, 504 

properties and function of, 478 
Cinchona alkaloids, 684 
Citronellal, 780 
Citronellol, 780 
Civetone, 769-772 
Claisen condensation, 716 

of esters, 352-354 
Claisen rearrangement, 632 
Claus, structure of benzene, 128 
Clemmensen reduction, 291-292, 718 

of civetone, 771 
Cocaine, 684 
Codeine, 684 

Codon, for phenylalanine, 485 
Codons, 483, 485 
Coenzyme A, in biogenesis, 792 
Coenzyme Q, 640 
Coenzymes, 503 

Collagen, leather from, 757-758 
Colors, complementary, 701 
Combustion, calculation of heat of, 24 
definition of, 23 

and elemental analysis, 156-157 
Complexes (see also Charge-transfer 
complexes) 
in acylation of arenes, 566 
halogens with arenes, 562-563 
7r-type, 612-613 
Conant, J. B., 681 
Condensation polymers, 748-752 
Condensation reactions, acyloin, 717 
benzoin, 656-657 
of carbonyl compounds, 288-290 
definition of, 288 
polymerization, 748-752 



general index 814 



of silanols, 535 
Conessine, 791 

Configuration, Cahn-Ingold-Prelog 
system, 382 

of cis-trans isomers, 86 

definition of, 84 

determination of, 38, 86 

dl designation of, 382 

of glucose, 402 

inversion of, 389 

at nitrogen, 425-426 

of optical isomers, 381-384 

rs designation of, 382 
Conformational analysis, and nmr 

spectroscopy, 177 
Conformations, asymmetric, 376 

of bromocyclohexane, 67 

of 4-?-butylcyclohexyl chloride, 71 

of chlorocyclohexane, 67 

of cyclohexanes, 63-67 

of decalins, 785-786 

definition of, 21, 84 

of cis- 1 ,4-di-?-butylcyclobutane, 
72 

eclipsed, 21 

of glucose, 405 

of methylcyclohexane, 65-66 

Newman convention, 22 

saw-horse convention, 22 

staggered, 21 
Conformers, definition of, 315, 369 
Congo Red, 759 
Coniine, 684 

Conjugate addition, 127, 133-135, 
718-721 

to 1,3 -butadiene, 221 

to quinones, 639-640 

to a,/3-unsaturated carbonyl com- 
pounds, 311, 355 
Conjugated dienes, electronic spectrum 
of, 166 

with singlet oxygen, 706 

stabilization of, 135-137 
Conjugated double bonds, bonding in, 
127-147 

definition of, 83 
Conjugation, and light absorption, 

699-703 
Conjugation, and spectra of carbonyl 

compounds, 277 
Conjugation, and structures for excised 
states, 168 



Conjugation, in a,/3-unsaturated car- 
bonyl compounds, 311 
Conjugation effects, in electrophilic 

aromatic substitution, 572 
Cope rearrangements, 723-724 
Copolymers, butadiene-styrene, 756 

definition of, 103 

ethene-propene, 748, 756 

vinyl chloride- vinyl acetate, 748, 
756 
Coprostane, 785 

oxidation, 786 
Corpus luteum, hormones from, 790 
Cortisone, 791 

synthesis, 788 

therapeutic uses, 788 
Cotton, cellulose in, 410 

dyeing of, 759, 760 
Cotton effect, 390 
Coumarin, 687 
Coumarins, occurrence and synthesis, 

687 
Couper, A. S., 3 
Covalent bonding (see also Bonding), 5 

and polarity, 6 
Covalent catalysis, 500 
Cracking, of kerosene, 58 
p-Cresol, acid dissociation constant, 
629 

uv spectrum, 629 
Cresols, from />-chlorotoluene, 595 

physical properties of, 634-635 
Crick, F. H. C, 428 
Crotonaldehyde, from acetaldehyde, 

306 
Crotonic acid (see 2-Butenoic acid) 
Crotyl chloride (see l-Chloro-2-bu- 

tene) 
Crystal Violet, resonance hybrid, 702 
Crystallites, of polymers, 739-741 
Crystallization, solvents for, 153 
Cumene (see Isopropylbenzene) 
Cumulated double bonds, definition 
of, 83 

in ketenes, 312 
Cupric acetylacetonate, 316 
Cuprous salts, in ethyne dimerization, 

114, 115, 119 
Curtius reaction, 430 
Cyanic acid, 422 

Cyanide ion, in cyanohydrin forma- 
tion, 283 



general index 815 



Cyano compounds, synthesis of, 618- 

619 
Cyanoacetic acid, 440 

acid dissociation constant of, 331 

decarboxylation of, 341 

physical properties of, 331 
Cyanocobalamin (see Vitamin Bi 2 ) 
Cyanohydrin formation, 282-283 

by 1,4-addition, 311 

equilibrium in, 284 
2-Cyanopropanoic acid, formation of, 

343 
p-Cyanotoluene, synthesis of, 619 
Cyclization reactions, 715-730 

of alkynes, 721 

of carbonyl compounds, 715-718 
Cycloaddition reactions, 718-723 

of 1,3-butadiene, 729 

cis-1,1, 224 

of ethene, 729 

and orbital symmetry, 726-730 

photochemical, 707-708 
Cycloalkanes, 62-74 

angle strain in, 67-68 

chemical properties of, 68-70 

heats of combustion of, 68 

cis-trans isomerism of, 69-71 

naming of, 62 

physical properties of, 63 
Cyclobutane, from cyclobutyl bro- 
mide, 230 

heat of combustion of, 68 

physical properties of, 63, 83, 226 
Cyclodecane, heat of combustion of, 

68 
Cycloheptane, heat of combustion of, 
68 

physical properties of, 63 
1,3,5-Cycloheptatriene, oxidation of, 
641 

X-irradiation of, 668 
Cycloheptatrienyl cation, resonance 

structures of, 144 
Cycloheptatrienyl radical, epr spec- 
trum, 668 
1,3-Cyclohexadiene, with tetracyano- 

ethylene, 719 
1,4-Cyclohexadiene, formation of, 721 
Cyclohexane, axial positions in, 65-67 

boat form of, 64 

chair form of, 64-67 

conformations of, 63-67 



equatorial positions in, 65-67 
heat of combustion of, 68 
nmr spectrum of, 178-179 
oxidation of, 750 
physical properties of, 63 
twist-boat form of, 65-66 

Cyclohexane-l,2-dicarboxylic acid, 377 

Cyelohexanethiol, synthesis of, 522 

Cyclohexanol, in acetal formation, 288 
physical properties of, 628 

Cyclohexanols, from phenols, 635 

Cyclohexanone, bromination of, 303 
cyanohydrin of, 284 
from pimelic acid, 358 
preparation of, 750 

Cyclohexene, with bromine, 128 
with dichloromethylene, 224 
electrophilic additions to, 89 
heat of combustion of, 129 
with iodomethylzinc iodide, 224 

Cyelohexyl bromide, with magnesium, 
522 

Cyclohexylamine, base dissociation 
constant of, 423 
infrared spectrum of, 424 
physical properties of, 423 

Cyclohexylcarbinyl chloride, 190 

Cyclononane, heat of combustion of, 
68 
physical properties of, 63 

Cyclooctane, 63 

heat of combustion of, 68, 148 
physical properties of, 63 

Cyclooctanone, synthesis 1 of, 772 

Cyclooctatetraene, additions to, 724 
from ethyne, 721 
heat of combustion of, 148 
structure of, 578 
valence tautomers of, 723-724 

1,3,5-Cyclooctatriene, valence tauto- 
mers of, 723 

Cyclopentadecane, heat of combustion 
of, 68 

Cyclopentadiene, polymerization of, 
737 

Cyclopentadienyl radical, cation, 
anion, relative stabilities of, 145 
resonance structures of, 144 

7T-Cyclopentadienyldiethenerhodium, 
235 

Cyclopentane, heat of combustion of, 
68 



general index 816 



physical properties of, 63 
c«-l,2-Cyclopentanediol, from cyclo- 

pentene, 99 
?ra«.y-l,2-Cyclopentanediol, from cy- 

clopentene, 99 
Cyclopentanone, from adipic acid, 358 
cyanohydrirt of, 284 
with phosphorus pen-tachloride, 

290 
with sulfur tetrafiuoride, 290 
Cyclopentene, oxidation of, 99 
Cyclopentyl chloride (see Chlorocyclo- 

pentane) 
Cyclopropane, heat of combustion of, 
68 
physical properties of, 63 
Cyclopropanecarbonitrile, reduction 

of, 280 
Cyclopropanecarboxaldehyde, prep- 
aration of, 280 
Cyclopropanecarboxylic acid, forma- 
tion of, 306 
Cyclopropyl methyl ketone, bromina- 

tion of, 306 
Cyclopropyne, 117 
Cysteine, 457, 522 

physical properties of, 461 
Cystine, 457 

from cysteine, 523 
physical properties of, 461 
in wool, 757 
Cytochromes, 508, 641 
Cytosine, in DNA, 480 
Cytosine deoxyribonucleotide, 481 
Cytosine deoxyriboside, 481 



2,4-D, ecological effect, 598 
Dacron, preparation of, 749 
DDE (see DDT) 
DDT, 4 

DDE from, 597 
ecological effects of, 596-597 
DDT dehydrochlorinase, in conversion 

of DDT to DDE, 602 
Decaborane, 537 
Decalin, 551 

conformations of, 785-786 
Decane, isomers of, 48 

physical properties of, 55 
Decarboxylation, of carboxylate rad- 
icals, 342 



of carboxylic acids, 341-342, 354 

Dehydration, of alcohols, 256-258 

Dehydrobenzene (see Benzyne) 

Derealization, in benzene, 549 

of electrons in conjugated systems, 

127-147 
in excited states, 556 

Derealization energy (see also Stabi- 
lization energy and Resonance 
energy), 132 

Delphinidin chloride, 687-688 

Delrin, 288 

Denaturation, of proteins, 469 

2-Deoxyribofuranose, in DNA, 479 

Deoxyribonucleic acid (DNA), in 
chromosomes, 477 

2-deoxy-2-ribose in, 400 
double-stranded helix of, 479 
equivalence of base pairs in, 483 
in genetic control, 483^486 
hydrolysis of, 482 
molecular weights of, 479 
nucleotides in, 482 
replication of, 483-486 
structure of, 477^183 
thermal dissociation of, 479 
Watson-Crick model of, 483 

Deoxyribonucleosides, as glycosides, 
407 

2-Deoxy- D-ribose, structure of, 400- 
401 

Desoxycholic acid, 784 
degradation of, 786 

Detergents, phosphates in, 528 
from sulfonic acids, 526 

Deuteration of arenes, 567 

Dewar, J., 128, 708 

Dewar benzene, 708 

Dextrose, 410 

Diacetone alcohol, dehydration of, 309 
formation of, 307-308 

Diamines, tetrazotization of, 618 

4,4'-Diaminobiphenyl (see Benzidine) 

Diastase, 412 

Diastereomers, 375-379 
definition of, 378 
physical properties of, 378 

Diatomic molecules, potential energy 
diagram, 697 

Diazirine, preparation and properties 
of, 444 

Diazo compounds, 443-444 



general index 817 



Diazo coupling, 619-620 

of heterocyclic compounds, 674- 

677 
of tropolones, 642 
Diazoalkanes, as 1,3-dipoles, 721-723, 

731 
Diazomethane, 422 
esters from, 444 
as methylating agent, 444 
methylene from, 224 
from N-nitroso-N-methyl amide, 

444 
with phenols, 631 
physical properties of, 444 
properties of, 443 
Diazonium salts, 617-620 
coupling of, 619-620 
covalent forms of, 618 
decomposition of, 591 
with hypophosphorous acid, 608 
radicals from, 618 
reduction of, 619 
replacement reactions, 618-619 
in Sn reactions, 194 
stability of, 618 
Diazotization, of/>-aminophenols, 618 

of arylamines, 591-592, 616-617 
Diborane, with alkenes, 96-97, 539-540 

structure of, 92, 537-538 
2,3-Dibromobutane, 377 
trans-1 ,2-Dibromocyclohexane, from 

cyclohexene, 88, 128 
1,2-Dibromoethane, from ethene, 37, 
90 
in formation of organomagnesium 

compounds, 592 
in gasoline, 58, 234 
naming of, 37 
1,1-Dibromoethene, 39 
1 ,2-Dibromoethene, cis and trans iso- 
mers, 38 
from ethyne, 38 
physical properties of, 38 
2,5-Dibromo-3-hexene, from 2,4-hexa- 

diene, 127 
4,5-Dibromo-2-hexene, from 2,4-hexa- 

diene, 127 
5,6-Dibromo-l-hexene, from 1,5-hexa- 

diene, 127 
2,3-Dibromo-4-methylhexane, from 4- 

methyl-2-hexene, 86 
frfl«j-l,2-Dibromo-l-propene, from 



propyne, 121 
Di-?-butyl ether, stability of, 257 
Di-n-butyl phthalate, as plasticizer, 748 
c«-.s)w-Di-?-butylethylene, repulsive 

interactions in, 85 
1,3-Dicarbonyl compounds, 314 
Dicarboxylic acids, 356-359 

acidic properties of, 356-358 
and Blanc rule, 786 
in ketone synthesis, 772 
thermal behavior of, 358 
Dichloroacetic acid, acid dissociation 
constant of, 331 
physical properties of, 331 
w-Dichlorobenzene, preparation of, 

592 
1 , 1 -Dichloro-2,2-bis(p-chlorophenyl) 

ethene (see DDT, from DDE) 
Dichlorocarbene, from chloroform, 
223 
electrophic nature of, 223 
Dichlorodiphenyltrichloroethane (see 

DDT) 
Dichloromethane, with hydroxide ion, 
223 
from methane chlorination, 28, 60, 
222 
Dichloromethane, physical properties 

of, 222, 535 
Dichloromethylene (see Dichlorocar- 
bene) 
1 ,5-Dichloro-2,6-naphthoquinone, 638 
Dichlorosilane, physical properties of, 

535 
Dicyclohexylcarbodiimide, in oxida- 
tion of alcohols, 526 
in peptide synthesis, 472-473 
Dieckmann reaction, 716 
Dielectric constant, of solvents in S N 

reactions, 204 
Diels-Alder reaction, 719-721 
of cyclooctatetraene, 724 
of cyclopentadiene, 737 
Dienes, conformations of, 719 
Dienophile, definition of, 719 
Diesel oil, composition of, 58 
Diethyl ether, from ethanol, 256 

from ethyl hydrogen sulfate, 256 
with methylsodium, 264 
peroxides from, 265 
physical properties of, 263 
as a solvent, 264 



general index 818 



as solvent for organomagnesium 

compounds, 230 
uses of, 263 
Diethyl maleate, with 1,3-butadiene, 

720 
Diethyl malonate, alkylation of, 354 
Diethyl phenylphosphonate, 530 
Diethyl sulfide, 521 

Diethylamine, base dissociation con- 
stant of, 423 
with 1 -chloro-2,4-dinitrobenzene, 

593 
nmr spectrum of, 425 
physical properties of, 423-424 
Diethyldiketopiperazine, from ethyl 

2-aminobutanoate, 468 
Diethylene glycol, 264 

from ethylene oxide, 266 
physical properties of, 263 
Diethyloxonium bromide, 264 
Difluorodichloromethane, formation 
of, 224 
reactivity of, 224-225 
utility of, 224 
o,o'-Difiuorodiphenic acid, optical iso- 
mers of, 381 
1,1-Difluoroethane, from ethyne, 114 
Digitogenin, 791 
Digitoxigenin, 791 
Diglyme, 264 

physical properties of, 264 
9,10-Dihydroanthracene, 551 
Dihydropentaborane, 537 
1,2-Dihydroxyanthraquinone, 760 
m-Dihydroxybenzene (see Resorcinol) 
o-Dihydroxybenzene (see Catechol) 
p-Dihydroxybenzene (see Hydroquin- 

one) 
Dihydroxymethane (see Formalde- 
hyde) 
2,4-Dihydroxypyrimidine (see Uracil) 
Diisobutylene, from 2-methylpropene, 

102 
Diisopropyl ether, peroxides from, 265 

as a solvent, 264 
Diisopropyl fluorophosphate, 506 
Diketene, structure of, 313, 321 
1,1-Dimethoxyethane, from acetalde- 
hyde, 284 
nmr spectrum of, 175 
Dimethyl acetal (see 1,1-dimethoxy- 
ethane) 



Dimethyl ether, bond angles in, 10 
physical properties of, 12 
solubility in water, 1 2 
Dimethyl sulfate, 521 
properties of, 528 
Dimethyl sulfoxide, acid strength of, 
526 
dielectric constant of, 205, 525 
double bonds in, 519 
oxidizing properties of, 526 
physical properties of, 525 
shape of, 519 
solvent properties of, 525 
N,N-Dimethylacetamide, physical 

properties of, 435 
Dimethylallylcarbinol, 191 
Dimethylamine, physical properties of, 

12 
/;-Dimethylaminoazobenzene, forma- 
tion of, 620 

carcinogenic properties of, 620 
Dimethylaminoborane, 539 
N,N-Dimethylaniline, basicity of, 614 
diazo coupling of, 620 
uv spectrum of, 614 
Dimethylbenzenes (see Xylenes) 
2,2-Dimethylbutane, 49, 55 
2,3-Dimethylbutane, 49, 55 
2,3-Dimethyl-2,3-butanediol (see Pin- 

acol) 
3,3-Dimethyl-2-butanol, dehydration 

of, 258 
3,3-Dimethyl-2-butanone (see Pina- 

colone) 
l,3-Dimethyl-5-f-butyl-2,4,6-trinitro- 

benzene, in perfumes, 612 
Dimethylchloroborane, 539 
1 ,4-Dimethylcyclohexane, naming of, 

62 
1,3-Dimethylcyclohexene, 82 
1 ,2-Dimethylcyclopropane, cis-trans 

isomers of, 69 
4,4-Dimethyl-l,2-dibromo-l-pentene, 

from 4,4-dimethyl-l-pentyne, 113 
N,N-Dimethylformamide, dielectric 
constant of, 205 
infrared spectrum of, 436 
solvent properties of, 437 
Dimethylglycolic acid, from acetone, 

282 
1,2-Dimethylhydrazine, 422 
N,N-Dimethylhydroxylamine, 422 



general index 819 



Dimethylmercury, formation of, 228- 
229 

2,2-Dimethylpentane, from 4,4-di- 
methyl-1-pentyne, 113 

2,3-Dimethylpentane, naming of, 51 

4,4-Dimethyl-l-pentene, from 4,4-di- 
methyl-1-pentyne, 113 

4,4-Dimethyl-l-pentyne, bromine ad- 
dition to, 113 
hydrogenation of, 113 

2,2-Dimethylpropane, 49 

physical properties of, 11, 535 

3,5-Dimethylpyridine, mass spectrum 
of, 773-775 

Dimethylsilanediol, physical proper- 
ties, 535 

Dimsyl ion, 526 

2,4-Dinitroaniline, reduction of, 609- 
610 

p-Dinitrobenzene, synthesis of, 606- 
607 

2,4-Dinitrobenzenesulfenyl chloride, 
521 

2,4-Dinitrochlorobenzene, formation 
of, 632 

o,o'-Dinitrodiphenic acid, optical iso- 
mers of, 380 

2,4-Dinitrofluorobenzene, with pep- 
tides, 593 
in N-terminal amino acid analysis, 
470 

2,4-Dinitro-l-naphthol (see Martius 
Yellow) 

2,4-Dinitrophenol, with phosphorus 
pentachloride, 632 

2,4-Dinitrotoluene, nitration of, 562 
reduction of, 609-610 

Diols (see Glycols) 

Diosgenin, cortisone from, 788 

1,4-Dioxane, from ethylene oxide, 
266 
peroxides from, 265 
physical properties of, 263 
solubility in water, 264 
as a solvent, 264 

Diphenyl disulfide, 521 

Diphenyl ether, 190 

Diphenyl sulfone, 521 

Diphenylamine, basicity of, 614 
uv spectrum of, 614 

l,10-Diphenyl-l,3,5,7,9-decapen- 
taene, light absorption of, 699 



1,2-Diphenylethene, light absorption 

of, 699 
Diphenylmethane, 550 
Diphenylmethyl cation, resonance 

stabilization of, 653 
Diphosphopyridine nucleotide 
(DPN ffi ) (see Nicotinamide-adenine 
dinucleotide NAD®) 
Diphosphoric acid, derivatives of, 528 
1,3-Dipolar additions, 721-723 
1,3-Dipolar reagents, 722 
Dipole moments, demonstration of, 7 

1,2-dibromoethene, cis and trans, 
39 

and molecular shape, 7 

of water, 7 
Disaccharides, 400, 407, 408-410 
Disilane, physical properties of, 535 
Dispersion forces (see van der Waals 

forces) 
Disproportionation, in alkene polym- 
erization, 102 
Disulfides, preparation of, 523 

in proteins, 457-458 

reduction of, 747 
Diterpenes, 779-781 
Dithioacetals, from thiols, 523 
Dithioacetic acid, 520 
Dithioketals, from thiols, 523 
DMSO (see Dimethyl sulfoxide) 
DNA (see Deoxyribonucleic acid) 
DNA-polymerase, 484 
Dodecane, physical properties of, 55 
Double bonds, with 3d orbitals, 519 
Double bonds, with 3p orbitals, 519 
Double bonds, in silicon compounds, 

534, 535 
Drugs, 4 

antibiotic, 527 

barbiturates, 685 

examples of, 658-659 
Durene, physical properties, 553 
Dyes, azo compounds, 620 

for cotton, 759, 760 

Crystal Violet, 702 

for Dacron, 759 

disperse, 759 

fading of, 160 

Martius Yellow, 700 

mordant, 759-760 

for polymers, 758-762 

for rayon and cotton, 762 



general index 820 



sulfonic acids in, 527 
vat, 760-762 

for wool and silk, 758-759 
Dynamite, 262 



E2 reaction (see Elimination reactions) 
Eicosane, isomers of, 48 

physical properties of, 55 
Einstein units, definition of, 703 
Elastomers, configuration, 742 
Electromagnetic radiation, absorption 
of, 159-161 
electric and magnetic forces in, 
369 
Electromagnetic spectrum, regions of, 

695 
Electron diffraction, of boron hy- 
drides, 538 
Electron paramagnetic resonance spec- 
troscopy, 659-661 
Electron paramagnetic resonance spec- 
trum, of cycloheptatrienyl, 668 
Electron repulsion, and acidity of 
alkynes, 116 
between bonding electron pairs, 9 
between nonbonding electron 

pairs, 9 
and electrophilic addition, 87 
and molecular shape, 8 
in singlet and triplet states, 698 
and tetrahedral methane, 130 
Electron spin, 130, 131 
Electron spin resonance spectroscopy 
(see Electron paramagnetic reso- 
nance spectroscopy) 
Electron transport chain, 641 
Electronegativity, and acid strength, 
116,338 
of carbon, 6 
of elements, 6 
of fluorine, 6 
of hydrogen, 6 

and hydrogen bonding, 11, 423 
of nitrogen, 6 
of oxygen, 6 
and polarity, 6 
scale of, 6 
Electronic absorption spectra, of 
arenes, 554-557 
of aldehydes, 334 
of amides, 437 



benzenoid bands, 557 

of carboxylic acids, 334 

of ketones, 334 

of polyenes, 699 

and rotatory dispersion, 391 

of a,/3-unsaturated carbonyl com- 
pounds, 311 
Electronic absorption spectroscopy, 

165-168 
Electron absorption spectrum, of an- 
iline, 556 

of anthracene, 557 

of benzaldehyde, 556 

of benzene, 556 

of biphenyl, 556 

of iodobenzene, 556 

of methyl ethyl ketone, 277-278 

of methyl vinyl ketone, 277-278 

of naphthacene, 557 

of naphthalene, 557 

of />-nitrophenol, 700 

of pentacene, 557 

of phenol, 556 

of stilbene, 556 

of styrene, 556 
Electronic configuration, of oxygen 

and sulfur, 518 
Electronic states, singlet, 697-699 

triplet, 697-699 
Electronic transitions, «->w*, 277, 
391, 704 

77->77*, 278, 311, 554 
Electrophile, definition of, 87 
Electrophilic addition, 87-92 

to alkenes, mechanism of, 87-92 
Electrophilic aromatic substitution, 
559-577 

acylation, 565-566 

alkylation, 564-565 

by aryl cations, 619 

catalysts in, 561, 563, 564 

chloromethylation, 652 

deuteration, 567 

diazo coupling, 619-620 

of disubstituted benzenes, 573- 
574 

examples of, 560 

formylation, 642 

of furan, 674-678 

halogenation, 562-563 

kinetic control in, 569 

mechanism of, 559-561 



general index 821 



nitration, 561-562 
orientation in, 567-574 
of phenoxide ion, 633 
of polynuclear aromatic hydro- 
carbons, 574-577 
of pyridine, 674-678 
of pyrrole, 674-678 
reactivity effects in, 570-573 
substituting agents in, 561 
sulfonation, 566 
of thiophene, 674-678 
of tropolones, 642 
Elemental analysis, 156-157 

from mass spectrometry, 159 
Elimination reactions, 205-208 
of alkyl hydrogen sulfate, 256 
1,1 or alpha, 223 
of aryl halides, 595-596 
the El reaction, 207-208 
the E2 reaction, 206-207 
in ether synthesis, 252 
Emulsin, 410 

Enamines, rearrangement of, 427 
Enantiomers, definition of, 373 
physical properties of, 374 
separation of, 384 
End-capping, of polymers, 288 
Endocyclic, definition of, 82 
Enolate anions, alkylation of, 310 
formation of, 304-306 
reactions of, 306-310 
in S N reactions, 310 
Enolization, acid-catalyzed, 304-305, 
501 
base-catalyzed, 304-306 
Enols, hydrogen bonding in, 315 

oxidation of, 293 
Enovid, 789 

Enthalpy, definition of, 23 
Entropy of activation, 496 

definition of, 199 
Enzymes, active sites of, 504 

alcohol :NAD® oxidoreductase, 

508 
and asymmetric synthesis, 387 
cytochromes, 476, 508 
epr spectra of, 661 
flavins, 508 

in hydrolysis of DNA, 482 
in hydrolysis of saccharides, 410 
hydrolytic, 504-506 
in mitochondria, 509 



oxidative, 506-509 
Enzymic processes, 495-513 
Epimers, 403 
Epinephrine, 659 
Equatorial positions and substituents 

(see Cyclohexane) 
Equilenin, 790 
Equilibrium constants, 26 
Ergosterol, 790 

vitamin D from, 789 
Ergot alkaloids, 683 
Erythrose, oxidation of, 379 

projection formulas of, 378 
Escherichia coli, DNA from, 483 
Esr or epr (see Electron paramagnetic 

resonance spectroscopy) 
Essential oils, 778-782 
Esterification (see Carboxylic esters, 

formation of) 
Esters (see Carboxylic esters) 
Estradiol, 790 
Estrone, 790 
Ethane, bond lengths in, 35 

heat of combustion of, 24, 113, 
510 

molecular shape of, 20 

in natural gas, 57 

physical properties of, 11, 55, 112, 
246, 535 

rotational conformations of, 21 

structure of, 19 
1,2-Ethanediol (see Ethylene glycol) 
Ethanesulfinic acid, 521 
Ethanethiol, 520 

acid dissociation constant, 522 

formation of, 522 

physical properties of, 522 
Ethanol (see Ethyl alcohol) 
Ethan- l-ol-2-thiol, 524 
Ethene, additions to, 34 

bent bonds in, 20, 34 

bond lengths in, 35 

bonding in, 34 

with chlorine, 223 

cycloaddition of, 729 

from ethyl alcohol, 256-257 

from ethyl chloride, 206 

geometry of, 35 

halogen addition to, 37 

heat of combustion of, 68, 113 

hydration of, 88, 250, 275 

hydroboration of, 96 



general index 822 



hydrogenation of, 35 

model of, 20, 34 

molecular orbitals of, 727-729 

molecular shape of, 20 

oxidation of, 261 

physical properties of, 112 

polymer from, 745 

polymerization of, 100-103 

structure of, 19 
Ethers, 262-266 

acid cleavage of, 264 

from alcohols, 252, 256-257, 263 

from alkyl halides, 263 

basic cleavage of, 264 

basic properties of, 264 

cyclic, 265-266 

as derivatives of water, 9 

naming of, 190 

peroxides from, 265 

physical properties of, 263 

preparation of, 263 

reactions of, 264-265 

as solvents, 264 
Ethyl acetate, in Claisen condensation, 

352-353 
Ethyl acetoacetate (see also Claisen 
condensation) 

acid dissociation constant of, 351 

alkylation of, 354 

nmr spectrum of, 350-352 

synthesis of, 352-353 

tautomers of, 351-352 
Ethyl alcohol, absolute, 250 

in acetal formation, 287 

with acetic acid, 254 

acid dissociation constant of, 251, 
315, 522 

acidity of, 337 

dehydration of, 256-257 

from ethene, 88, 250 

in ether formation, 256-257 

from ethyl chloride, 206 

by fermentation, 245, 250 

from formaldehyde, 231 

heat of combustion of, 510 

infrared spectrum of, 247-248 

manufacture of, 250 

NAD® oxidation, 506-507 

nmr spectrum of, 169 

physical propert ies of, 522 

salt of, 251 

with sodium amide, 251 



with sodium hydride, 251 
toxicity of, 245 
Ethyl 2-aminobutanoate, thermal de- 
composition of, 468 
Ethyl 4-aminobutanoate, thermal de- 
composition of, 468 
Ethyl benzoate, alkaline hydrolysis of, 
661 
nitration of, 568 
Ethyl benzoylacetate, synthesis of, 353 
Ethyl bromide, physical properties of, 

217 
Ethyl butanoate, mass spectrum of, 

778 
Ethyl chloride, with sodium hydroxide, 

206 
Ethyl 2-chloro-3-butenoate, 191 
Ethyl ether (see Diethyl ether) 
Ethyl formylphenylacetate, synthesis 

of, 353 
Ethyl hydrogen sulfate, ethene from, 

256 
Ethyl iodide, nmr spectrum of, 171 
Ethyl methyl sulfoxide, 521 
2-(Ethyl oxalyl)-cyclohexanone, syn- 
thesis of, 353 
Ethyl phenylacetate, 191 
Ethyl thioacetate, formation of, 523 
Ethylamine, base dissociation constant 
of, 423 
physical properties of, 423 
Ethylbenzene, 550 
formation of, 564 
physical properties of, 553 
Ethylene (see Ethene) 
Ethylene chlorohydrin, from ethylene 

oxide, 266 
Ethylene dibromide (see 1,2-Dibromo- 

ethane) 
Ethylene glycol, from ethylene oxide, 
261, 266 
physical properties of, 261 
polymer from, 747 
uses of, 261 
Ethylene oxide, with 1-bromohexane, 
265 
commercial importance of, 266 
from ethane, 261 

with organomagnesium com- 
pounds, 234, 265 
physical properties of, 263 
polymerization of, 755 



general index 823 



reactivity of, 265 

strain in, 265 
Ethylenediamine, base dissociation 
constant of, 423 

physical properties of, 423 
Ethylmagnesium bromide, formation 

of, 228 
Ethylmercuric chloride, fungicidal 

properties, 234 
l-Ethyl-3-methylcyclopentane, 62 
4-Ethyl-3-methylheptane, 52 
4-Ethyl-3-methyl-;ra«^-3-heptene, 84 
3-Ethylpentane, physical properties of, 

424 
Ethylphenylcarbinylamine, 190 
Ethylpyridines, mass spectra of, 773- 

775 
Ethyne, acid ionization constant of, 
122 

additions to, 34, 38 

bond lengths in, 35 

from calcium carbide, 112 

with chlorine, 223 

dimerization of, 114, 115, 119 

heat of combustion of, 24, 113 

hydration of, 114 

hydrogenation of, 35 

model of, 20 

molecular shape of, 20 

from petroleum gases, 112 

physical properties of, 112 

stability of, 112 

structure of, 19 

for welding, 112 
Euclid, 3 

Eugenol, oxidation of, 658 
Evolution, chemical, 486-487 
Excited states, of ferric phenoxide, 631 
Exocyclic, definition of, 82 
Explosives, 4, 611-612 

from cellulose, 410 

dynamite, 262 

nitro compounds, 442 

nitrocellulose, 262 

nitroglycerin, 262 

peroxides, 265 



Farnesol, 780 

Fats, glycerol from, 262 

hydrolysis of, 330 
Fatty acids, 330 



Fehling's solution, 293 , 404 

Ferric chloride, complexes with tropo- 
lones, 642 
with phenols, 631 

Ferrocene, bonding in, 235 
from cyclopentadiene, 235 
physical properties of, 235 

Fibers, from cellulose, 410 
dyeing of, 758-762 

Fibrogenin, properties and function of, 
478 

Fischer, E., 402-403 

Fischer, H., 681 

Flavins, 508, 641 

Flavones, 687 

Flavorings, examples of, 658-659 

Flax, cellulose in, 410 

Fluorescence, 696-699 

Fluorine, heat of reaction with meth- 
ane, 75 
reaction with alkanes, 60 

Fluorine oxide, dipole moment of, 7 
shape of, 7 

Fluorobenzene, formation of, 619 
physical properties of, 590 

l-Fluoro-4-bromonaphthalene, forma- 
tion of, 591 

Fluorocarbons, 225-226 

physical properties of, 225-226 

Fluorochloromethane, formation of, 
224 

2-Fluoroethanol, toxicity of, 226 

Fluorosulfonic acid, super acid, 13 

Fluxional systems, 723-725 

Formal bonds, 136 

Formaldehyde, with acetaldehyde, 325 
with acetone, 308-309 
in aldol addition, 308-309 
carbonyl bond strength of, 308 
in chloromethylation, 652 
from dichloromethane, 223 
hydration of, 261, 275 
with lithium aluminum hydride, 

291 
with organomagnesium com- 
pounds, 231-232 
physical properties of, 277 
polymerization of, 287-288 
polymers from, 747, 751-752 
with sodium borohydride, 291 

Formaldoxime, 289 

Formamide, physical properties of, 435 



general index 824 



Formamidine, in adenine synthesis, 487 
Formic acid, 190 

acid dissociation constant of, 331 

physical properties of, 331 
Formyl chloride, 340 
Formylation, of arenes, 652 

of heterocyclic compounds, 674- 
677 
Franck-Condon principle, 696-697 
Free energy of activation, 496 
Free energy of reaction, 26-28 
Friedel-Crafts acylation (see also Acyl- 
ation), 565-566 

of benzene, 651 
Friedel-Crafts alkylation (see also 
Alkylation), 564-565 

of benzene, 651 
Fructose, phenylosazone from, 403 

structure of, 401 
Fumaric acid, 86 

acid dissociation constant of, 357 

physical properties of, 357 
Fungicides, natural, 641 
Fukui, K., 726 
Furan, aromatic character of, 672-673 

chemical properties of, 673-679 

physical properties of, 671 
Furanoses, 406 
Fusel oil, 245 



D-Galactose, structure of, 401 

Gallstones, 782 

Gas constant, 27 

Gasoline, composition of, 57-58 

Gelatin, 757 

Gentiobiose, in amygdalin, 657 

Geometrical isomerism, 39-41 

Geraniol, 780 

Geranyl pyrophosphate, in biogenesis, 

792 
D-Glucaric acid, from glucose, 404 
Gluconic acid, from glucose, 404 
D-Glucosamine, from chitin, 412 
Glucose, 402-404 

anomers of, 404^405 

configuration of, 383 

conformations of, 405 

heat of combustion of, 24, 510 

mutarotation of, 406 

phenylosazone from, 403 

photosynthesis of, 387 



properties of, 403^-04 

stereoisomers of, 383 

structure of, 401-402 
Glucovanillin, 658 
Glutamic acid, isoelectric point, 463 

physical properties of, 461 
Glutamine, physical properties of, 461 
Glutaric acid, acid dissociation con- 
stant of, 357 

anhydride from, 358 

physical properties of, 357 
Glutaric anhydride, formation of, 358 
Glyceraldehyde, absolute configuration 
of, 382 

from acrolein, 286 
Glyceric acid, in photosynthesis, 399 
Glycerides, hydrolysis of, 330 
Glycerol, esters of, 330 

from fats, 262 

physical properties of, 261 

polymer from, 750 

from propene, 220, 262 

uses of, 261 
Glyceryl tristearate, heat of combus- 
tion of, 510 
Glycine, acid-base properties of, 462 

N-acyl derivatives, 784 

heat of combustion of, 510 

isoelectric point, 463 

physical properties of, 459 
Glycogens, 412 

Glycolic acid, from glyoxal, 313, 321 
1,2-Glycols, from alkenes, 261 

definition of, 260-261 

oxidation of, 279, 336 

rearrangements of, 280-281 
1,1-Glycols, dehydration of, 261 
Glycosides, 407-408 
N-Glycosides, in DNA, 480 
Glycylalanine, 469 
Glyoxal, 313, 321 
Glyptal resin, 750 
Gomberg, M., 654 
Grignard, V., 229 
Grignard reagents (see Organomag- 

nesium compounds) 
Guanidine, base strength of, 444 
Guanine, in DNA, 480 



2-Halo acids (see 2-Halocarboxylic 
acids) 



general index 825 



Haloalkanes, naming of, 53 
Haloalkynes, in S N reactions, 220 
Haloamines, formation of, 432 

properties of, 432 
2-Halocarboxylic acids, preparation of, 
343-344 

substitution of, 343-344 
Haloform reaction, 292, 305-306 

with />-tolyl methyl ketone, 650 
Halogenation {see also Electrophilic 
aromatic substitution and Radical, 
halogenation) 

of arenes, 562-563 

of carbonyl compounds, 303-306 

of carboxylic acids, 342, 343 

of heterocyclic compounds, 674- 
677 
Halogens, addition to alkenes, 87-91 

additions to multiple bonds, 37 

charge-transfer complexes of, 562- 
563 

reactivity in aromatic substitution, 
563 
Halonium ions, 92 
Halosilanes, examples of, 532 
Hammett equation, 663 
Heat of activation, 496 

definition of, 199 
Heats of combustion, of various 

substances, 510 
a-Helix, hydrogen bonding in, 475 
Hemiacetals, of carbohydrates, 404 

cyclic, 287 

formation of, 283-287, 497, 499 

hydrolysis of, 497 

from hydroxyaldehydes, 716 
Hemin, structure of, 680 
Hemlock, coniine in, 684 
Hemoglobin, 680 

properties and function of, 478 

quaternary structure of, 475 
Hemp, cellulose in, 410 
Heptane, isomers of, 48 

octane rating of, 58 

physical properties of, 53-55, 63 
Herbicides, 4 

Heterocycle, definition of, 427 
Heterocyclic compounds {see also 
individual types such as Pyrrole, 
Furan, etc.), 671-689 

by 1,3-dipolar addition, 722 

naming of, 671 



natural products, 680-689 

nucleophilic substitution reactions 
of, 678-679 

saturated types, 671 

unsaturated types, 672 
Heterolytic bond breaking, definition 

of, 88 
Hexaborane, 537 
1,5-Hexadiene, with bromine, 127 

Cope rearrangement of, 724 
2,4-Hexadiene, with bromine, 127 

conjugate addition of, 720 
l,3-Hexadien-5-yne, 111 
Hexafluoroacetone hydrate, 261 
Hexafiuoropropene, polymers from, 

745 
Hexamethylenediamine, nylon from, 
747, 750 

preparation of, 751 
Hexamethylethane, 51 
Hexamethylphosphoramide, dielectric 

constant of, 205 
Hexane, isomers of, 48-49, 55 

physical properties of, 53-55, 63, 
112,424 
Hexanedioic acid (see Adipic acid) 
Hexaphenylethane, formation and dis- 
sociation, 654 
1 -Hexene, physical properties of, 1 1 2 
cw-3-Hexene, heat of combustion of, 85 

from 3-hexyne, 115 

physical properties of, 85 
trans-3-Hexem, heat of combustion of 
85 

physical properties of, 85 
l-Hexene-3,5-diyne, 111 
1-Hexyne, physical properties of, 112 
3-Hexyne, hydroboration of, 115 
Hinokitiol, 641 
Histidine, in ester hydrolysis, 505 

in hydrolytic enzymes, 500 

physical properties of, 461 
HMP {see Hexamethylphosphoramide) 
Hodgkin, D. C, 476 
Hoffmann, R., 726 
Hofmann reaction, 430 
Homology, concept of, 53-56 
Homolytic bond breaking, definition 

of, 88 
Hormones, 4 

adrenal cortex, 791 

sex, 790-791 



general index 826 



Hiickel, E., 141 

Hiickel's (An + 2) rule, 145, 578 
Hund's rule, 142, 698 
Hunsdiecker reaction, 342 
Hydrazine, 289 

in reduction of carbonyl com- 
pounds, 292 
Hydrazines, formation and properties 

of, 442^143 
Hydrazobenzene, from nitrobenzene, 
611 

rearrangement of, 61 1 
Hydrazoic acid, 422 
Hydrazones, formation of, 289 
Hydroboration of alkenes, 96-97, 
539-540 

of propene, 97 
Hydrocarbons (see also individual types 
such as Alkanes, Alkenes, etc.) 

C-l and C-2 compounds, 19-42 

chemical reactions of, 23 

definition of, 9, 19 

nomenclature of, 47-52 

substitution of, 26 

solubility in water, 12 
Hydrogen, addition to 4-methyl-2- 
hexene, 86 

addition to muliple bonds, 35, 
113, 115 
Hydrogen bonding, in adenine-thy- 
mine, 482 

in alcohols, 246-249, 334 

in amides, 435 

in amines, 423^425 

in carboxylic acids, 330, 334 

and decarboxylation, 341 

in dicarboxylic acids, 356-357 

in DNA, 479 

in enols, 315 

in guanine-cytosine, 482 

intramolecular, 634-635 

in N-methylamine, 12 

N-H---N, 12 

O-H-O, 11-12 

in peptides, 475 

in phenols, 628 

and physical properties, 10 

in polyamides, 739 

in pyrrole, 671 

in silanols, 535 

in silk, 757 

and spectroscopic properties of 



alcohols, 247-249 
strength of, 1 1 
in thiols, 522 
in tropolone, 642 
Hydrogen bromide, addition to 3- 
methyl-1-butene, 119 
addition to 3-methyl-l-butyne, 

119 
electrophilic addition to propene, 

94 
with methanol, 251 
radical addition to propene, 
94 
Hydrogen cyanide, in adenine synthe- 
sis, 486^-87 
with aldehydes and ketones, 282- 

284 
conjugate addition of, 31 1 
Hydrogen fluoride, acid dissociation 
constant of, 13 
addition to ethyne, 1 14 
base ionization constant of, 13 
physical properties of, 10 
Hydrogen halides, in alkyl halide 
formation, 255 
with carbonyl groups, 288 
conjugate addition of, 311 
Hydrogen molecule ion, 92 
Hydrogen peroxide, in alkene oxida- 
tion, 261 
singlet oxygen from, 706 
Hydrogen sulfide, acid dissociation 

constant of, 522 
Hydrogenation, of acyl halides, 279 
in analysis, 36 
of carbon monoxide, 251 
catalytic, 35 
with diborane, 96-97 
of diisobutylene, 102 
of ethene, heat of, 35 
of ethyne, heat of, 35 
mechanism of catalytic, 36 
Hydrolysis, of acetals, 287 

of acid derivatives, 336, 344, 439 

of alkyl halides, 249 

of amides, 336, 439 

of ATP, 509, 529 

of carboxylate salts, 329 

of esters, 336 

of nitriles, 336, 439 

of organosilicon compounds, 534 

of peptides, 469 



general index 827 



Hydroperoxides, from organomagne- 

sium compounds, 231 
Hydroquinone, acid dissociation con- 
stant of, 629 
structure of, 635 
uv spectrum of, 629 
Hydroxyaldehydes, hemiacetals from, 

716 
Hydroxycarboxylic acids, lactones 

from, 715 
jj-Hydroxybenzaldehyde, acid disso- 
ciation constant of, 629 
uv spectrum of, 629 
Hydroxybenzaldehydes, physical prop- 
erties of, 634-635 
o-Hydroxybenzoic acid {see Salicyclic 

acid) 
jS-Hydroxybutyraldehyde {see Acetal- 

dol) 
3-Hydroxyindole, indigo from, 760-761 
Hydroxyl groups {see also Alcohols, 
Carboxylic acids, etc.) 
chemical shift of, 247-249 
hydrogen bonding in, 11, 245-249 
infrared bands of, 247 
and solubility, 12 
Hydroxylamines, 289, 422 

from secondary amines, 434 
tautomers of, 434 
5-Hydroxylysine, physical properties 

of, 459 
2-Hydroxy-5-methyl-3-hexenoic acid, 

190 
2-Hydroxy-l -naphthaldehyde, forma- 
tion of, 653 
1 1 -Hydroxyprogesterone, cortisone 

from, 788 
Hydroxyproline, in ninhydrin test, 465 

physical properties of, 461 
3-Hydroxypropanoic acid, from acry- 
lic acid, 355 
2-Hydroxypyridine, tautomers of, 723 
2-Hydroxypyrimidine, derivatives of 

•in DNA, 480 
a-Hydroxysulfonate salts, from alde- 
hydes, 527 
5-Hydroxytryptamine, in the brain, 

682-683 
Hypochlorous acid, addition to cyclo- 

hexene, 89 
Hypohalous acids, addition to alkenes, 
249 



I effect {see Inductive effects) 
Identification, of organic compounds, 

156-159 
Imidazole, 679 

basicity of, 500 
in ester hydrolysis, 500 
Imides, acid strengths of, 439 
Imines, from enamines, 427 

from nitriles, 349 
Indican, 760 
Indigo, configuration of, 762 

preparation and occurrence, 760- 

762 
properties, 762 
• Indole, natural products related to, 
682-683 
structure of, 679 
Indole alkaloids, 682-683 
quebrachamine, 772 
Indoxyl {see 3-Hydroxyindole) 
Inductive effects, and acid strengths, 
337-338 
of carboxyl groups, 356-357 
and decarboxylation, 341 
in electrophilic aromatic substitu- 
tion, 570-573 
of ester groups, 350-351 
Infrared spectra, of alcohols, 247- 
248 
of alkyl halides, 217-218 
of amides, 435 
of amines, 424-425 
of arenes, 554-555 
of aryl halides, 590 
of carboxylic acids, 334-335 
Infrared spectroscopy, 161-165 

absorption frequencies in, 166-167 
of C-H bonds, 164-165 
fingerprint region, 162, 164 
of multiple bonds, 164-165 
Infrared spectrum, of acetaldehyde, 
334-335 
of acetanilide, 436 
of acetic acid, 334-335 
of 1-butene, 163 

of carbon tetrachloride, 217-218 
of chloroform, 217-218 
of cyclohexylamine, 424 
of N,N-dimethylformamide, 436 
of ethanol, 247-248, 334-335 
of N-methylaniline, 424 
of octane, 163 



general index 828 



of phenylacetylene, 164 
of propanamide, 436 
of toluene, 554-555 
of w-xylene, 554-555 
of o-xylene, 554-558 
of p-xylene, 554-555 
Inhibitors, in chain reactions, 32 
Initiation, of chain reactions, 31 
Initiators, in radical polymerization, 

754 
Inositol, 418 
Insecticides, 4 

Insulin, amino acids in, 475 
primary structure of, 475 
properties and function of, 478 
quaternary structure of, 475 
X-ray diffraction of, 476 
Intermediates, in electrophilic aromatic 
substitution, 559, 675 
in nucleophilic aromatic substitu- 
tion, 593 
Intermolecular forces, and physical 

properties, 10 
Intersystem crossing, 698 

in benzophenone, 704 
Invert sugar, 410 
Invertase, 410 

Iodide ion, catalysis by, 496 
Iodine monochloride, in aromatic 

substitution, 563 
Iodobenzene, physical properties of, 
590 
uv spectrum of, 556 
Iodoethane (see Ethyl iodide) 
Iodoform, from methyl ketones, 305- 

306 
2-Iodo-3-hydroxybenzaldehyde, Can- 

nizzaro reaction of, 650 
2-Iodo-3-hydroxybenzoic acid, forma- 
tion of, 650 
2-Iodo-3-hydroxybenzyl alcohol, form- 
ation of, 650 
Iodomethane (see Methyl iodide) 
Iodonium ions, in S N reactions, 194 
1-Iodophenanthrene, formation of, 

591 
2-Iodopropanoic acid, formation of, 

343 
Iodotoluenes, formation of, 563 
Ion-exchange resins, in amino acid 

analysis, 466 
Ionic bonding, 5 



Ionic polymerization, 755-756 

IPP (see Isopentenyl pyrophosphate) 

Iso, definition of, 49 

Isobutane (see 2-Methylpropane) 

Isobutyl alcohol, physical properties 

of, 247 
Isobutyl bromide, 53 

equilibration with f-butyl bro- 
mide, 94 
Isobutylene (see 2-Methylpropene) 
Isobutyraldehyde, acetals from, 287 

from 2-methylpropene, 281 
Isobutyric acid (see 2-Methylpropanoic 

acid) 
Isoelectric points, of amino acids, 459- 
461 

definition of, 463 
of proteins, 476, 478 
Isoelectronic, definition of, 8 
Isohexane (see 2-Methylpentane) 
Isoleucine, physical properties of, 459 
Isomerism (see also individual types 
such as Optical, Geometric, etc.) in 
alkenes, 84 
in arenes, 550 
conformational, 38, 84 
in cycloalkanes, 69 
geometrical, 39, 69, 84 
structural, 38, 84 
Isomers, configurational, 38 
conformational, 38 
definition of, 369 
ortho, meta, and para, 550 
structural, 38 
Isooctane (see 2,2,4-Trimethylpentane) 
Isopentenyl pyrophosphate (IPP), in 

biogenesis, 792 
Isoprene, 83 

polymer from, 742-743, 746 
polymerization of, 755 
Isoprene units, in naturally occurring 
quinones, 641 
in terpenes, 779 
Isopropyl alcohol, in acetal formation, 
287 
manufacture of, 251 
oxidation of, 259 
in photoreduction of ketones, 704 
Isopropyl bromide, from propane, 
219 
from 2-propanol, 219 
from propene, 94, 219 



general index 829 



Isopropyl chloride, physical proper- 
ties of, 217 
from propene, 93 
Isopropyl hydrogen chromate, 259 
Isopropyl methyl ether, formation of, 

252 
/j-Isopropylbenzaldehyde, formation 

of, 652 
Isopropylbenzene, 550 
acetone from, 627 
formation of, 564 
formylation, 652 
hydroperoxide, 627-628 
oxidation of, 627 
phenol from, 627 
physical properties of, 553 
Isopropyllithium, triisopropylcarbinol 

from, 234 
Isoquinoline, 679-680 

natural products related to, 684 
Isovaleryl chloride {see 3-Methyl bu- 

tanoyl chloride) 
IUPAC {see also Nomenclature), 19 



Jute, cellulose in, 410 



Kekule,A,3, 127 

Kel-F, from chlorotrifluoroethene, 225 
Kerosene, composition of, 58 
Ketene, from acetone, 312, 703 

as an acetylating agent, 313 

with amines, 312 

cumulated double bonds in, 312— 
313 

dimers of, 313, 321 

with hydroxylic compounds, 312 

physical properties of, 312 
Ketimmes, hydrolysis of, 233 
/3-Keto esters, hydrolysis of, 354 

synthesis of, 352-354 
12-Ketocholanic acid, 798 
Ketones, with acid halides, 289-290 

from acyl chlorides, 279 

by acylation of arenes, 565 

from alcohols, 259, 279 

aldol addition to, 307-309 

from alkenes, 279 

from alkylacetoacetic esters, 354 

from alkynes, 279 

with amines, 288-289, 433 



Clemmensen reduction of, 291- 

292 
in condensation reactions, 354 
from dicarboxylic acids, 716 
from 1,2-glycols, 279, 280 
halogenation of, 303-306 
with hydrogen cyanide, 283 
hydrogenation of, 290 
nomenclature of, 275-276 
from organocadmium compounds, 

279 
from organomagnesium com- 
pounds, 232-233 
physical properties of, 277 
preparation of, 278-281, 294 
reactions of, 281-294 
reduction of, 290, 291-292, 566 
spectroscopic properties of, 277- 

278 
synthesis of large ring, 771 
a,j3-unsaturated, 311-312 
uv spectra of, 334 
Wolff-Kishner reduction of, 292 

Kharasch, M. S., 94 

Kiliani-Fischer cyanohydrin synthesis, 
417 

Kinetic control, 94 

in electrophilic aromatic substitu- 
tion, 569 

Kinetics, of S N reactions, 196-197 

Kolbe electrolysis, 342 

Kolbe reaction, 633 

Korners absolute method, 567 

Kraft process, in paper manufacture, 
411 

Kiirster, W., 681 



Lactams, 467-468 

from amino acids, 715 

nylon (6) from, 751 
^-Lactams, in penicillin G, 468 
Lactic acid, from 2-bromopropanoic 
acid, 502 

configuration of, 381-382 

formation of, 343 
Lactides, from hydroxy acids, 356 
Lactones, from hydroxy acids, 356, 715 

from unsaturated acids, 355 
/3-Lactones, from ketenes, 313 
Lactose, structure of, 408-409 
Ladenburg, A., 128, 708 



general index 830 



Ladenburg benzene, 707 
Lanosterol, 781 

biogenesis of, 792-795 

isomer of, 794-795 

from squalene, 792-795 

in wool fat, 792 
LeBel, J. A., 3 

and tetrahedral carbon atom, 385 
Lead fluoride, physical properties of, 14 
Lead oxide, as antiknock agent, 58 
Leaving groups, in S N reactions, 202 
Leucine, physical properties of, 459 
Lewis acids, in aromatic halogenation, 
563 

organoboranes, 536 
Lexan, preparation of, 749 
Light absorption {see also Electronic 
absorption spectra), 696-699 

and chemical structure, 699-703 
Lignin, 411 
Limonene, 780 
Lindlar catalyst, 726 
Linear free-energy relations, 661-664 
Linoleic acid, 330 
Lipids, 262, 413 

Lithium aluminum hydride, with 3- 
butenoic acid, 340 

with carboxylic acids, 340 

decomposition of, 292 

with formaldehyde, 291 

with ketones, 291 

with nitriles, 280 

reduction of acid derivatives by, 
349 
Lithium hydride, 291 
Lithium nitride, formation of, 511 
Longuet-Higgins, H. C, 726 
LSD {see Lysergic acid diethylamide) 
Lucas reagent, 255 
Luciferin, 688 
Lycopene, 782 

uv spectrum of, 166 
Lysergic acid, 683 

Lysergic acid diethylamide (LSD), 683 
Lysine, isoelectric point of, 463 
physical properties of, 459 
Lysozyme, properties and function of, 
478 



Magnesium, as inorganic coenzyme, 
484 



Maleic acid, acid dissociation con- 
stant of, 357 
anhydride from, 358 
dehydration of, 86 
physical properties of, 357 
Maleic anhydride, formation of, 358 

from maleic acid, 86 
Malic acid, as resolving agent, 385 
Malonic acid, acid dissociation con- 
stant of, 357 
decarboxylation of, 341, 359 
physical properties of, 357 
Malonic ester synthesis, of carboxylic 

acids, 336, 361 
Malonic esters, barbituric acids from, 

686 
Malonic esters, in synthesis of carbox- 
ylic acids, 336, 361 
Maltase, 410 
Maltose, hydrolysis of, 410 

structure of, 408-409 
Mandelic acid, as resolving agent, 385 
D-Mannose, phenylosazone from, 403 

structure of, 401 
Markownikoff 's rule, 92-96 

in additions to alkynes, 114 
Martius Yellow, 700, 759 
Mass spectrometry, 158-159 
molecular weights from, 773 
rearrangements in, 778 
in structure determination, 772- 
778 
Mass spectrum, of aspidospermine, 
776 
and CH 5 ®, 13 

of 3,5-dimethylpyridine, 773-775 
of ethyl butanoate, 778 
of 2-, 3-, and 4-ethylpyridine, 773- 

775 
of quebrachamine, 772-778 
Mayo, F. R., 94 
Menadione {see 2-Methyl-l,4-naphtho- 

quinone) 
Menthol, 780 

Menthone, cyanohydrin of, 284 
Mercaptans {see Thiols) 
Mercaptoethanol {see Ethan- l-ol-2- 

thiol) 
Mercuric salts, in alkyne hydration, 

114 
Merophan, 429 
Merrifield, R. B., 472 



general index 831 



Mescaline, 659 

Mesityl oxide, formation of, 309 

Mesitylene (see 1,3,5-Trimethylben- 

zene) 
Messenger RNA, 485 
Mestranol, 789 
Metabolism, 495-513 
Metal chelates, of carbonyl com- 
pounds, 315 
Metal hydrides, in reduction of car- 
bonyl compounds, 291 
Metallocenes, 235 
Metals, as prosthetic groups, 476 
Methacrolein, 276 
Methacrylic acid, 276 
Methanal (see Formaldehyde) 
Methane, acid dissociation constant of, 
13, 229 
base ionization constant of, 13 
bond angles in, 8 
bonding in, 6 
calculation of heat of combustion 

of, 24 
chlorination of, 222 
derivatives of, 9 
heat of combustion of, 24 
heat of monofluorination, 75 
from methylmagnesium iodide, 

230 
model of, 20 
molecular shape of, 20 
in natural gas, 56-57 
physical properties of, 10, 55, 535 
thermodynamic values for the 
chlorination of, 27 
Methanesulfonic acid, 521, 528 
Methanesulfonyl chloride, 521 
Methanethiol, 521 
Methanol (see Methyl alcohol) 
Methionine, 457 

oxidation of, 525 
physical properties of, 461 
2-Methoxyacetanilide, nitration of, 

574 
Methoxyacetic acid, acid dissociation 
constant of, 331 
physical properties of, 331 
Methoxybenzaldehydes, physical prop- 
erties of, 634-635 
4-Methoxybenzoin, formation of, 656 
l-Methoxy-l-buten-3-yne, from 1,3- 
butadiyne, 115 



1-Methoxyethanol, from acetaldehyde, 

284 
Methyl alcohol, with acetic acid, 253- 
254 
bond angles in, 10 
with hydrogen bromide, 251 
from methyl chloride, 196-198, 

200 
physical properties of, 12, 246 
preparation of, 245 
solubility in water, 12 
toxicity of, 245 
Methyl-2-aminobenzoate, 659 
Methyl anion, 33 
Methyl azide, 422 
Methyl bromide, reactivity toward 

nucleophiles, 204 
Methyl f-butyl ketone, cyanohydrin of, 

284 
Methyl cation, 32 
Methyl cellosolve, physical properties 

of, 264 
Methyl chloride, with chloride ion, 205 
with hydroxide, 195-198, 200, 496 
mechanism of formation from 

methane chlorination, 28 
from methane chlorination, 27, 60 
Methyl ethyl ether, formation of, 252 
Methyl ethyl ketone, cyanohydrin of, 
284 
with lithium aluminum hydride, 

373 
synthesis of, 354 
uv spectrum of, 277-278 
Methyl glucoside, formation of, 404 
Methyl iodide, physical properties of, 
217 
radicals from, 661 
Methyl isocyanate, 422 
Methyl isohexyl ketone, from choles- 
terol, 784 
Methyl isopropyl ketone, cyanohydrin 

of, 284 
Methyl methacrylate, 747 

polymerization of, 755 
Methyl nitrate, 422 
Methyl nitrite, 422 
Methyl propionate, 191 
Methyl radicals, from acetone, 703 

epr spectrum of, 661 
Methyl vinyl ether, physical properties 
of, 263 



general index 832 



Methyl vinyl ketone, with hydrogen 
cyanide, 311 

uv spectrum of, 278 

a-Methylallyl chloride (see 3-Chloro-l- 
butene) 

Methylamine, base dissociation con- 
stant of, 423 
from nitromethane, 431 • 
physical properties of, 12, 423 

N-Methylaniline, base dissociation con- 
stant of, 614 

infrared spectrum of, 424 
uv spectrum of, 614 

1-Methylanthracene, 551 

2-Methyl-l,3-butadiene (see Isoprene) 

2-Methylbutane, 49 
bromination of, 61 
chlorination of, 60 

2-Methyl-2-butanol, nmr spectrum of, 
173 

3-Methylbutanoyl chloride, forma- 
tion of, 340 

2-Methyl-l-butene, from /-pentyl 
chloride, 208 

2-Methyl-2-butene, formation of, 258 
from neopentyl iodide, 208 
from ?-pentyl chloride, 208 

2-Methyl-3-buten-2-ol, 191 

Methyl-7-butylcarbinol, dehydration 
of, 258 

l-Methyl-3-f-butyl-2,4,6-trinitroben- 
zene, in perfumes, 612 

3-Methyl-l-butyne, hydrogen bro- 
mide addition to, 1 19 
hydrogenation of, 1 19 

a-Methyl-a-carboxyglutaric acid, 786- 
787 

Methylchlorosilane, preparation of, 
534 

Methylcholanthrene, 798 

Methylcyclohexane, chair forms of, 
65 

frww-3-MethylcyclopentanoI, from 3- 
chloro-l-methylcyclopentane, 201 

Methylcyclopentanophenanthrene, 787 

Methylcyclopropane, physical proper- 
ties of, 83 

10-Methyl-2-decalone, optical rotatory 
dispersion curves for, 390 

Methylene (see Carbene) 

Methylene dichloride (see Dichloro- 
methane) 



Methylene glycol, dehydration of, 261 

from formaldehyde, 261 
Methylenecyclobutane, 82 
Methylenecyclobutene, formation of, 

428 
Methylenecyclohexane, synthesis of, 

531 
Methylenetriphenylphosphorane, with 

cyclohexanone, 531 
Methylethylamine, 191 
3-Methylhexane, from 4-methyl-2- 

hexene, 86 
4-Methyl-2-hexene, bromine addition 
to, 86 
hydrogenation of, 86 
5-Methyl-2-hexene, 103 
Methyllithium, formation of, 228 
Methylmagnesium iodide, bonding in, 
230 
with ethanol, 230 
with formaldehyde, 231 
from methyl iodide, 229 
Methylmalonic acid, formation of, 343 
a-Methylnaphthalene, 551 
/S-Methylnaphthalene, 551 
2-Methyl-l ,4-naphthoquinone, 640 
Methyloxonium bisulfate, formation 

of, 254 
Methyloxonium bromide, formation 

of, 251 
2-Methylpentane, 49, 55 
3-Methylpentane, 49, 55 
4-Methyl-3-penten-2-one (see Mesityl 

oxide) 
2-Methylpropane, in natural gas, 57 
naming of, 48 

physical properties of, 11, 83 
shape of, 49 
3-Methylpropanoic acid, 190 

acid dissociation constant of, 331 
physical properties of, 331 
2-Methyl-l -propanol (see Isobutyl 

alcohol) 
2-Methyl-2-propanol (see /-Butyl alco- 
hol) 
2-Methylpropanal (see Isobutyralde- 

hyde) 
2-Methylpropenal (see Methacrolein) 
2-Methylpropene, from ?-butyl alcohol, 
257 
from ?-butyl chloride, 206-207 
dimerization of, 101 



general index 833 



hydration of, 88 
isobutyraldehyde from, 281 
physical properties of, 84 
polymerization of, 101, 755 
polymers from, 746 
2-Methylpropenoic acid (see Metha- 

crylic acid) 
2-Methyl-2-propenonitrile, from ace- 
tone, 282 
5-(l-Methylpropyl)decane, naming of, 

52 
N-Methylpyrrolidone, solvent proper- 
ties of, 437 
Methylsilane, physical properties of, 

535 
Methylsodium, ether cleavage with, 264 
a-Methylstyrene, polymerization of, 

755 
2-(Methylthio)ethanol, 524 
Methyltrichlorosilane, physical prop- 
erties of, 535 
Mevalonic acid, in biogenesis, 792 
Micelles, 332-333 

catalysis by, 333 
Michaelis-Menten effect, 504 
Microwave spectroscopy, bond lengths 

and bond angles from, 160 
Mitochondria, 509, 640 
Models, ball-and-stick, 20, 22 
of cyclohexanes, 64-67 
space-filling, 22 
Molecular orbital energies, of 1,3- 
butadiene, 142 
of conjugated cyclic systems, 143 
of ethene, 142 
of trimethylenemethyl, 143 
Molecular orbital theory, 140-145 
and concerted reactions, 726-730 
LCAO approach, 727 
Molecular orbitals, antibonding, 727 
bonding, 727 

of 1,3-butadiene, 698, 727-729 
energies of, 160 
of ethene, 727-729 
in methane, 1 30 
Molecular rotation, definition of, 372 
Molecular shapes, importance of, 8 
Molecular weights, from boiling point 
elevations, 157 

by end group analysis, 157 
from freezing-point depressions, 
157 



from light scattering, 157 

by mass spectrometry, 157-158 

osmotic pressure, 1 57 

of proteins, 476 

from sedimentation rates, 157 

from vapor density measurements, 
157 

from viscosity measurements, 1 57 
Molozonide, 98 
Monosaccharides, 400-408 
Monoterpenes, 779-780 
Morphine, 684 
Muscone, 769 

Mustard gas, toxicity of, 429 
Mutarotation, of glucose, 406 
Myoglobin, properties and function of, 

478 
Myrcene, 779 



NAD (see Nicotinamide-adenine di- 

nucleotide) 
NADH (see Nicotinamide-adenine di- 

nucleotide) 
Naphthacene, 552 

uv spectrum of, 557 
Naphthalene, acylation of, 575-576 
from azulene, 578 
bond lengths in, 574 
monosubstitution products of, 551 
physical properties of, 553 
reactions of, 575-576 
resonance hybrid of, 575 
sulfonation of, 575-576 
uv spectrum of, 557 
1-Naphthalenesulfonic acid, from 

naphthalene, 575-576 
1-Naphthol, acid dissociation con- 
stant of, 629 

uv spectrum of, 629 
2-Naphthol, acid dissociation constant 
of, 629 
formylation of, 653 
2-Naphthyl methyl ketone, as photo- 

sensitizer, 707 
1-Naphthylamine, basicity of, 614 

uv spectrum of, 614 
2-Naphthylamine, basicity of, 614 

uv spectrum of, 614 
1 -Naphthylphenylmethylsilane, enan- 

tiomers of, 533 
Narcotine, 684 



general index 834 



Natta, G., 103 

Natural gas, composition of, 56-57 

Natural products, 769-796 
polyhetero, 688 
related to indole, 682-683 
related to isoquinoline, 684 
related to pteridine, 686 
related to purine, 686 
related to pyran, 686-688 
related to pyridine, 684 
related to pyrimidine, 685-686 
related to pyrrole, 680-682 
related to quinoline, 684 

Natural rubber, 742 
with ozone, 98 

Neo, definition of, 49 

Neohexane (see 2,2-Dimethylbutane) 

Neopentane (see 2,2-Dimethylpropane) 

Neopentyl halides, in S N 2 reactions, 
202 

Neopentyl iodide, formation of, 231 
solvolysis of, 208 

Neoprene, 746 

Nerol, 780 

Newton, I., 3 

Niacin (see Nicotinic acid) 

Nickel, in catalytic hydrogenation, 35 

Nicol prism, and polarized light, 369 

Nicotinamide, in NAD®, 506 

Nicotinamide-adenine dinucleotide 
(NAD®), 506, 684 

Nicotine, 684 

Nicotinic acid, 684 

Ninhydrin, color test for a-amino 
acids, 463^64 
structure of, 464 

Nitration (see also Electrophilic aro- 
matic substitution) 
of alkanes, 61 
of arenes, 561-562 
of heterocyclic compounds, 674- 

677 
of 2-methoxyacetanilide, 574 
of monosubstituted benzenes, 568 
of p-nitrotoluene, 573 
of phenanthrene, 577 
of trifluoromethylbenzene, 568 

Nitric acid, 422 

Nitric oxide, 422 

Nitrile oxide, as 1,3-dipole, 721-723, 
731 

Nitriles, alkylation of, 440 



hydrolysis of, 336, 439 

with lithium aluminum hydride, 
280, 349 

with organomagnesium com- 
pounds, 233 

properties of, 440 

reduction of, 280, 349, 430 
Nitrite esters, 442 

Nitro compounds (see also Nitro- 
alkanes), 441-442 

from alkyl halides, 442 

from amines, 433, 442 

aromatic, 606-613 

reduction of, 430, 591, 608-611 

synthesis of, 606-608, 618-619 
Nitro groups, effects of, 608 

resonance in, 441 
Nitroacetic acid, decarboxylation of, 

341 
Nitroalkanes (see also Nitro com- 
pounds), from alkanes, 61 

electronic spectra of, 441 

infrared spectra of, 441 

naming of, 53 
m-Nitroaniline, basicity of, 614 

uv spectrum of, 614 
p-Nitroaniline, basicity of, 614 

formation and diazotization of, 
606 

oxidation of, 606 

resonance in, 616 

uv spectrum of, 614 
Nitrobenzene, 568 

chlorination of, 569, 571 

nmr spectrum of, 558 

physical properties of, 608 

radical anion from, 613 

reduction of, 609 
m-Nitrobenzoic acid, 568 
/j-Nitrobenzoic acid, synthesis of, 606 
3-Nitrobiphenyl, 619 
3-Nitro-4'-chlorobenzoin, 656 
Nitroethane, 61 
Nitrogen, fixation of, 511 
Nitrogen dioxide, 422 
Nitrogen mustards, in cancer therapy, 
429 

with DNA, 429 
Nitroglycerin, from glycerol, 262 
5-Nitro-2-indanone, oxidation of, 649 
Nitromethane, 61, 422 

reduction of, 431 



general index 835 



3-Nitro-2-methylpentane, 53 
2-Nitro-2-methylpropane, formation 

of, 433 
o-Nitrophenol, physical properties of, 

633 
p-Nitrophenol, acid dissociation con- 
stant of, 629 

resonance, in 700 

uv spectrum , 629, 700 
jP-Nitrophenolate, resonance in, 700 

uv spectrum of, 700 
Nitrophenols, nmr spectra of, 635, 637 

physical properties of, 634-635 
4-Nitrophthalic acid, formation of, 649 
Nitropropanes, 61 
Nitroso compounds, 440-441 

from aromatic amines, 617 

rearrangement of, 617 
Nitrosobenzene, 609 

dimer of, 610 

formation of, 610 

with N-phenylhydroxylamine, 
610-611 
/7-Nitroso-N,N-dimethylaniline, 617 
Nitrosomethane, 422 
2-Nitrosopropane, rearrangement of, 

441 
m-Nitrotoluene, preparation of, 608 
p-Nitrotoluene, nitration of, 562, 573 

oxidation of, 606 
Nitrotoluenes, formation of, 562 
Nitrous acid, 422 

with amines, 432 

with aromatic amines, 616-617 
Nitrous oxide, 422 
Nmr (see Nuclear magnetic resonance 

spectroscopy) 
Nomenclature, of alcohols, 187-190 

of alkenes, 81-83 

of alkyl halides, 187-190 

of alkynes, 111 

of amines, 421^423 

of ammonium salts, 421-423 

of annulenes, 725 

of benzene derivatives, 549-552 

of benzoins, 656 
Nomenclature, of carboxylic acids, 
190, 329 

of carboxylic esters, 191 

of ethers, 190 

of halogenated hydrocarbons, 37 

of heterocyclic compounds, 671 



IUPAC rules for, 19 

of organoboron compounds, 539 

of organosulfur compounds, 520- 

521 
single- or multiple-word names, 

191-192 
use of Greek letters, 191 

Nonane, isomers of, 48 

physical properties of, 55, 63 

Nonbenzenoid compounds, 578 

Norbornene, with phenyl azide, 722 

Norethindrone, 789 

Nuclear magnetic resonance spectra 
(nmr), of alcohols, 247-249 
of amides, 435 
of amines, 425 
of arenes, 558 
of carboxylic acids, 334 
of fiuxional systems, 723-724 
and hydrogen bonding, 635 
of a, /J-unsaturated carbonyl com- 
pounds, 311 

Nuclear magnetic resonance spectros- 
copy, 168-179 

the chemical shift, 170-171 
in qualitative analysis, 174-176 
and rate processes, 176-178, 696 
spin-spin splitting, 171-174 

Nuclear magnetic resonance spectrum, 
of l-chloro-2-fluoro-l ,1,2,2-tetra- 
bromoethane, 178 
of cholesterol, 787 
of dietbylamine, 425 
of 1,1-dimethoxyethane, 175 
of ethyl acetoacetate, 350-352 
of ethyl alcohol, 169 
of ethyl iodide, 171 
of 2-methyl-2-butanol, 173 
of nitrobenzene, 558 
o-, m-, p-nitrophenols, 635, 637 
of 2,4-pentanedione, 314-315 
of propanamide, 437 

Nucleic acids (see Ribonucleic acid and 
Deoxyribonucleic acid) 

Nucleophiles, base dissociation con- 
stants, of, 204 
definition of, 87, 192 
in electrophilic addition, 89 

Nucleophilic addition, to alkynes, 115 

Nucleophilic aromatic substitution, of 
activated aryl halides, 593-594 
elimination-addition mechanism, 



general index 836 



594-596 
rearrangements in, 595 

Nucleophilic catalysis, 496-497, 499-500 

Nucleophilic displacement reactions 
(see also Nucleophilic aromatic sub- 
stitution), 192-205 
of activated aryl halides, 593-594 
of alkyl derivatives, 193-195 
of alkyl halides, 193-195 
catalysis by heavy metal salts, 203 
energetics of, 197-200 
of heterocyclic compounds, 678- 

679 
kinetics of, 196-197 
the leaving group in, 202-203 
mechanisms of, 195-197 
nature of solvent, 204-205 
reagents for, 193-195 
S N /, 502 

solvents for, 193-195 
stereochemistry of, 200-201, 389 
structural effects in, 201-204 
synthetic utility of, 193-195 

Nucleophilic reagents, and basicity, 
203-204 
list of, 193-194 

Nucleosides, 480-481 

Nucleotides, 481 

Nylon, 750-751 

hydrogen bonding in, 739 



Octafluorocyclobutane, physical prop- 
erties of, 226 

as a propellant, 226 
Octane, heat of combustion of, 24, 57, 
510 

infrared spectrum of, 163 

isomers of, 48 

octane rating of, 58 

physical properties of, 53-55, 63 
Octane rating, 57-58 
1-Octanol, from 1-hexanol, 265 
Oil of wintergreen, 658 
Olefins (see also Alkenes), 81 
Oleic acid, 330 
Opium alkaloids, 684 
Optical activity, origin of, 369-371 
Optical isomerism, 369-392 

and absolute and relative con- 
figuration, 381-384 

of allenes, 379-380 



of amine oxides, 434 

of amines, 426 

asymmetric induction, 386-387 

of biphenyls, 380-381 

in 2-butanol, 369 

conventions for, 374-375, 381-384 

and optical rotatory dispersion, 

389-391 
of organophosphorus compounds, 

525 
of organosilicon compounds, 

533 
of organosulfur compounds, 524- 

525 
projection formulas, 374-375 
resolution (see Resolution) 
and restricted rotation, 379-381 
of spiranes, 379-380 
Optical isomers, definition of, 369 
diastereomers, 378 
enantiomers, 384 
meso forms, 377 
resolution of, 384 
threo and erythro forms, 377 
Optical rotatory dispersion, 389-391 

of 10-methyl-2-decalone, 390 
Oral contraceptives, 789 
Orbital symmetry, and cycloaddition 

reactions, 726-730 
Orbitals, antibonding, 141, 277 
bonding, 145 
definition of, 7, 129 
in ethene, 34 
nonbonding, 141 
shapes of, 129, 517-518 
d Orbitals, and chemical bonds, 517- 
520 
in organosilicon compounds, 533- 
534 
p Orbitals, 129 
s Orbitals, 129 
d 2 sp 3 Orbitals, 519 
sp 2 Orbitals, 131 
sp 3 Orbitals, 130, 519 
ORD (see optical rotatory dispersion) 
Organic chemistry, definition of, 3 
Organic nitrogen compounds, 421- 

426 
Organic synthesis (see Synthesis) 
Organoboranes, as Lewis acids, 536 
Organoboron compounds, 536-540 
nomenclature of, 539 



general index 837 



Organocadmium compounds, with 
acyl chlorides, 233, 279 
ketones from, 233 
Organofluorine compounds, physio- 
logical properties of, 226 
Organolithium compounds, 234 
Organomagnesium compounds, with 
acids, 230 
with 1-alkynes, 229 
with carbonyl compounds, 231- 

233, 250 
carboxylic acids from, 336 
with cyclopentadiene, 229 
with halogens, 230-231 
with metallic halides, 229 
with nitriles, 233 
with oxygen, 230-231 
with sulfur, 230-231 
thiols from, 522 
Organomercury compounds, from 
Grignard reagents, 229 
from methyl iodide, 228 
as seed fungicides, 234 
Organometallic compounds {see also 
individual types such as Organomag- 
nesium compounds), 226-235 
from aryl halides, 592-593 
bonding in, 227, 230 
preparation of, 228-229 
toxicity of, 229 
use of, 234 
Organonitrogen compounds {see Or- 
ganic nitrogen compounds) 
Organophosphorus compounds, 528- 
531 
asymmetry of, 525 
Organosilicon compounds {see also 
individual types such as Silanes, Sil- 
anols, etc.), 531-536 
asymmetry of, 533 
preparation and properties, 534- 

535 
types of, 532 . 
Organosodium compounds, 234 

Wurtz coupling of, 228 
Organosulfur compounds, 520-528 

nomenclature of, 520-521 
Orientation, in addition polymeriza- 
tion, 754 
in electrophilic aromatic subsitu- 

tion, 567-574 
in nitration of arenes, 568 



Orion {see Polyacrylonitrile) 
Ortho-Novum, 789 
Osazones, formation of, 403 
Osmium tetroxide, in alkene oxidation, 

261 
Oxalic acid, acid dissociation constant 
of, 357 
decarboxylation of, 359 
physical properties of, 357 
Oxidation, of alcohols, 259-260, 523, 
526 
of alkenes with osmium tetroxide, 

98-99 
of alkenes with ozone, 97 
of alkenes with permanganate, 

98-99 
of alkylboranes, 97 
of amines, 432-433, 523 
of cholesterol, 784 
of civetone, 770-771 
of cumene, 627 
of cyclohexane, 750 
of enols, 292 

of ethanol by NAD®, 506-508 
of eugenol, 658 
of glucose, 509 
of hydroquinones, 638 
and metabolic processes, 510 
microbiological, 788 
photochemical, 705-706 
of sulfides, 525 
of thiols, 523-524 
Oxidation levels, of nitrogen, 421 
Oxidative phosphorylation, 509 
Oximes, formation of, 289 
reduction of, 430 
rearrangement of, 429^130, 439 
Oxomalonic acid, 282 
Oxonium ions, from carbonyl com- 
pounds, 285-286 
in enol formation, 304 
from ethers, 264 

in hemiacetal and acetal forma- 
tion, 285-286 
Oxygen, excited singlet state of, 705- 
706 
as inhibitor in chain reactions, 32 
triplet state of, 705 
Ozone, air pollution from, 58 

as 1,3-dipole, 721-723, 731 
Ozonides, 97 
Ozonization, 97-98, 279 



general index 838 



Palladium, in catalytic hydrogenation, 
35 
in reduction of acyl chlorides, 279 
Palmitic acid, 330 

acid dissociation constant of, 331 
physical properties of, 331 
Papain, properties and function of, 

478 
Papaverine, 684 
Paper chromatography, in amino acid 

analysis, 466 
Paraffin, 56 

Paraformaldehyde, from formalde- 
hyde, 287-288 
Partial rate factors, in nitration of 

arenes, 569 
Pasteur, L., 385 
Patterson, A. M., 552 
Pauli exclusion principle, 1 30 
Pelargonidin chloride, 687-688 
Pencillin, in resolution of enantiomers, 

385 
Penicillin G., /3-lactam structure of, 468 
X-ray diffraction analysis of, 468 
Pentaborane, structure of, 537-538 
Pentacene, uv spectrum of, 557 
Pentadecane, physical properties of, 55 
1,3-Pentadiene, 83 
1,4-Pentadiene, 83 

Pentaerythritol, preparation of, 324 
Pentane, isomers of, 48-49 

physical properties of, 55, 63, 424 
2,3-Pentanediol, optical isomers of, 

392 
2,4-Pentanedione, acid dissociation 
constant of, 315 
Be" salt of, 316 
Cu" salt of, 316 
enol form of, 314 
nmr spectrum of, 314-315 
tautomerization of, 315, 723 
Pentanoic acid, acid dissociation con- 
stant of, 331 
physical properties of, 331 
m-2-Pentene, heat of combustion of, 
85 
physical properties of, 85 
?ra«.s-2-Pentene, heat of combustion of, 
85 
physical properties of, 85 
4-Pentenoic acid, y-valerolactone 
from, 355 



f-Pentyl chloride (see 2-Chloro-2- 

methylbutane) 
n-Pentylamine, physical properties of, 

424 
Peptide synthesis, coupling reagents 
in, 472 

protecting groups in, 471 

solid-phase, 472 

yields in, 471 
Peptides, 468-474 

amino acid sequence in, 470 

analysis of, 469-470 

with 2,4-dinitrofluorobenzene, 593 

naturally occurring polymers, 
756-757 

synthesis of, 470-474 
Peracids, with amines, 433 

in radical polymerization, 754 
Perfluoro-2-methylpropene, toxicity of, 

226 
Perfumes, 4 

civetone in, 769 

coumarins in, 687 

cyclopentadecanone in, 772 

exaltone in, 772 

examples of, 658-659 

ingredients in, 612 

muscone in, 769 

terpene alcohols in, 780 
Perhydrophenanthrene, 551 
Periodic table, 517 

Permanganate oxidation, of alco- 
hols, 260 

of alkenes, 261, 287 

of amines, 432 
Peroxides, as radical initiators, 

95 
Peroxy radicals, 32 

Persulfuric acid, in radical polym- 
erization, 754 
Pesticides, 4 

analysis for, 156 

organochlorine derivatives, .596- 
598 
Petroleum, 56-58 
Phenacetin, 659 
Phenalenyl, 552 

Phenanthrene, monosubstitution prod- 
ucts of, 551 

physical properties of, 553 

reactions of, 577 

resonance hybrid of, 575 



general index 839 



1-Phenanthrylamine, diazotization of, 

591 
Phenobarbital, 659, 685 
Phenol, acid dissociation constant of, 
629 
from chlorobenzene, 595 
commercial synthesis of, 627 
physical properties of, 628-629 
polymers from, 751-752 
resonance hybrid of, 630 
stabilization energy of, 605 
tautomer of, 605 
uv spectrum of, 556, 629 
Phenol-formaldehyde resins, 751-752 
Phenolphthalein, structure of, 710 
Phenols, acidity of, 630 
alkylation of, 631-632 
from aromatic amines, 617 
bromination of, 632-633 
chemical properties of, 630-635 
with ferric chloride, 631 
hydrogen bonding in, 634 
nmr spectra of, 635, 637 
oxidation of, 635-636 
physical properties of, 627-629, 

633 
polyhydric, 635-636 
quinones from, 635-636 
reduction of, 635 
separation from carboxylic acids, 

631 
synthesis of, 627-629 
Phenoxide anion, resonance hybrid 

of, 630 
Phenyl acetate, formation of, 630 
Phenyl allyl ether, from phenol, 631 

rearrangement of, 632 
Phenyl azide, addition reactions of, 722 
Phenyl halides, in S N reactions, 220 
Phenyl radicals, from diazonium salts, 

618 
Phenylacetic acid, acid dissociation 
constant of, 331 
physical properties of, 331 
Phenylacetylene, 550 

infrared spectrum of, 164 
Phenylalanine, physical properties of, 
459 
synthesis of, 458 
Phenylamine {see Aniline) 
/i-Phenylbenzaldehyde, formation of, 
652 



l-Phenyl-2-butanol, 187 
l-Phenyl-3-buten-l-ol, 189 
Phenyldichloroborane, 539 
^-Phenylenediamine, uv spectrum of, 

614 
Phenylethanone {see Acetophenone) 
Phenylethene {see Styrene) 
2-Phenylethyl bromide, 589 
Phenylhydrazine, osazones from, 403 
N-Phenylhydroxylamine, 609 

from nitrobenzene, 610 
Phenyllithium, 191 

preparation of, 593 
Phenylmagnesium bromide, prepara- 
tion of, 592 
Phenylmagnesium iodide, 191 
2-Phenylpyridine, from pyridine, 678 
Phenyltrichloromethane, physical 

properties of, 535 
Phenyltrichlorosilane, physical prop- 
erties of, 535 
Phenyl trimethylammonium salts, ni- 
tration of, 568 
Phenyltrimethylmethane, physical 

properties of, 535 
Phenyltrimethylsilane, physical prop- 
erties of, 535 
Phillips process, in coordination poly- 
merization, 103 
Phloroglucinol, reactivity of, 636 

structure of, 635 
Phosphate groups, in DNA, 479 
Phosphines, asymmetric, 525 
Phosphonate esters, hydrolysis of, 502 
Phosphonium salts, with basic re- 
agents, 530 
formation of, 530 
reactions of, 530 
Phosphorescence, 696-699 
Phosphoric acid, derivatives of, 528- 

530 
Phosphorous acid, structure of, 

529 
Phosphorus, in halogenation of acids, 
343 
organic compounds of, 528-531 
Phosphorus halides, with carboxylic 

acids, 340 
Phosphorus pentachloride, in Beck- 
mann rearrangement, 430 
with cyclopentanone, 290 
Phosphorus pentafluoride, 517 



general index 840 



Phosphorus tribromide, in alkyl bro- 
mide formation, 256 
Photochemistry, 695-709 
Photodissociation, 703-704 
Photographic developers, 639 
Photosensitizers, benzophenone, 706 
definition of, 706 
2-naphthyl methyl ketone, 707 
in photoisomerization, 707 
Photosynthesis, of carbohydrates, 399 
of glucose, 387 
radicals in, 661 
study of, 466 
Phthalic acid, acid dissociation con- 
stant of, 357 
anhydride from, 358, 715 
physical properties of, 357 
Phthalic anhydride, formation of, 358, 
715 
polymer from, 750 
Phytol, 781 

in chlorophyll, 781 
in vitamin K, 781 
Picric acid, 611-612 

acid dissociation constant of, 629 
charge- transfer complexes of, 612 
uv spectrum of, 629 
Pimelic acid, acid dissociation constant 
of, 357 
from civetone, 770 
cyclohexanone from, 358 
physical properties of, 357 
Pinacol, rearrangement of, 280 
Pinacolone, from pinacol rearrange- 
ment, 280 
a-Pinene, camphor from, 780 
Piperidine, base dissociation constant 
of, 423 

physical properties of, 423 
Piperonal, 659 

Plane-polarized light (see also Optical 
Isomerism), and origin of optical 
rotation, 369-371 
Plasticizers, 748 

camphor, 780 
Plastics (see also Polymers), from cellu- 
lose, 410 
from formaldehyde, 288 
thermal moulding of, 740 
Platinum, in catalytic hydrogenation, 

35 
Polarimeter, description of, 370-371 



Polarizability, and S N reactions, 203 
Pollution, atmospheric, 58 

by detergents, 526 

from paper pulping, 41 1 

from petroleum, 58 

by phosphates, 528 

trace analysis for, 156 
Polyacrylonitrile, 740 
Polyamides, definition of, 469 

nylons, 750-751 
Polycyclopentadiene, 737 
Polyesters, preparation of, 749-750 
Polyethylene (see also Polythene), 102 
Polyethylene glycol, from ethylene 
oxide, 266 

uses of, 756 
Polyhalogen compounds, 222-226 
Polyhetero systems, examples of, 679- 

680 
Polyisobutylene, 101 
Polyisoprene (see also Natural rubber), 

98, 742-743 
Polymerization, by addition, 753-756 

anionic, 100 

azo initiators in, 443 

cationic, 101, 755-756 

by condensation, 748-752 

coordination, 103, 753 

of cyclopentadiene, 737 

ionic, 755-756 

of isopentenyl pyrophosphate, 792 

of isoprene, 743 

radical, 102, 743, 753-755 

of tetrafluoroethene, 225 

vinyl, 753 

Ziegler, 743 
Polymers, 737-766 

from aldehydes, 287-288 

atactic, 744 

cold-drawing, 740 

cross linking, 738 

crystalline state of, 740 

definition of, 99 

DNA, 479 

dyeing of, 758-762 

elastomers, 738, 741 

glass temperature of, 740 

glyptal resin, 750 

isotactic, 744 

melting temperature, 740 

naturally occurring, 756-758 

nylons, 750 



general index 841 



in peptide synthesis, 472 
phenol-formaldehyde resins, 751— 

752 
physical properties of, 738-748 
preparation, 748-756 
representative, 745-747 
silicones, 533, 536 
stereoregular, 753 
syndiotactic, 744 
thermoplastic, 738 
thermosetting, 738 
types of, 737-738 
vulcanization of, 743 
Polynitro compounds, 611-612 

charge-transfer complexes of, 612- 
613 
Polynuclear aromatic hydrocarbons, 
double bond character in, 574 
naming of, 551-552 
substitution reactions of, 574 
Polyoxymethylene {see Formaldehyde, 

polymers from) 
Polypeptides {see Peptides) 
Polypropene, 100 
isotactic, 744 
Polypropylene {see Polypropene) 
Polysaccharides {see Cellulose, Starch, 

etc.) 
Polystyrene, configurations of, 744 
Polystyrene resin, in peptide synthesis, 

472 
Polytetrafluoroethene {see Teflon) 
Polythene, 745 

from ethene, 102, 745 
crystalline character of, 739 
Polyvinyl chloride, 745 
"Polywater", 11 

Potassium bromide, in infrared spec- 
troscopy, 161 
Potassium f-butoxide, in E2 elimina- 
tion, 252 
Potassium permanganate {see Perman- 
ganate oxidation) 
Potential energy, of diatomic mole- 
cules, 697 
Progesterone, 790 

cortisone from, 788 
Projection formulas {see also Optical 
isomerism), of 2-butanol, 374 
conventions for, 375, 383 
Fischer type, 405 
Haworth type, 405 



Proline, in ninhydrin test, 465 

physical properties of, 461 
Propadiene {see Allene) 
Propanamide, infrared spectrum of, 
436 

nmr spectrum of, 437 
Propane, with bromine, 219 

in natural gas, 57 

nitration of, 61 

physical properties of, 11, 55, 63 
1,3-Propanedithiol, with acetone, 523 
Propanethiol, occurrence, 522 
Propanoic acid, 190 

acid dissociation constant of, 331 

cleavage of alkyl boranes, 97 

physical properties of, 331 
Propanone {see Acetone) 
Propanoyl chloride, with benzene, 566 
Propargyl chloride {see 3-Chloro-l- 

propyne) 
Propenal {see Acrolein) 
Propene, addition reactions of, 104 

with chlorine, 220 

hydroboration of, 539 

with hydrogen halides, 219 

from isopropyl iodide, 252 

polymer from, 745 

from propyne, 121 
Propanoic acid {see Acrylic acid) 
2-Propen-l-ol {see also Allyl alcohol), 

262 
Propionamide {see Propanamide) 
Propionic acid {see Propanoic acid) 
n-Propyl alcohol, from tri-«-propyl- 

borane, 97 
Propyl bromide, from propane, 219 

from 1-propanol, 219 

from propene, 95 
re-Propyl chloride, physical properties 

of, 217 
ra-Propylbenzene, formation of, 566 

physical properties of, 553 
Propylene {see also Propene), 82 
3-Propyl-l-heptene, naming of, 81 
Propyne, acid dissociation constant of, 
229 

addition reactions of, 121 

hydration of, 114 
Prosthetic groups, 503 

in hemoglobin, 680 

metals as, 476 

in proteins, 469 



general index 842 



Protecting groups, benzyloxycarbonyl, 
471 

f-butoxycarbonyl, 472 
in peptide synthesis, 471 
removal of, 471, 473 
Proteins, 4 

biological functions of, 477 
biosynthesis of, 477 
conformations of, 474 
denaturation of, 469 
heat of combustion of, 510 
in hemoglobin, 680 
isoelectric points of, 476, 478 
as peptides, 469 
structures of, 474-477 
Proton magnetic resonance spectros- 
copy (see Nuclear magnetic reson- 
ance spectroscopy) 
Protonium ions, 91-92 
Prototropic change, 315 
Pteridine, 679, 686 

Purification, by chromatography, 156 
by crystallization, 153 
by distillation, 153 
by extraction, 153 
Purine, 679 

derivatives of, 407, 480 
natural products related to, 686 
Pyran, natural products related to, 
686-688 
structure of, 406 
Pyranose, definition of, 406 
Pyrene, 552 

3,10-Pyrenequinone, 638 
Pyridine, base strength of, 423, 427, 
431, 674 
chemical properties of, 673-679 
natural products related to, 684 
nucleophilic substitution reactions 

of, 678-679 
physical properties of, 423, 431, 
671 
Pyridinium salts, formation of, 674 
a-Pyridone, catalysis by, 503 
from pyridine, 678 
tautomers of, 503, 723 
Pyridoxine, 684 
Pyrimidine, 679 

derivatives of, 407 
natural products related to, 685- 
686 
Pyrogallol, structure of, 635 



a-Pyrone, 686 

y-Pyrone, resonance hybrid of, 686-687 
Pyrrole, acid dissociation constant of, 
674 
aromatic character of, 672-673 
atomic orbital description of, 672- 

673 
chemical properties of, 673-679 
hydrogen bonding in, 671 
physical properties of, 671 
natural products related to, 680- 

682 
resonance hybrid of, 672 
Pyrrylmagnesium bromide, formation 

of, 674 
Pyrrylpotassium, 674 
Pyrryl anion, resonance hybrid of, 674 



Quantum yield, definition of, 703 
Quartz, in separation of tautomers, 352 
Quebrachamine, mass spectrum of, 
772-777 

pyridines from, 773-775 

structure determination of, 772- 
777 
Quercitin, 687 
Quinhydrone, 638 
Quinine, as resolving agent, 385 
Quinoline, 679-680 

natural products related to, 684 
Quinone-hydroquinone equilibrium, 

638 
Quinones, 636-641 

addition reactions of, 639-640 

naturally occurring, 640 

from phenols, 635-636 

reduction of, 638 
Quinuclidine, basicity of, 622 



Racemization, by acid-catalyzed enol- 
ization, 388 

by base-catalyzed enolization, 388 

definition of, 373 

mechanisms of, 388 

by an S N 1 process, 388 
Radiation, effects of, 695 
Radical addition, to alkenes, 94-96 
Radical halogenation, of alkanes, 28- 
32, 59-61 

of alkylbenzenes, 650 



general index 843 



Radical polymerization, 753-755 
Radicals, acetyl, 703 

anion, 613 

alkoxy, 95 

benzhydrol, 704 

coupling of, 655 

methyl, 32, 703 

in photosynthesis, 661 

semiquinone, 638-639 

stabilities of, 96 

stable triarylmethyl, 654 
Raney nickel, 36 
Rayon acetate, 410 
Reaction (p) constants, of Hammett 

equation, 663-664 
Reaction intermediates, types of (see 

also Intermediates), 32 
Reaction rates, diffusion controlled, 495 

and entropy of activation, 30 

and heat of activation, 30 

and mechanism, 28 

and rate-determining step, 29 

and S N reactions, 196-197 
Reactions (see also individual types 
such as Hydrogenation, Nucleo- 
philic displacement, etc.) 

addition of unsaturated hydro- 
carbons, 34 

additions to alkynes, 113-115 

alkylation of esters, 354 

chain-termination in radical types, 
31 

Claisen condensation, 352-354 

combustion, 23 

condensation of carbonyl com- 
pounds with amines, 288-289 

conjugate addition, 127, 133-135 

cyclization, 715-730 

cycloaddition, 718-723 

electrophilic addition to alkenes, 
87-96 

electrophilic aromatic substitu- 
tion, 559-577 

elimination, 205-208, 252, 256 

endothermic, 23 

esterification, 252-255 

of excited states, 695-709 

exothermic, 23 

heat of, 23 

hydrogenation, 35 

nucleophilic addition to alkynes, 
115 



nucleophilic displacement, 192- 
205 

oxidation of alcohols, 259-260 

polymerization of alkenes, 99-103 

propagation, 31 

solvolysis, 197 

substitution, 26, 252 

thermodynamics of, 26 
Rearrangement, of benzilic acid, 313, 
321 

of cumene hydroperoxide, 627- 
628 

of glyoxal, 313, 321 

of hydrazobenzene, 611 

of a-pinene, 798 

of squalene, 795 
Rearrangements, of alkyl groups, 258, 
565 

in alkylation of arenes, 565 

Beckmann, 429^130 

Claisen, 632 

Cope, 723-724 

of dienes, 723-724 

of 1,2-glycols, 280 

in mass spectrometry, 778 

of N-nitroso compounds, 617 

of oximes, 429^130 

of phenyl allyl ethers, 632 

photochemical, 707 

of terpenes, 798 

of /3,y-unsaturated carbonyl com- 
pounds, 312 
Reducing sugars, definition of, 406 
Reduction, of carboxylic acids, 279- 
280, 340 

of civetone, 771 

of diazonium salts, 619 

of nitriles, 280 

of nitro compounds, 608-61 1 

photochemical, 704-705 

of quinones, 638 
Relative configuration, of optical iso- 
mers, 381-384 
Reserpine, 683 

Resolution, crystallization procedure 
for, 385 

definition of, 373 

of enantiomers, 384 

of optically active acids, 385 

of optically active alcohols, 385 

of optically active bases, 385 

Pasteur method for, 385 



general index 844 



Resonance, in carboxylate anions, 337 

in excited state of 1,3 -butadiene, 
168 
Resonance energy (see also Dereal- 
ization energy and Stabilization 
energy), 132 

of benzene, 549 
Resonance method, 13 

rules for, 140 
Resorcinol, acid dissociation constant 
of, 629 

reactivity of, 636 

structure of, 635 

uv spectrum of, 629 
Restricted rotation, and optical activ- 
ity, 379-381 
Rf value, definition of, 466 
Ribofuranose, in vitamin B 12 , 681- 682 
Ribonuclease, properties and function 
of, 478 

synthesis of, 474 
Ribonucleic acid (RNA), 484 

bases in, 485 

classes of, 485 

differences from DNA, 489 

D-ribose in, 400 
Ribonucleosides, as glycosides, 407 
D-Ribose, in NAD®, 506 

in RNA, 400, 484 

structure of, 400-401 

in vitamin B 12 400 
Ribosomal RNA, 485 
Ribosomes, 484 
RNA (see Ribonucleic acid) 
Robinson, R., 683 
Rosenmund reduction, 279-280 
Ruff degradation, 416 
Ruzicka, L., 769, 779 



S N reactions (see Nucleophilic dis- 
placement reactions) 
S— S linkages, in proteins, 457^158 
Saccharides, classification of, 400-402 
Salicylaldehyde, acid dissociation con- 
stant, 629 

coumarin from, 687 

uv spectrum of, 629 
Salicylic acid, acetyl derivative of, 658 

formation of, 633 

methyl ester of, 658 

physical properties of, 633 



Salmonella bacteria, 413 
Sandmeyer reaction, 591, 618-619 
Sanger, F., 475 
Santonin, 780 
Saponification, 345 
Schiemann reaction, 591, 619 
Schiff 's base, formation of, 289 
Schmidt reaction, 430 
Selection rules, spectroscopic, 698 
^-Selinene, 780 
Semicarbazide, 289 
Semicarbazones, formation of, 289 
Semiquinone radicals, 638-639 

epr spectra of, 661 
Sensitizers (see Photosensitizers) 
Serine, in ester hydrolysis, 505 

in hydrolytic enzymes, 500 

physical properties of, 459 
Serotonin (see 5-Hydroxytryptamine) 
Sesquiterpenes, 779-780 
Sex attractants, farnesol, 780 
Sex hormones, 788-789 
Silane, physical properties of, 535 
Silanes, examples of, 532 

physical properties of, 535 
Silanols, examples of, 532 

preparation and properties of, 
534-535 
Silica gel, in alcohol dehydration, 257 

in thin-layer chromatography, 
467 
Silicon compounds, bond strengths of, 
532 

organic compounds of, 531-536 

resolution of, 533 

tetrahedral structure of, 533 
Silicones, 536 
Silk, 756-757 
Siloxanes, examples of, 532 

preparation and properties of, 
534-535 
Silver, as catalyst in ethene oxidation, 

261 
Silver bromide, in photography, 639 
Silver ion, complexes with alkenes, 235 
Singlet state, nature of, 697 

of oxygen, 705 
Sitosterol, 783 
Skew-boat (see Twist-boat) 
Smog (see Pollution, atmospheric), 58 
Soap, cleansing action of, 332 

formation of, 330 



general index 845 



Sodium bisulfite, with aldehydes, 294, 

527 
Sodium borohydride, decomposition 
of, 291 

diborane from, 97 

with formaldehyde, 291 

with ketones, 291 

in reduction of metal salts, 36 
Sodium chloride, bonding in, 5 

in infrared spectroscopy, 161 

physical properties of, 5 
Sodium dichromate, in oxidation of 

alcohols, 259 
Sodium ethoxide, formation of, 251 
Sodium fluoroacetate, toxicity of, 226 
Sodium phenoxide, with carbon diox- 
ide, 633 
Sodium phenyl carbonate, rearrange- 
ment of, 633 
Sodium sulfide, in Kraft process, 41 1 
Sodium sulfonates, as detergents, 526 
Sodium triphenylmethide, formation 

of, 654 
Soluble RNA, 485 

Solvents, polar aprotic in S N reactions, 
205 

for S N reactions, 194-195 
Sorbitol, from glucose, 404 
Specific rotation, definition of, 371 
Spectroscopy, 159-178 

electronic absorption, 165-168 

elucidating structures with, 4 

infrared, 161-165 

methods in structure determina- 
tion, 772-778 

microwave, 160 

nuclear magnetic resonance, 168— 
178 

X-ray diffraction, 159 
Spiranes, examples of, 380 

optical isomerism of, 379 
Spiro[2.2]pentane, structure of, 380 
Squalene, 781 

biogenesis of, 792-795 

conversion to lanosterol, 792-795 

epoxide, 794 
Stabilization energy (see also Dereal- 
ization energy a/w/Resonance energy) 

of benzene, 128-129,549 

of iJ-benzoquinone, 636 

of 1,3-butadiene, 135 

of heterocyclic compounds, 673 



Staggered conformations (see Con- 
formations, staggered) 
Starch, heat of combustion of, 510 
hydrolysis of, 408 
structure of, 412 
Stearic acid, 330 

acid dissociation constant of, 331 
physical properties of, 331 
Stereochemistry {see also individual 
types such as Optical isomerism and 
Geometrical isomerism) 
of allenes, 379-380 
of amines, 425-426 
of biphenyls, 380-381 
definition of, 369 
of electrophilic addition to al- 

kenes, 89 
geometric isomerism or cis- trans, 

69-71, 84 
of nucleophilic displacement re- 
actions, 200-201, 389 
and optical isomerism, 369-392 
of organophosphorus compounds, 

525 
of organosilicon compounds, 533 
of spiranes, 379-380 
Stereoisomers, definition of, 39 
Stereospecificity, of biochemical re- 
actions, 387 
Steric acceleration, in S N reactions, 

202 
Steric hindrance, in acetal formation, 
287 
in additions to a,/3-unsaturated 

carbonyl compounds, 311 
in alkylation of phenols, 631 
in biphenyls, 381 
in boat cyclohexane, 65 
in carbonyl additions, 281-282 
in ds-alkenes, 85 
in hemiacetal formation, 287 
relief of, 10 
Steroids, 782-789 

biogenesis of, 789-794 
representative types, 788-791 
Stigmasterol, 790 

Stilbene, photoisomerization of cis- 
trans isomers of, 707 
uv spectrum of, 556 
Stoll, synthesis of civetone, 772 
Strecker synthesis, 458 
Streptomycin, 402 



general index 846 



Structural isomers, 83, 369 

Structure, definition of, 21, 84 

Structure determination, 156-178 

Strychnine, 683 

as resolving agent, 385 

Styrene, 128, 550 

polymerization of, 754-755 
polymers from, 746 
uv spectrum of, 556 

Suberic acid, from civetone, 770 

Substituent (a) constants, of Hammett 
equation, 662-663 

Substituent effects, in base strengths 
of amines, 616 
in electrophilic aromatic substi- 
tution, 567-574 
in light absorption, 701-702 
in nucleophilic aromatic substitu- 
tion, 594 

Substitution reactions (see also Nucle- 
ophilic displacement reactions, Nu- 
cleophilic aromatic substitution, 
etc.) 
electrophilic aromatic, 559-577 
of 2-halo acids, 343-344 
radical mechanisms of 28-32 
S N 1 with aryl compounds, 617 
of saturated hydrocarbons, 26 

Succinic acid, acid dissociation con- 
stant of, 357 
anhydride from, 358, 715 
catalyst in aldol addition, 309 
formation of, 358, 715 
physical properties of, 357 

Succinimide, 439 

Sucrose, structure of, 408-409 

Sugars (see also Monosaccharides, Di- 
saccharides, Carbohydrates, etc.), 4 

Sulfa drugs, 527 

Sulfadiazine, 527, 685 

Sulfaguanidine, 527 

Sulfate esters, 528 

Sulfenic acids, and derivatives, 526 

Sulfinic acids, and derivatives, 526 

Sulfonation, of arenes, 566 

of heterocyclic compounds, 674- 

677 
of naphthalene, 575-576 
of phenanthrene, 577 

Sulfones, 525-526 

from thioethers, 524-526 

Sulfonic acids, acid strengths of, 527 



and derivatives, 526 

preparation of, 527 
Sulfonium salts, formation of, 524 

resolution of, 524-525 
Sulfonyl chlorides, 527 
Sulfoxides, 525-526 

from thioethers, 524-526 
Sulfur, in catalyst poisoning, 279 

electronegativity of, 522 

elemental, 519 

organic compounds of, 520-528 
Sulfur hexafluoride, 517 
Sulfur-sulfur bonds, in proteins, 457- 

458 
Sulfur tetrafluoride, with cyclopenta- 

none, 290 
Sulfur trioxide, in aromatic sulfona- 
tion, 567 
Sulfuric acid, in ester formation, 253- 
254 

esters of, 256-258 
Super acids, 1 3 

and carbonium salts, 200 
Synthesis, of cyclic ketones, 771 

of organic compounds, general 
considerations, 117-120 

overall yields in, 471 

of peptides, 470-474 

prebiotic, 486 
Synthetic fibers (see also Polymers, 
types of), 4 

2,4,5-T, ecological effect of, 598 
Tartaric acid, absolute configuration 
of, 382 
optical isomers of, 377-378 
physical properties of, 374 
resolution of, 385 
as resolving agent, 385 
from threose, 379 
meso-Tartaric acid, from ( — )-erythrose, 

379 
Taurine, 784 

Tautomer, definition of, 315 ' 
Tautomers, of aniline, 427, 605 
of barbaralane, 724 
of barbituric acid, 723 
of bullvalene, 725 
of 1,3,5-cyclooctatriene, 723 
of ethyl acetoacetate, 351-352 
of hydroxypyridine, 503, 723 
of hydroxypyrimidines, 480-481 



general index 847 



of 2,4-pentanedione, 723 
of phenol, 605 
valence bond, 723 
of vinyl alcohols, 605 
of vinylamines, 605 
Teflon, 743, 745 

properties of, 225 
from tetrafluoroethene, 225 
Terephthalic esters, polymers from, 

747 
Terpenes, 778-782 

biogenesis of, 789-794 
in wood, 411 
Testosterone, 791 
Tetraborane, 537 

1 , 1 ,2,2-Tetrabromo-4,4-dimethylpen- 
tane, from 4,4-dimethyl-l-pentyne, 
113 
1,2,5,6-Tetrabromohexane, from 1,5- 

hexadiene, 127 
Tetra-r-butylmethane, 1 17 
2,3,7,8-Tetrachlorobenzodioxin, tera- 
togenic agent, 598 
Tetrachloromethane (see also Carbon 

tetrachloride), 28 
Tetrachlorosilane, physical properties 

of, 535 
Tetracyanoethene, charge-transfer 
complexes of, 613 
as a dienophile, 719 
Tetracyanomethane, 117 
Tetraethyllead, from ethyl chloride, 234 
in gasoline, 58, 234 
physical properties of, 234 
Tetrafluoroethene, from chloroform, 
225 

dimerization of, 726 
polymer from, 745 
polymerization of, 225 
Tetrafluoromethane, bond angles in, 8 

bonding in, 6 
Tetrahedrane, 117 

Tetrahydrofuran, peroxides from, 265 
properties of, 266 
physical properties of, 263 
solubility in water, 264 
as a solvent, 264 
Tetralin, 551 
Tetramethylene glycol (see 1,4-Butane- 

diol) 
Tetramethylene oxide (see Tetrahy- 
drofuran) 



Tetramethylene sulfone, dielectric con- 
stant of, 205 
Tetramethylefhylene, 81 

formation of, 258 
2,2,5,5-Tetramethyl-cw-3-hexene (see 

cw-.syw-Di-r-butylethylene) 
Tetramethyllead, physical properties 

of, 14 
2,2,4,4-Tetramethyl-3-pentanol, 1 89 
Tetramethylsilane, in nmr spectros- 
copy, 170 
physical properties of, 535 
Tetranitromethane, 442 
N,2,4,6-Tetranitro-N-methylaniline, 

611-612 
Tetraphenylmethane, physical proper- 
ties of, 535 
Tetraphenylsilane, physical properties 

of, 535 
1 ,2,4,6- Tetraphenylthiabenzene, 520 
Tetraterpenes, 782 

Tetrazotization, of diamines, 618 
Tetryl (see N,2,4,6-Tetranitro-N-meth- 

ylaniline) 
Thalidomide, 440 

Thermodynamics, equilibrium con- 
stants, 26-28, 93 
free energy of reaction, 26-28 
entropy of reaction, 26-28 
heat of reaction, 26-28 
Thiamine, 685 

pyrophosphate, 685 
Thiazole, 679 
Thietane, 520 
Thin-layer chromatography, in amino 

acid analysis, 467 
Thiobenzophenone, 521 
Thioesters, from thiols, 523 
Thioethers, nucleophilic properties of, 
524 
preparation of, 524-525 
from thiols, 523 
Thiols, 521-524 

acid strength of, 522 
hydrogen bonding in, 522 
infrared spectra of, 522 
mercury salts of, 522 
from organomagnesium com- 
pounds, 230-231 
oxidation of, 523-524, 527 
in paper manufacture, 522 
in petroleum, 522 



general index 848 



preparation, 522 
Thionyl chloride, with carboxylic 
acids, 340 
in preparation of alkyl chlorides, 
255-256 
Thiophene, aromatic character of, 672- 
673 
chemical properties of, 673-679 
physical properties of, 671 
Thiophenol, 520 

Thorium oxide, in ketone synthesis, 772 
Threonine, configuration of, 384 

physical properties of, 459 
Threose, oxidation of, 379 

projection formulas of, 378 
jS-Thujaplicin, 641 
Thymine, in DNA, 480 
Thyroxine, physical properties of, 461 
Titanocene, with nitrogen, 511 
TMS {see Tetramethylsilane) 
Tobacco mosaic virus, properties and 

function of, 478 
Tollen's reagent, 404 
Toluene, 550 

infrared spectrum of, 554-555 
iodination of, 563 
nitration of, 562, 568, 606 
physical properties of, 553 
radical chlorination of, 650 
sulfonation of, 566 
p-Toluenesulfonic acid, formation of, 

566 
/>-Toluenesulfonyl chloride, in sulfon- 
ate ester synthesis, 527 
/i-Toluic acid, preparation of, 650 
o-Toluidine, diazotization of, 591 
j?-Toluidine, basicity of, 614 
nitration of, 608 
uv spectrum of, 614 
/)-Tolyl methyl ketone, oxidation of, 

650 
Transfer RNA, 485 
Transition states, in electrophilic aro- 
matic substitution, 569-571 
in S N reactions, 198-200 
Triacontane, isomers of, 48 

physical properties of, 55 
2,4,6-Tribromoaniline, formation of, 

616 
2,3,3-Tribromo-2-methylpropanoic 

acid, 191 
2,4, 6-Tribromophenol formation of, 633 



Tri-«-butylamine, base dissociation 
constant of, 423 
physical properties of, 423 

Trichloroacetaldehyde, 282 

Trichloroacetic acid, acid dissociation 
constant of, 331, 338 
decarboxylation of, 341 
physical properties of, 331 

Trichloroethane, physical properties 
of, 535 

Trichloroethene, 81 

from ethene and ethyne, 223 

Trichloromethane {see Chloroform) 

Trichlorosilane, physical properties of, 
535 

Triclene, 223 

Tricresyl phosphate, as plasticizer, 748 

Triethyl phosphate, 530 

Triethylamine, base dissociation con- 
stant of, 423 
physical properties of, 423-424 

Triethylborane, from ethene, 96 

Triethylene glycol, 264 

Trifiuoroacetic acid, acid strength of, 
331, 337 
physical properties of, 331 

Trifluoromethylbenzene, nitration of, 
568 

Triglyme, 264 

2,3,4-Trihydroxybutanal, disastereo- 
mers of, 378 

Triisopropylcarbinol, from isopropyl- 
lithium, 234 

Triisopropylmethane, 51 

Trimethylaluminum, 228 

Trimethylamine, physical properties 
of, 12 

Trimethylamine borane, 539 

1 ,3,5-Trimethylbenzene, charge-trans- 
fer complexes of, 613 
physical properties of, 553 

Trimethylborane, 536 

2,2,3-Trimethylbutane, naming of, 52 

Trimethylene oxide, 266 

with organomagnesium com- 
pounds, 234 

Trimethylenemethyl, 77-electron sys- 
tems of, 141-143 
resonance structures of, 142 

2,2,5-Trimethyl-3-hexyne, naming of, 
111 

2,2,4-Trimethylpentane, heat of 



general index 849 



combustion of, 57 
in gasoline, 57 
octane rating of, 58 
Trimethylphosphine, basic properties 

of, 530 
Trimethylsilanol, physical properties 

of, 535 
Trimethylsulfonium bromide, 521 
1,3,5-Trinitrobenzene, charge-transfer 
complexes of, 613 
preparation of, 606-607 
1,3,5-Trinitrobenzoic acid, decarboxyl- 
ation of, 606-607 
2,4,6-Trinitrophenol (see Picric acid) 
2,4,6-Trinitrotoluene (TNT), 442, 611- 
612 
formation of, 562 
Trioxymethylene, 288 
Triphenylamine, uv spectrum of, 614 
Triphenylborane, 539 
Triphenylcarbinol, with sulfuric acid, 

654 
Triphenylmethane, 191 

with sodium amide, 654 
Triphenylmethyl chloride, physical 

properties of, 653 
Triphenylmethyl peroxide, 655 
Triphenylmethyl radical, coupling of, 
655 
epr spectrum of, 661 
resonance hybrid of, 655 
Triphosphoric acid, derivatives of, 528 
Triplet state, of benzophenone, 705 
of oxygen, 705 
nature of, 697 
Tri-w-propylborane, from propene, 

97 
Triterpenes, 781-782 
Trityl chloride (see Triphenylmethyl 

chloride) 
Tropane alkaloids, 684 
Tropilidene (see 1,3,5-Cycloheptatri- 

ene) 
Tropolone, acid dissociation constant 
of, 641 
hydrogen bonding in, 642 
preparation of, 641 
resonance hybrid of, 642 
Tropolones, 641-642 

complexes with ferric chloride, 642 
electrophilic substitution reac- 
tions of, 642 



Tropylium cation, formation of, 642 

hybrid structure of, 642 
Tryptophan, 682 

physical properties of, 461 
Tschitchibabin reaction, 678 
Twist-boat form, of cyclohexane, 65 

of c«-l,4-di-?-butylcyclohexane, 
72 
Tyrian purple, 760 
Tyrosine, physical properties of, 461 



Ultraviolet light, absorption of, 160 

Ultraviolet spectra (see Electronic 
absorption spectra) 

Undecane, physical properties of, 55 

<x,/3-Unsaturated carbonyl compounds 
(see also Carboxylic acids, unsat- 
urated, Aldehydes, a^-unsaturated, 
etc.), conjugate addition to, 311, 355 

Unsaturated compounds (see Alkenes, 
Alkynes, Aldehydes, Carboxylic 
acids, etc.) 

Uracil, in RNA, 485 

Urea, 3 

barbituric acid from, 686 
heat of combustion of, 511 
hydrolysis of, 504 

Urease, 504 



Valence tautomerization, 723 
n- Valeric acid (see Pentanoic acid) 
y-Valerolactone, formation of, 355 
Valine, physical properties of, 459 
van der Waals forces, 36, 39, 475, 739 

and physical properties, 10 
Van Slyke aminonitrogen determina- 
tion, 463 
Vanillin, structure, occurrence and 

synthesis, 658 
van't Hoff, J. H., 3 

and tetrahedral carbon, 386 
Vaseline, from petroleum, 58 
Veronal, 685 

Vibrational and rotational states, 695 
Vinyl acetate, polymer from 746 
Vinyl alcohol, acetaldehyde from, 262 

polymer from, 746 
Vinyl alcohols, derivatives of, in poly- 
merization, 262 
tautomers of, 605 



general index 850 



Vinyl amines, resonance in, 140 

tautomers of, 605 
Vinyl bromide, 589 
Vinyl butyral, polymer from, 746 
Vinyl chloride, from ethene, 220 

from ethyne, 220 

polymers from, 745 
Vinyl ethers, bromine addition to, 139 

resonance structures in, 139 
Vinyl fluoride, from ethyne, 114 

polymer from, 745 
Vinyl halides, preparation of, 219-220 

resonance structures in, 139 

in S N reactions, 220 
Vinyl polymerization, 753 
Vinylacetic acid (see 3-Butenoic acid) 
Vinylacetylene (see Butenyne) 
Vinylboranes, alkenes from, 115 

from alkynes, 115 
Vinylidene halides, polymers from, 743 
Viscose rayon, 410 
Visible light, absorption of, 160 
Vitamin A, 781 
Vitamin Bi (see Thiamine) 
Vitamin B 6 (see Pyridoxine) 
Vitamin B 12 , D-ribose in, 400 

structure of, 681-682 
Vitamin C (see Ascorbic acid) 
Vitamin D 2 , 790 

from ergosterol, 789 
Vitamin K 1; 640 

phytol in, 781 
Vitamins, 4 
Viton, 225 
Vulcanization, 743 



Waxes, from petroleum, 58 
Whiskey, proof of, 245 
Whisky (see Whiskey) 
Williamson synthesis, 252, 263, 267 
Willstatter, R., 681 
Wittig reaction, 531 
Wohl degradation, 417 
Wohler, A., 3 

Wolff-Kishner reduction, 292 
Wood, cellulose in, 410 
composition of, 411 
Wood alcohol (see Methanol), 245 
Woodward, R. B., 681, 683, 726 
Wool, 757 
Wurtz coupling, 654 



X-ray diffraction, of boron hydrides, 
538 

of cholesterol, 787 

of glucose, 405 

of insulin, 476 

of naphthalene, 574 

of penicillin G, 468 

and structure determination, 4, 
159 

of vitamin D 2 , 789 
m-Xylene, infrared spectrum, 554-555 

physical properties of, 553 
o-Xylene, infrared spectrum, 554-555 

physical properties of, 553 
p-Xylene, infrared spectrum of, 554- 
555 

physical properties of, 553 
Xylenes, 550 
D-Xylose, structure of, 400-401 



Walker, D. F., 552 

Wallach, O., 778 

Water, acid dissociation constant of, 

- 13, 122 

base ionization constant of, 1 3 

bond angles of, 7-8 

derivatives of, 9 

dielectric constant of, 204 

dipole moment of, 7 

physical properties of, 10 

shape of, 7 
Watson, J. D., 483 
Wave equations, 130 



Yields, in organic synthesis, 471 
Ylids (see Alkylidene phosphoranes) 



Ziegler, K., 103 

Ziegler catalysts, in polymerization, 
744, 753 

Ziegler process, in coordination poly- 
merization, 103 

Zinc chloride, in alkyl chloride forma- 
tion, 255 

Zwitterions, 462