Skip to main content

Full text of "كتب في علم الميكروبيولوجي"

See other formats



1237_C05.fm Page 197 Wednesday, November 12, 2003 12:54 PM 





~V 




C Challenge of Food and 
the Environment 



Tim Brocklehurst 




CONTENTS 

5.1 Role of Food Heterogeneity 

5.1.1 Aqueous Phase 

5.1.2 Gelled Aqueous Phase 

5.1.3 Oil-in-Water Emulsions 

5.1.4 Water-in-Oil Emulsions 

5.1.5 Gelled Emulsions 

5.1.6 Surfaces 

5.2 Modeling the Food Environment 

5.2.1 Organic Acids 

5.2.2 Dissociation 

5.2.3 Partitioning into Oil Phases 

5.2.4 Water Activity 

5.3 Hurdle Concept 

5.4 Competition with Other Microorganisms 

5.4.1 Interactions Based on the End-Products of Metabolism 
of One Species 

5.4.2 Mixed Culture 

5.5 Adaptation and Injury 

5.5.1 Effects of Environment on Adaptation 

5.5.2 Effects of Sublethal Injury 

5.5.2.1 Enumeration of Sublethally Injured Bacteria 

5.6 Validation in Foods 

5.6.1 Bias and Accuracy 

5.6.2 Validation Using Literature Values 

5.6.3 Validation in Foods 
References 




5.1 ROLE OF FOOD HETEROGENEITY 

Foods are typically not homogeneous. The structure of the food creates local chem- 
ical or physical environments that affect the spatial distribution of microorganisms 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 198 Wednesday, November 12, 2003 12:54 PM 







TABLE 5.1 

Examples of the Heterogeneity of Foods 



Structure of the Food 

Liquid 

Gel 



Oil-in-water emulsion 

Water-in-oil emulsion 
Gelled emulsion 



Examples of Food 

Soups, juices (with some suspended 

material) 
Pate, jellies, skimmed milk cheeses, 

such as cottage cheese 

Dairy cream, milk, salad cream, 

mayonnaise 
Butter, margarine, low fat spread 
Whole milk cheese 



Surface 



Vegetable tissues, meat tissues 



Model Experimental Systems 

Used to Mimic This Food 

Structure 

Broth culture medium 

Cells immobilized in agar or gelatin 
(including in a specifically designed 
Gel Cassette System) 

Alkanexulture medium emulsions 

Culture medium:alkane emulsions 
Alkanexulture medium emulsions, 

where the aqueous phase is gelled 

with agarose 
Agar or gelatin (including a modified 

version of the Gel Cassette System) 




as well as their survival and growth. 197 Microorganisms occupy the aqueous phase 
of foods, and structural features of this phase (Table 5.1) relevant to the length scale 
of microorganisms can influence their growth. The effects of these structural features 
on microbial growth include constraints on the mechanical distribution of water, 77 ' 78 
the redistribution of organic acids, including those used as food preservatives, 3132 
and constraints on the mobility of microorganisms. 30 ' 60 ' 61 ' 109 ' 110 ' 139 ' 152 ' 153 ' 201 

Many foods will contain a number of microstructural features, and the behavior 
of microorganisms is influenced differently in each. For example, Parker et al. 140 
described the effect of microstructure on the distribution and growth of microorgan- 
isms in Serra cheese. Some growth occurred in liquid regions, while other micro- 
organisms formed colonies on surfaces and within the protein gel of the curd (Figure 
5.1). Predictions based on data obtained from broth systems can be applied success- 
fully to organisms growing in structured foods. However, where the structure of the 
food results in a different behavior, this is described below, together with model 
experimental systems for its study. In many cases growth is "fail-safe," in that 
organisms grow more slowly in structured systems than in broths. Wilson et al. 197 
suggested that this may explain the differences that food manufacturers sometimes 
observe, where challenge testing of real foods indicates growth at a slower rate than 
suggested from predictive models. Additionally, the complexity of food structure 
has been identified as a major contribution to the "overall error" included in micro- 
biological modeling predictions. 144 

5.1.1 Aqueous Phase 

Growth in a liquid aqueous phase is typically planktonic, with motility allowing 
taxis to preferred regions of the food. Diffusive transport of nutrients to 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 199 Wednesday, November 12, 2003 12:54 PM 







a) 




b) I 






FIGURE 5.1 Light micrographs demonstrating some structural heterogeneity in hard cheese, 
and showing (a) a colony embedded within the gelled protein of the cheese curd and (b) a 
colony growing on the surface. The black irregular shapes are embedded globules of milk fat. 

microorganisms and of their metabolites away can result in a locally stable equi- 
librium environment until accumulation of microbial biomass and metabolites 
cause bulk chemical changes. This is typically manifested by changes in pH or in 
gaseous composition. When broth culture medium is used in microbiological 
experiments it is this environment that is mimicked, and, with few exceptions, 
models for bacterial growth and death have been developed in such simple broth 
systems. The complexity of foods has been recognized for many years, and it has 
been suggested that the development of detailed models to account for all aspects 
of microbial growth in foods may be too costly, and will not yield useful 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 200 Wednesday, November 12, 2003 12:54 PM 








intermediate models. 14 Simplifying assumptions can be made, and models derived 
in this way have proved useful. 14 

However, with improving knowledge and the advent of mechanistic modeling 
approaches it is possible to make predictions of the behavior of microorganisms in 
structured foods. 

5.1.2 Gelled Aqueous Phase 

In gelled regions microorganisms are immobilized. This can occur as single isolated 
cells, or when these multiply, they are constrained to grow as colonies. 60 ' 61,84 ' 88 ' 140 ' 201 

Model experimental systems for studying colonial growth include agar 17124 and 
gelatin in a specifically designed Gel Cassette System. 28 Immobilized growth as 
colonies results in local depletion of oxygen 200 ' 201 and local accumulation of end- 
products of metabolism, which results in a local decrease in pH within and around 
the colonies. 104192201 Immobilized bacteria also differ from planktonic cultures in 
their susceptibility to antimicrobial compounds, their energy metabolism, and their 
metabolic end-products. 165193 Accordingly, in gelled regions of foods, the growth of 
microorganisms will result in local changes in the concentration of their growth 
requirements and metabolites. This results in growth at a slower rate and to a lower 
yield than planktonic, or free-living cells. 30152 A unifying theory of microbial growth, 
which includes proposed equations for a structured-cell mathematical model, influ- 
ences of local environmental conditions on growth, influences of the microorganisms 
themselves on the environment, transport of solutes between phases, and physical 
expansion of colonies, 152 has been developed to attempt explanation of these growth 
characteristics. 79 Experimental data demonstrate both a decrease in growth rate and 
shrinkage of habitat domain in the case of Listeria monocytogenes, Listeria innocua, 
and Bacillus cereus. In all of these cases, the use of a predictive model based on 
data from the broth experiments would lead to a "fail-safe" prediction in the gelled 
system. However, Wilson et al. 197 described the growth of Staphylococcus aureus as 
a function of sucrose concentration. In the absence of sucrose, growth was slower 
than in the broth cultures when the cells were immobilized in gel. However, as the 
concentration of sucrose was increased, the growth rate in broth decreased, but 
remained unaffected in gel. Hence, these authors identified conditions of a concen- 
tration of sucrose above ca. 15% (w/v) at pH 6 where growth was faster in the case 
of cells immobilized in gel than for cells in broth (i.e., "fail-dangerous" if a model 
prediction was based on data from broth cultures). 

Growth of cells immobilized in gelatin has been examined under nonisothermal 
conditions. 28 This study showed that immobilized cells differ from planktonic bac- 
teria during temperature cycling when stressed by high salt or low pH. A finite- 
difference scheme has been used to combine thermal inactivation modeling with 
thermal conduction modeling to simulate inactivation of bacteria immobilized within 
agar blocks. 17 

The local accumulation of metabolic end-products within and around colonies 
can result in interaction between them. Such competition resulting from close spatial 
distribution has been termed propinquity, and occurs up to a separation distance of 
between 1400 and 2000 |im. 177 ' 201 The authors of these works go on to emphasize 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 201 Wednesday, November 12, 2003 12:54 PM 








that a gap exists between model systems and food, and that to bridge it requires the 
combined efforts of food microbiologists and microbial physiologists. 201 

5.1.3 Oil-in-Water Emulsions 

Here, structure is affected by the concentration and form of the oil phase. The 
concentration of oil in food varies considerably, 32 and in milk is typically between 
3 and 5% (v/v), but in mayonnaise may be between 26 and 85% (v/v). The oil phase 
exists as polydispersed droplets with a mean diameter that is typically between 0.15 
and 8 |i.m. In concentrated emulsions, the space of the interstices between the droplets 
is of the same order of size, which is also the same order of size as many bacteria. 
In model experimental systems a relationship exists between the concentration 
of oil and the form of growth of microorganisms. 139 Where the concentration of lipid 
phase was low (30% v/v) the growth of bacteria was as free-living (or planktonic) 
cells. An increase in the concentration of the oil phase had no effect on the form of 
growth of bacteria until it was increased to 83% (v/v). Here the bacteria became 
immobilized between the close-packed oil droplets. This entrapment resulted in 
growth not as planktonic cells, but as discrete colonies. The droplets within emul- 
sions confer opacity, and hence visualization of microorganisms is difficult. A mix- 
ture of chloroform and methanol was used to selectively remove the oil phase and 
allow the examination of colonies in situ. 30 - 139 The investigators showed that the 
colonies are formed from a single bacterium, and as they expanded they displaced 
the emulsion droplets. Immobilization of bacteria by the lipid component and sub- 
sequent growth as colonies resulted in a decreased rate of growth and a shrinkage 
of the habitat domain compared with growth as planktonic cells — essentially, 
similar results to the consequences of colonial growth in gels. 

5.1.4 Water-in-Oil Emulsions 

These consist of an internal aqueous phase dispersed as discrete spherical or irreg- 
ularly shaped droplets within an outer oil phase, which may contain a mixture of 
fluid and crystalline fats. In the case of margarines the droplets of aqueous phase 
are typically irregular in shape, and can range between 0.3 and 30 \i in diameter. 186 

Droplets can be contaminated with microorganisms at the point of emulsion 
manufacture. 186 The proportion of droplets occupied by microorganisms is small, 
and a model to predict microbiological contamination based on a function of the 
initial contamination, and the numbers of droplets exceeding the minimum size for 
occupancy, has been developed. 186 

Classical theories to describe microbial growth rely on the maintenance of 
discrete compartmentalized droplets that restrict the availability of water, space, and 
nutrients for growth. On the basis of these assumptions, Verrips and Zaalberg 186 and 
Verrips et al. 187 used a mechanistic approach to predict the growth of bacteria within 
discrete droplets related to the dimensions of the occupied droplets. This was 
expanded further by modeling the energy demands of the contained bacteria. 175 
Models are useful here to predict states that are difficult to measure, and predictions 
confirm that bacteria in the droplets can grow well, but that their numbers remain 



2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 202 Wednesday, November 12, 2003 12:54 PM 








small when expressed per unit volume of emulsion (although their local number 
density within a droplet is extremely high). Additionally, microorganisms cease to 
grow when the concentration of metabolic end-products (typically organic acids) 
becomes toxic or if a requirement for growth, such as oxygen or a carbon source, 
is exhausted. Models confirm that bacterial growth is restricted when the food 
structure remains intact (i.e., when coalescence of the droplets does not occur). This 
was observed in model experimental systems where an increase in numbers of 
bacteria in water-in-oil emulsions was always accompanied by coalescence of the 
droplets of aqueous phase. 31 

5.1.5 Gelled Emulsions 

Many food emulsions are gelled. This can occur by the deliberate addition of gums 
or thickeners to increase the bulk viscosity (such as in sausages) or the denaturation 
of protein to form protein micelles (such as in cheese). Microorganisms are immo- 
bilized and constrained to form colonies much as in gelled systems described 

above. 60 ' 61 ' 140 

5.1.6 Surfaces 

The simplest form of food structure is the surface. Growth of bacteria on the surface 
of food has been measured on Canadian wieners, 118 pate, 69 and vegetable tissues. 27 
Model experimental systems are numerous and include agar gels, 53 ' 115 ' 168 ' 179 ' 199 ' 202 
agar film, 115 two-dimensional gradient plates, 178-180 ' 203 ' 204 and a modification of the 
Gel Cassette mentioned above. 29 

Nicolai et al. 132 modeled surface growth with the assumption that it was in a 
surface film of liquid. However, growth on a surface is typically colonial. Hence, 
constraints on growth are similar to those described in the case of gels. Some key 
differences are important in modeling. Crucially, diffusion limitations are greater at 
a surface than within an enveloping gel. This was confirmed by Wimpenny and 
Coombs, 200 Peters et al., 141 and Robinson et al. 155 who measured the depletion of 
oxygen and accumulation of protons immediately beneath the colony and extending 
into the substratum. Colonial growth on surfaces results in decreased growth rates, 
and comparisons of the growth rates of Salmonella typhimurium affected by increas- 
ing salt or sucrose followed the order: broth > immersed colonies > surface colo- 
nies. 29 This suggests that the rate of growth on surfaces may not be well predicted 
by models derived from broth systems. 29 Spatial distribution on a surface leads to 
interactions between colonies. 176 Spatial and temporal variations have a major influ- 
ence on the potential of surfaces to support bacterial growth. In foods, it is partic- 
ularly the availability of water. 50 Drying of a food may be deliberate to inhibit growth, 
and desiccation of microorganisms has been reviewed. 146 A solid surface model 
system was developed to study the effect of gas atmosphere on growth of several 
psychrotrophic pathogens. 21 This system demonstrated that increased C0 2 markedly 
inhibited the growth of all pathogens. The model system can be applied to exami- 
nation of the growth of pathogens on minimally processed produce under modified 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 203 Wednesday, November 12, 2003 12:54 PM 








atmospheres. Radial growth of colonies of B. cereus on a solid agar surface was 
dependent on interaction between agar concentration and water activity. 168 

5.2 MODELING THE FOOD ENVIRONMENT 

In order to predict the growth of microorganisms in foods reliably, it is vital to use 
the correct initial chemical conditions. The structural heterogeneity of foods results 
in a chemical heterogeneity, which is often complicated by dynamics within the 
food that create a "new" chemical environment. Models of varying complexity exist 
that can predict the true initial chemical state of foods. Microorganisms occupy the 
aqueous phase of foods, 30184 and hence, it is the chemical composition of this phase 
that requires accurate prediction. 

Many foods rely for their preservation on the concentration of organic acids 
(e.g., acetic, lactic, benzoic, or sorbic acid). In addition, the concentration of sugars 
or salts can contribute to preservation. It is, therefore, no surprise that many predic- 
tive models use combinations of pH and water activity (although often expressed as 
concentration of NaCl) together with temperature as the three major determinants 
of growth. What follows is a summary of available models that can predict the initial 
environmental conditions within foods. 

5.2.1 Organic Acids 

Acetic, lactic, benzoic, and sorbic acid (and their salts) are added as preservatives 
in many foods, although acetic and lactic acids are also produced in fermented foods 
as end-products of microbial metabolism. Their preservative action is by virtue of 
a combination of their effect on the pH of the food and the antimicrobial properties 
of the undissociated form of the molecule. Accordingly, their antimicrobial effect is 
influenced by the fundamental thermodynamic characteristics of dissociation and 
partition. It is these that must be modeled to predict the potential of foods to inhibit 
the growth of microorganisms. 

5.2.2 Dissociation 

Weak organic acids dissociate (or separate) into their component parts. In the case 
of acetic acid, this occurs as: 




CH3COOH 


<^ 


CH3COO- 


and 


H + 


acetic acid 




acetate 




hydrogen ion 


(undissociated) 




(dissociated) 




(proton) 



This dissociation is key to prediction of the concentration of the undissociated form 
of the acid, which has the predominant antimicrobial effect in foods. 11 ' 65 ' 166 

The Henderson-Hasselbalch equation relates the pH of the food to the pK a and 
the relative proportions of dissociated and undissociated acid in foods have been 
predicted 198 as follows: 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 204 Wednesday, November 12, 2003 12:54 PM 








[acid].. . t , 

pH = pK + \Og w . 'dissociated ^ ^ 

\acid\ ,. . t , 

L ■* undissociated 

Rearrangement gives the concentration of weak acid in its undissociated (i.e., micro- 
biologically active) form, [HA] aq , given the pH, ipK a , and total concentration of weak 
acid, [HA] T , as follows: 

[HA] = [HA } T K : (5.2) 

q 1 + 10 a 

where [HA] aq is the concentration of undissociated organic acid in the aqueous phase 
and [HA] T is the total concentration of organic acid. pK a is the negative logarithm 
of the dissociation constant K a , which is a thermodynamic constant controlling the 
dissociation equilibrium shown above: 

p£=-log(£ a ) (5.3) 

K a is typically a small number, and published values are available. 205 ])K a varies 
slightly with temperature, and an empirical equation that predicts this variation has 
been published 154 : 

V K a =(^)-B + (CT) (5.4) 

where T is the temperature in Kelvin (K), and A, B, and C are shown in Table 5.2. 
Such predictions are important preliminaries in dealing with the challenge of food 
and the environment. Without such knowledge it is quite simple to apply an incorrect 
initial environmental condition when using predictive microbiology tools, and this 
can easily result in erroneous predictions. 

Predictions must also be reiterative. For example, once dissolved, the organic 
acid will dissociate depending upon local pH, but will then perturb this pH. The 
dissociation is also dependent on local buffering capacity of the food, and this is 




TABLE 5.2 

Values of A, B, and C to Be Inserted into 
Equation 5.4 for Calculation of the Effect 
of Temperature on p^ a 154 

Acid ABC 



Acetic 


1170.48 


3.1649 


0.013399 


Lactic 


1286.49 


4.8607 


0.014776 


Benzoic 


1590.2 


6.394 


0.01765 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 205 Wednesday, November 12, 2003 12:54 PM 








extremely difficult to predict. However, Wilson et al. 198 developed a method for 
performing calculations describing the reiterative dissociation of organic acids, and 
hence predicting the true chemical composition of foods. This not only allows 
microbial growth models to predict growth, but also allows the changes in pH caused 
by microbial metabolism to be predicted. These authors used a theory describing 
the behavior of weakly dissociating systems, and knowledge of dissociation con- 
stants and concentrations. They make the point that food is too complex for solutions 
to be achieved through complex calculation. Hence, the authors characterized the 
buffering behavior of food by a titration with a strong (i.e., completely dissociating) 
acid, and then used knowledge of the dissociation constants of weak acid preserva- 
tives to predict the behavior of these in the food. Their calculation scheme may also 
be applied to a mixture of weak acids including polyacid species such as the 
tricarboxylic acids (e.g., citric acid). 198 

5.2.3 Partitioning into Oil Phases 

In biphasic foods, which contain aqueous and lipid phases, the antimicrobial undis- 
sociated acids partition between the aqueous and lipid components. 32 This decreases 
the concentration of undissociated acid in the aqueous phase. Partition coefficients 
of acetic, lactic, and sorbic acids between sunflower oil and water have been reported 
as 0.02, 0.033, and 2.15, respectively, 32 demonstrating the potential for, particularly, 
the undissociated form of sorbic acid to decrease in the oil phase of biphasic foods. 

As a complication, the pH of foods preserved using organic acids is typically 
in a region where weak organic acids are present in both the undissociated and the 
dissociated forms. Calculation of the residual concentration of the undissociated 
form following partition is thus difficult because the concentration is subject to the 
effects of partition, and to the dissociation equilibrium based on the new pH of the 
system and the new residual concentration of undissociated acid. 

A modified form of the Henderson-Hasselbalch equation has been developed, 198 
which takes these effects into account and gives the proportion of the total weak 
acid in a two-phase system that is present in its undissociated form in the aqueous 
phase, given the pH, the volume fraction of oil, and the partition coefficient for the 
undissociated weak acid. It was cast as: 




[HA] an 1_ 

4> 



= - — (5.5) 

[HA] T 



l + K< 



vl"*y 



+ 10 



(pH-ptf a ) 



where K P is the partition coefficient and <|) is the fraction volume of the oil phase. 
Predictions have been validated in aqueous and biphasic foods. 198 

5.2.4 Water Activity 

Water activity (a w ) is a measure of the concentration of available water in a food 
and can be defined as the tendency of water to escape from a solution relative to its 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 206 Wednesday, November 12, 2003 12:54 PM 








ability to escape from pure water at a specific temperature. Water activity is equal 
to the equilibrium relative humidity divided by 100. Pure water has an a w of 1.000, 
and an environment where water is absent has an a w of 0.000. 49 ' 182 Most microor- 
ganisms require a high a w for growth, and a w is included in many predictive micro- 
biology models. The a w of foods can be adjusted by the addition of solutes (humec- 
tants), such as sodium chloride, sucrose, or glycerol. In some cases, the solute itself 
may have toxic effects, and the inhibition of growth of microorganisms when sodium 
chloride is used to adjust a w can be greater than when glycerol is used, due to the 
toxicity of high concentrations of sodium chloride. 1275182 Care must be taken, there- 
fore, to use only those predictive models that use the same humectant as the food 
of interest. Prediction of the initial a w of the food can be achieved from first principles 
using a variety of equations, such as Raoult's law, 4993 which was derived by 
Christian 49 as: 



t -vmty 

Log e «. = — (5.6) 

where m is the molal concentration of the solute, v is the number of ions generated 
by each molecule of the solute, and <|) is the molal osmotic coefficient. Commercial 
software to predict water activity from a list of food ingredients in a recipe is available 
(e.g., ERH CALC™). 



5.3 HURDLE CONCEPT 

Hurdle technology involves the use of combinations of physical or physicochemical 
preservation techniques at subinhibitory levels to control the growth of food-borne 
microorganisms. 98 This has the effect of conferring microbial safety and stability 
while maintaining acceptable nutritional and sensorial attributes, 160 an approach that 
is important for minimally processed extended shelf life foods. 108 With the develop- 
ment of new food products that depend on multiple barriers to ensure safety, it 
becomes necessary to develop the means to apply predictive microbiology to hurdle 
technology. 43 ' 97 Careful definition of the conditions defining the boundaries of growth 
or survival will allow industry to design foods with the appropriate level of 
safety; 149 ' 160 however, there have been few attempts to provide a quantitative assess- 
ment of hurdles. 160 

Examples of interactions include C0 2 , pH, and NaCl on L. monocytogenes', 71 
temperature, pH, citric acid, and NaCl in reduced calorie mayonnaise on Salmonella 
spp.; 121 pH, acid, and salt on Staphylococcus aureus; 64 salt, pH, and nitrite on 
Escherichia coli 0157:H7 in pepperoni; 151 temperature and pH on E. coli 0157:H7 
in Lebanon bologna; 66 and nisin and leucosin on L. monocytogenes . 138 

While it is clear that combinations of hurdles can influence food-borne micro- 
organisms, it is not clear to what extent these factors interact. When the square root 
model is used to describe the effect of several hurdles such as temperature, pH, and 
a w , these factors are usually considered to act independently, with no interactions. 119 
Ratkowsky and Ross 149 described a combined probability /kinetic model for Shigella 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 207 Wednesday, November 12, 2003 12:54 PM 








flexneri in which temperature, pH, a w , and nitrite were shown to act individually. It 
would be expected, however, that interactions must occur between certain hurdles. 
For example, interactions between organic acids and pH would be expected (due to 
the influence of pH on the extent of dissociation as described in Section 5.2) and 
have been observed. 38121147 Effects on heat resistance of E. coli due to the interactions 
between combinations of temperature, pH, NaCl, and sodium pyrophosphate have 
been modeled. 86 ' 87 

Polynomial models can be used to describe interactions between a wide variety 
of hurdles. This is because the regression methods used facilitate the search for 
quadratic or interactive effects. Combination effects have also been modeled using 
Belehradek and Arrhenius models. 9091 The growth of L. monocytogenes at 9°C as 
influenced by sodium nitrite, pH, sodium chloride, sodium lactate, and sodium 
acetate has been modeled, 130 and predictions compared with the growth of organisms 
in real sausage and predictions from Food MicroModel. Food MicroModel is a 
software package developed in the U.K. that contains secondary models of the effects 
of environmental factors (mainly pH, concentration of NaCl, and temperature) on 
the survival, growth, and thermal death of major food-borne pathogenic bacteria in 
broth. Predictions were on average within 20% of the Food MicroModel predictions 
based on 10 experiments although predictions of growth in sausage were, on average, 
16% below the observed values based on inoculation of four sausages. This is perhaps 
related to the effects of structure as described in Section 5.1. The effect of previous 
growth temperature, previous cell concentration, and previous pH on the lag time 
and specific growth rate of Salmonella typhimurium. has been investigated using 
response surface models. 135-137 In all cases the previous growth history did not 
influence the predictions of the model. 

Some authors contend that predictive models of the combined effects of tem- 
perature and water activity and the combined effects of temperature and pH suggest 
that the effect of the combinations on growth rate is independent. 120 However, these 
authors go on to state that the factors are interactive at the no-growth interface (i.e., 
the point where growth ceases). Such interface models quantify the probability of 
growth and define conditions at which the growth rate is zero or the lag time is 
infinite. Such new growth interface or habitat domain models have been pub- 
lished. 116 ' 181 Square root models and response surface models were developed to 
look at the effects of interactions between dissolved carbon dioxide and water activity 
on the growth and lag time of Lactobacillus sarcae. 52. The response surface models 
showed the best correlation although at low water activities, predictions were illog- 
ical. Both models, however, proved to be useful in the prediction of the shelf life 
of meat products, and were validated by comparison with an existing model. 196 
Similarly, a quadratic response surface model was built to predict the combined 
effects of temperature and modified gaseous atmosphere on the growth of Yersinia 
enterocolitica. 143 

Predictive models have been used to predict the response of Listeria monocyto- 
genes exposed to acid, alkaline, or osmotic shock at the time of inoculation on the 
subsequent effects of temperature, concentration of NaCl, and pH. 47 The authors 
found that predictive models were unreliable, highlighting potential problems of 
variable conditions, but failing to consider the implications of adaptation of the 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 208 Wednesday, November 12, 2003 12:54 PM 








organisms to osmotic or pH effects. An important development is the use of the 
gamma concept, which assumes that the effects of controlling variables can be 
multiplied and that the cardinal parameters of temperature, pH, and water activity 
are not a function of other variables. 196 Accordingly, these authors developed a model 
based on the prediction of growth rate as a function of temperature and water activity 
and another where growth rate was predicted as a function of temperature and pH. 
The two models were multiplied to produce one overall model, which was validated 
against new experiments. Additive interaction between inhibitors has been 
observed. 24 These authors used a response surface methodology to model the 
response of L. monocytogenes to a bacteriocin (curvaticine) and sodium chloride: 
the model showed that the combination of the two inhibitors was greater than the 
effect of each individually. Interactions between inhibitory compounds were also 
investigated 8 by using a series of secondary models 7 describing independently the 
effects of environmental factors. 8 The authors of the latter work then went on to 
show that, by taking into account interactions between environmental factors, the 
model decreased the frequency of fail-safe growth predictions from 13.5 to 12.1%, 
while the frequency of fail-dangerous no-growth predictions decreased from 16.1 to 
7.1%. These findings suggest that interactions are occurring within the system, and 
that the models were taking them into account. 8 However, even with multiplicative 
models the predictions are less accurate to describe lag time and growth rate near 
the limits of growth of microorganisms, 7 and lag time models were particularly 
vulnerable to error. 

Inactivation modeling is less common in response to a combination of hurdles. 
Death kinetics as a function of pH, storage temperature, and concentration of essen- 
tial oil have been described using a quadratic function, and used to predict success- 
fully the death of Salmonella in home-made salads. 89 A regression model describing 
the heat inactivation of L. monocytogenes was based on the Gompertz Equation. 48 
The equation enabled separate characterization of the parameters of the shoulder, 
the maximum slope, and the tail. Interactive effects were then derived from the 
regression model. This showed that the shoulder region of the survival curve was 
affected by pH, and the maximum slope by temperature, fat content, and interaction 
of temperature and milk fat. Model validation was successful for temperatures only 
above 62°C, however. The combined effects of pH and ethanol on the heat inacti- 
vation of B. cereus, S. typhimurium, and Lactobacillus delbrueckii were modeled 
using a series of second-order polynomial equations to describe variations in D 
values resulting from changes in pH or added ethanol. 45 The heat inactivation of B. 
cereus spores was modeled using a new concept of z- value modeling using a z(pH) 
value, 96 where z(pH) was defined as the difference in pH from a reference pH value 
required to effect a 10-fold reduction in the D value. A linear relationship between 
the calculated z(pH) value and the lowest of the pK values of organic acids used to 
effect heat resistance was found. The heat resistance of Listeria monocytogenes in 
logarithmic phase cells that had been heat shocked at 42°C for 1 h and subcultures 
of cells that were resistant to prolonged heating has been modeled. 9 A better fit for 
the survivor curves was found using sigmoidal equations compared with the classical 
log-linear models. Comparisons between models showed that an increase of thermal 
tolerance was induced by sublethal heat shock or by the selection of the heat-resistant 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 209 Wednesday, November 12, 2003 12:54 PM 








population. Both isothermal and nonisothermal heat inactivation effects on the ger- 
mination and heat resistance of B. cereus spores have been modeled. 70 An inactivation 
model was developed for Salmonella enteritidis. 96 It modeled the response of the 
organism to a range of concentrations of oregano essential oil and temperatures at 
two pH values. Quadratic functions were then used to predict the growth of this 
organism in home-made salads. The inactivation kinetics of E. coll 0157:H7 were 
modeled using the Baranyi model (based on a set of nonautonomous differential 
equations) 13 as a function of time to estimate the kinetic parameters. 164 Quadratic 
models were then developed with natural logarithms taken of the shoulder and death 
rate as a function of temperature, pH, and concentration of oregano essential oil. 
The predicted values from the model were validated using viable count measure- 
ments made within real salads. 

Modeling spore responses (other than inactivation) is unusual. The germination 
kinetics of spores of proteolytic Clostridium botulinum. 56 A as a function of temper- 
ature, pH, and concentration of sodium chloride have been modeled. 46 The germina- 
tion kinetics were collected and expressed as the accumulated fraction of germinated 
spores with time and each environmental condition, and this accumulated fraction 
was then described by an exponential distribution. Quadratic polynomial models were 
developed by regression analysis of the exponential parameter and the extent of 
germination as a function of the variables under study. Validation experiments con- 
firmed that the predictions were acceptable, and in most cases were fail-safe. 

5.4 COMPETITION WITH OTHER MICROORGANISMS 

Existing published models include a wide range of environmental, physical, or 
chemical factors; however, the competitive influence of microorganisms has not yet 
been incorporated into them. Competition may not be an issue in many foods, since 
interactions would not be expected until cell numbers had reached a potential hazard 
or caused spoilage. 160 On the other hand, growth of L. monocytogenes in dairy 
products is influenced by the natural microflora, and interactions may be difficult to 
model. 33 Therefore, it has been suggested that competition must be considered in 
the development of predictive models. 163 

Competition between microorganisms in a solid matrix such as food depends to 
a large extent on proximity of colonies to each other. 201 Cells growing on surfaces 
generate gradients of redox potential, pH, oxygen concentration, and nutrients, which 
can influence the growth of neighboring colonies. This phenomenon can be observed 
in foods, for example, where "nests" of lactic acid bacteria in fermented sausage 
influence the survival of food-borne pathogens, 201 and also in dairy products where 
interactions between natural microflora and L. monocytogenes are influenced by the 
nature of the food matrix. 33 

A related concept is the idea of "maximum carrying capacity" of a food prod- 
uct, 160 in which inhibition of pathogens by other microorganisms takes place when 
the competing flora have reached numbers at which the environment can support 
no further growth. This was observed with cocultures of L. monocytogenes and 
Carnobacterium piscicola. 35 In this study, the maximum population density of L. 
monocytogenes was reduced by the competing lactic acid bacteria, and this was 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 210 Wednesday, November 12, 2003 12:54 PM 








attributed to nutrient depletion. It is by no means clear to what extent competition 
is related to depletion of nutrients. The thermal tolerance of S. typhimurium was 
enhanced by the presence of competing microflora, and it was suggested that the 
presence of competitors may have influenced the pathogen to induce stationary- 
phase gene expression. 62 

The interaction of spoilage microorganisms has recently been quantitated by Pin 
and Baranyi. 142 Polynomial models were developed for a number of microorganisms, 
and the growth of groups of strains was compared individually and in the total 
mixture. This approach allowed the identification of the dominant group on the basis 
of its growth rate and lag time. These authors also showed that reduced growth rate 
could be attributed to microbial interactions. Competition from naturally occurring 
microflora has been documented. 94 Here, predictions of the growth of Pseudomonas 
and Listeria in meat were made. Predictive models worked well in predicting the 
growth of both organisms in decontaminated meat and in decontaminated meat 
inoculated with each organism, together or individually. However, the presence of 
naturally occurring microflora in non-decontaminated meat prevented the initiation 
of growth of Listeria and the predictive models failed. 

A related aspect of interaction is that of the potential for quorum sensing between 
microorganisms. 101 At low inoculation concentrations, modifications to modeling 
approaches were necessary to take into account inoculum size variation. Modeling 
the effects of inoculum size stochastics, however, confirmed that the growth rate 
was independent of inoculum concentration but that variability occurred as the 
inoculum concentration decreased. 209 ' 210 

5.4.1 Interactions Based on the End-Products 
of Metabolism of One Species 

This is a complex modeling task, but stoichiometric modeling can be used to relate 
the end-products of metabolism to the inhibition of the same or an accompanying 
organism. It assumes a "reaction scheme," and seeks to choose the simplest repre- 
sentation of a system that embodies the behavior of interest. 

Thus, a stoichiometric model can predict the local changes in weak acid con- 
centration resulting from microbial growth. This must then be used to predict changes 
in local pH. This can be done by an empirical characterization, merely by using a 
titration of the growth environment with the acid of interest, and fitting a curve to 
these data. Alternatively a quasi-mechanistically based approach may be taken, 132 
or use made of a Buffering theory 198 described in Section 5.2. An advantage of the 
latter is that the model may be easily applied to systems of differing buffering 
capacity, and can combine the effects of mixtures of weak acids. Diffusion is an 
integral part of such modeling, and a standard model of Fickian diffusion using 
published diffusion coefficients in aqueous solution is usually appropriate. 

For growth in liquid systems, a cardinal growth model has been combined with 
cardinal pH data. 99 Cardinal models use the cardinal values (minimum, optimum, 
and maximum values) of the environmental factors that constrain growth. Instanta- 
neous growth rates from this model were used in a modified Baranyi growth model, 13 
together with stoichiometric parameters determined from bioreactor experiments. 197 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 211 Wednesday, November 12, 2003 12:54 PM 








The change in pH from production of lactic acid was determined by use of a 
Buffering theory. 198 Very close agreement was found between the model and the data. 

5.4.2 Mixed Culture 

Application of stoichiometric approaches to mixed cultures also works well. Wilson 
et al. 197 showed the growth of a mixed culture of Lactococcus lactis and Listeria 
innocua in a bioreactor at pH 4.5. Predictions used cardinal model parameters, 99 
and stoichiometric parameters from bioreactor experiments. 197 A Buffering 
theory 198 was used to predict changes in pH. Such an approach provided good 
prediction of both the rate and extent of growth of the two organisms. Of interest 
in these approaches is that a stationary phase was not incorporated into the primary 
growth model, but emerged from the prediction in response to the accumulation 
of metabolites. 

Interactions resulting from the production of antimicrobial bacteriocins by lactic 
acid bacteria in conjunction with the inhibition resulting from production of lactic 
acid have been modeled. 44 These authors used a modification to logistical equations 
that described the combined (although not additive) effects of two or more inhibitory 
compounds. They then applied their findings to the inhibition of Leuconostoc 
mesenteroides. The inhibition of growth of Enterobacter cloacae by Lactobacillus 
curvatus resulted from the production of lactic acid by the latter, and the concomitant 
decrease in pH, 105 which was also inhibitory to L. curvatus. This interaction has been 
modeled using a set of first-order differential equations describing growth, consump- 
tion, and production rates for both microorganisms. 107 Parameters were obtained from 
pure culture studies and from the literature, and the equations were solved using a 
combination of analytical and numerical methods. Predictions of growth of mixed 
cultures used parameters from pure culture experiments, which were close to the 
experimental data. The models also showed that interactions occurred when the 
antagonistic bacterium, in this case L. curvatus, reached 10 8 cfu/ml. 



5.5 ADAPTATION AND INJURY 

5.5.1 Effects of Environment on Adaptation 

Predictive microbiology should deal with bacterial stress within populations. 6 An 
example is the extension of the lag time of Listeria monocytogenes under suboptimal 
conditions when the inoculum was stressed. 6 More important, considerable interest 
has arisen recently in the problems of adaptive responses of bacteria and in the cross- 
resistance that this can confer. For example, adaptation of bacteria to methods of 
preservation can result in survival or growth that is better than predicted if the 
adaptive response is ignored. Accordingly, adaptation of bacteria can lead to unsafe 
or spoiled food. 34 The implications of adaptation can be demonstrated by reference 
to the acid tolerance response (ATR). The ATR in L. monocytogenes has been 
attributed to the de novo synthesis of proteins (sometimes referred to as acid shock 
proteins) when exposed to a decrease in extracellular pH. 134 Such biochemical 
changes confer acid resistance on the organisms, but O'Driscoll et al. also noted 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 212 Wednesday, November 12, 2003 12:54 PM 








that L. monocytogenes that had been induced to show the ATR also had an increased 
resistance to thermal, osmotic, and cold stresses. 134 ATR has been defined as the 
resistance of cells to low pH when they have been grown at moderately low pH or 
when exposed to a low pH for some time, 59 and is typically demonstrated in broth 
culture, where a pH of 4.8 to 5.0 is reported to give an optimum ATR. 56 Many foods 
fall into this region of pH, and, more important, many microorganisms can experi- 
ence this pH transiently during food production or sanitation protocols. Adapted 
populations could then result. 

Additionally, it is clear from the above sections that one of the key effects of 
food structure is the immobilization of microorganisms and their resultant growth 
as colonies. This results in local changes in the concentration of substrates 201 and, 
particularly, a local accumulation of acidic metabolic end-products leading to a 
decline in pH within and around the colony 104 ' 192 with a pH gradient extending into 
the surrounding menstrum. 192 ' 201 In the case of S. typhimurium, the pH gradient 
extended from the original pH 7.0 in the surrounding medium to pH 4.3 inside the 
colony. 192 Such a local decline in pH within the colony is greater than the change 
required to stimulate an ATR in Salmonella and other Gram-negative enteric 
bacteria 95 and in L. monocytogenes . 91 It is conceivable, therefore, that cells of food- 
borne pathogenic bacteria immobilized as colonies embedded in a food matrix may 
undergo a self-induced ATR stimulated by a localized pH that has declined by virtue 
of the colony's own metabolic processes. It is known that acid shock proteins are 
synthesized and exported from cells experiencing adaptation in broths. Should this 
also be the case in colonies, it would result in cells within the colony becoming 
acid tolerant. 

Despite the importance of adaptation in food microbiology, attempts to model 
it are rare. Authors have acknowledged that organisms behaved differently when 
exposed to changes in pH or sodium chloride concentration, and that exposure to 
these agents during exponential phase had a more dramatic effect than during the 
lag phase when adaptation was possibly induced. 47 However, no attempt to incor- 
porate adaptive responses into models was made. A cross-resistance between high 
hydrostatic pressure and mild heat, acidity, oxidants, and osmotic stresses was 
demonstrated for E. coliO\51. 20 Differences were most dramatic in stationary-phase 
cells; the only exception being acid resistance where differences were also apparent 
in the exponential phase, although, again, no attempt to incorporate these into a 
model was made. In one attempt to model adaptation, a model to describe the 
influence of temperature and the duration of preincubation on the lag time of L. 
monocytogenes was developed. 10 

5.5.2 Effects of Sublethal Injury 

Subjection of bacteria to inimical processes can result in the cumulative injury of 
the bacteria, resulting in death. Sublethal injury is the reversible damage inflicted 
on bacteria that is insufficient to cause a loss of viability, and from which the bacteria 
can recover. 580 ' 150 It is an important phenomenon to recognize when collecting data 
for modeling, because bacteria can often fail to form colonies on conventional 
selective microbiological culture medium used for their enumeration. 2127 They can 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 213 Wednesday, November 12, 2003 12:54 PM 








also fail to respond positively to viability stains. 26 However, the cells can remain 
viable and the injury can be repaired in foods, where the bacteria can then increase 
in numbers. 82 ' 206 The severity of treatment that results in sublethal injury differs 
between species, although serotypes of Salmonella have been found to respond 
similarly to one another. 128 

5.5.2.1 Enumeration of Sublethally Injured Bacteria 

A range of methods have been used to determine the extent of injury of microorga- 
nisms. These include differential plate counts on selective and nonselective agars 15 ' 150 
or on minimal and more complex media, 102 extension of the lag phase, 4 ' 102 and 
changes in bioluminescence. 67 Such methods can be used to optimize both the 
recovery medium and the time and temperature of incubation. For example, it has 
been shown that cells of L. monocytogenes that were subjected to sublethal injury 
by heat exhibited a broad optimum temperature for recovery, with an optimum 
between 20 and 25°C, but that incubation at 2 or 5°C failed to allow repair. 103 The 
time taken for repair of injury to complete can be determined by measuring the time 
before equivalent counts are found on a selective medium (which will not support 
the growth of sublethally injured bacteria) and a nonselective culture medium (which 
will allow the growth of sublethally injured bacteria). 103 Some modeling of resusci- 
tation has been published. 117 Predictions of response might be possible: for example, 
a relationship was found between the concentration of sodium chloride in the heating 
menstrum and its concentration in the growth medium used for the resuscitation and 
subsequent enumeration of S. typhimurium. 106 

5.6 VALIDATION IN FOODS 

One of the most important aspects of model development is ensuring that predictions 
made by the model are applicable to real situations. This is the validation process. 
It should involve comparisons of the predictions of the model with observed mea- 
surements, which should be different data to those used to construct the original 
model. Although some predictive models have been constructed in real foods (see 
later in this chapter), the vast majority of models have been constructed from 
experiments performed in laboratory culture media (typically broth). In all cases the 
validation process should, ideally, include comparisons with the behavior of micro- 
organisms in real foods or during real food processes. However, due often to cost 
but also other factors, validation can be done in model systems, or using previously 
published data. A validated model should be consistently "fail-safe," that is, predic- 
tions should fail on the side of safety (i.e., predicted growth rate and lag time should 
be faster and shorter, respectively, than experimental values). Predictive models can 
be crucial aspects of HACCP protocols. Imaginary scenarios depicting the way in 
which predictive models can be incorporated into HACCP concepts have been 
published, 122 as has a useful review of the application of predictive food microbiology 
in the meat industry. 113 Similarly, predictive microbiology is an important element 
of Quantitative Microbial Risk Assessment (QMRA). Models are useful decision 
support tools, but it should be remembered that models are, at best, only a simplified 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 214 Wednesday, November 12, 2003 12:54 PM 








representation of reality. The application of model predictions should be tempered 
with previous experience and with knowledge of other microbial ecology principles 
that may be experienced in the food by the organism. 158 Sources of data and models 
relevant to the growth of L. monocytogenes in seafood and that could be part of a 
QMRA have been published. 158 

5.6.1 Bias and Accuracy 

Some criticism of the term "validation" revolves around the difficulty in quantifying 
just how well models perform their predictive role. Error occurs implicitly in the 
use of data for modeling and the use of those models for the prediction of growth 
of microorganisms. There are a number of potential sources of error: the homoge- 
neity of foods; the completeness of the environmental factors used to collect the 
data; conversion of empirical results to a mathematical function; and fitting the 
models to the data. 159 For example, the overall errors in the application of growth 
models to the growth of Pseudomonas species in food and in laboratory media have 
been quantified. 144 The authors made the point that the error was small in the case 
of culture medium but great in the case of food, and went on to quantify the influence 
of food structure and composition on the overall error. Sutherland et al. 172 found 
that much of the published work on E. coli 0157:H7 was done under conditions 
outside of the experimental values used to develop their growth model. These 
workers also reported that validation with data from cheeses and meats was difficult 
because the original authors often did not report experimental conditions such as 
NaCl content or pH. In these cases, poor predictions were often made. Similar 
observations were made when a growth model for B. cereus was being validated. 170 
It is clear in the above cases that some quantification of the deviation of the 
predictions from the observed values would be useful. Many measures of such 
quantification of error in the validation process have been made. 157 Additionally, 
however, Ross 157 has proposed using simple indices of the performance of models 
as a step towards an objective definition of the term "validated model." These indices 
give an indication of the confidence with which those models can be used (accuracy 
factor), and whether the model displays bias towards fail-dangerous predictions (bias 
factor). The accuracy factor is defined as: 




a 



lo g( GT predicted /GT observed )| /n > 



a r . ir\° predicted ooservea ' i ' /r '~i\ 

Accuracy factor = 10 (5.7) 

where GT predicted is the predicted generation time and GT observed is the observed 
generation time, and n is the number of observations. The less accurate the predic- 
tions the larger the accuracy factor. 
The bias factor is defined as: 

Bias factor = i (Ilog<G W<' G W,, d >'<>> (J g) 

If no disagreement between predicted and observed values occurs then the bias factor 
is equal to l. However, a value of the bias factor greater than l indicates a fail- 

2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 215 Wednesday, November 12, 2003 12:54 PM 








dangerous model because it will predict generation times longer than actually 
observed. It should be noted, however, that when rate values are used to compute 
the bias factor, a fail-dangerous model will have a bias factor of less than 1. 

As mathematical techniques advance, so does the process of comparing models. 
The use of artificial neural networks has been identified as a useful alternative 
technique for modeling microbial growth. Neural networks also lend themselves to 
quantifying comparisons between models and suitable indices have been suggested. 85 

5.6.2 Validation Using Literature Values 

The most common method of validation is the use of literature data. This is based 
on the assumption that if the published experiments were performed under well- 
defined conditions that do not differ markedly from those used to develop the model, 
then the model predictions should be reasonably reflected in the published data. A 
large number of models have been validated using published information including 
models for Y. enterocolitica 22 ' 100 ' 171 Aeromonas hydrophila, n2 Clostridium botuli- 
num, 76 S. enteritidis 23 and E. coli 0157:H7, 23 ' 173 L. monocytogenes 71 ' 111 and a number 
of other microorganisms. 55 

There are, however, some potentially serious limitations to the use of literature 
data for validation of predictive models. Additional food components are frequently 
responsible for deviations between predicted and observed values in validation exper- 
iments. For example, Tienungoon et al. 181 predicted the growth limits of L. monocy- 
togenes as a function of temperature, pH, NaCl, and lactic acid. The authors used two 
strains of L. monocytogenes, Scott A (a pathogenic strain) and L5 (a wild-type strain 
isolated from cold-smoked salmon). Experiments were carried out in broth culture at 
a wide range of environmental conditions. Aliquots of the inoculated media were 
observed for a period of 90 days to determine whether the conditions supported growth. 
Data from the experimental program were modeled using a probability model for 
growth. Figure 5.2 shows the growth boundary predicted by the model for the case of 
no added lactic acid, and a water activity of 0.992 (representing 0.5% NaCl in a typical 
culture medium) as a function of temperature and pH. This boundary is plotted 
alongside the data from the literature (Table 5.3). Generally, the model predicted values 
that were in good agreement with literature values. However, where deviation from 
the observed measurements occurred, this was usually explained by additional iden- 
tifiable preservative factors in the system, and these are described in Table 5.3. 

A similar issue arises using the growth boundary model of McKellar and Lu, 116 
which predicts the growth limits of E. coli 0157:H7 as a function of temperature, 
pH, NaCl, sucrose, and acetic acid. These authors used five strains of E. coli 0157:H7 
growing in broth culture for a period of 72 h to determine whether the conditions 
supported growth. Data from the experimental program were modeled using a 
probability model for growth. 

This boundary is plotted alongside the data from the literature (Table 5.4) in 
Figure 5.3. As above, the model predicted values that were in good agreement with 
literature values. Again, however, deviation from the observed measurements 
occurred, due to additional identifiable preservative factors, which are described in 
Table 5.4. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 216 Wednesday, November 12, 2003 12:54 PM 









Temperature (C) 

FIGURE 5.2 The growth boundary predicted by the model of Tienungoon et al. 181 for the 
case of no added lactic acid, and a water activity of 0.992 (representing 0.5% NaCl in a 
typical culture medium) as a function of temperature and pH. This boundary is plotted 
alongside the data from the literature described in Table 5.3. 

Other sources of error associated with the use of literature data include lack of 
information on preincubation conditions that might result in the development of acid 
tolerance; use of selective media for enumerating microorganisms; lack of estimates 
on variability; and presence of factors in foods that are not taken into account in 
models (e.g., preservatives). 39 It appears that the most appropriate method for vali- 
dation might be to use data derived under well-controlled conditions, so that the 
model's performance will not be unfairly biased. 157 Unsafe predictions and lack of 
published information on error also limit the usefulness of literature data, and 
emphasize the need to validate against new data. 57 

5.6.3 Validation in Foods 

The most common method for validating models using new data is to carry out 
experiments directly in the food product of concern. Thus, several models have been 
validated directly in food products including survival of L. monocytogenes in 
uncooked-fermented meat, 195 fishery products, 158 or pate; 69 survival of Campylo- 
bacter jejuni in a variety of foods; 54 growth of L. monocytogenes in dairy products; 129 
growth of L. innocua in Bologna-type sausage; 81 growth of Staphylococcus aureus 
in sterile foods; 190 growth of E. coli 0157:H7 on raw ground beef; 188 growth of L. 
monocytogenes in sterile foods; 189 growth of Shigella flexneri in sterile foods; 207 
growth of E. coli on raw displayed pork; 73 growth of Y. enterocolitica in seafood; 143 
and growth of Listeria in a range of foods. 174 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 217 Wednesday, November 12, 2003 12:54 PM 








TABLE 5.3 

Literature Values Used in the Validation of the Growth Boundary Model 

Shown in Figure 5.2 



Data 


Temp. 










Obs. 


Ref. 


(°C) 


pH 


Other Hurdles 3 
Listeria monocytogenes - 


Matrix b 
— Growth Data 


Ref. 


Time" 


1 





6.4 




Chicken broth 


191 




2 


4 


5.5 


0.5% NaCl 


TSBYG 


71 




3 


4 


6.5 


4% NaCl 


TSBYG 


71 




4 


5 


6 


4.5% NaCl 


Tryptose phosphate broth 


42 




5 


5 


6.27 


0.05% NaCl 


Minced beef 


111 




6 


5 


6.6 


0.05% NaCl 


UHT milk 


111 




7 


5.3 


5.6 


0.05% NaCl 


Vacuum packed lean beef 


111 




8 


6 


6.52 


0.05% NaCl 


Chicken legs 


111 




9 


7 


5.8 


0.05 [0.004]% acetic acid d 


Tryptose broth 


3 




10 


7 


5.8 


0.05% citric acid 


Tryptose broth 


3 




11 


7 


6.4 


0.05% NaCl 


Nonfat milk 


111 




12 


8 


5.8 


0.05% NaCl 


Minced beef 


111 




13 


8 


6.4 


0.05% NaCl 


Skimmed milk 


111 




14 


10 


4.5 




TSBYE 


181 




15 


10 


4.6 


Poised with citric acid 


TSB 


167 




16 


10 


4.8 


Poised with lactic acid 


TSB 


167 




17 


10 


5.6 




Tryptic meat broth 


16 




18 


10 


6.63 


0.277% NaCl + 170 ppm nitrite 


Vacuum packed ham 


111 




19 


10 


7 


a w = 0.96 

Listeria monocytogenes — 


Tryptic meat broth 
No Growth Data 


16 




20 


4 


5 




TSBYG 


72 


28 d 


21 


4 


6.5 


8% NaCl 


TSBYG 


71 


70 d 


22 


5 


6 


4.5% NaCl + nitrite 


Tryptose phosphate broth 


42 


NS 


23 


7 


4.4 


0.2% citric acid 


Tryptose broth 


3 


400 h 


24 


7 


4.6 




TSBYG 


72 


28 d 


25 


10 


4.25 




TSBYE 


181 


NS 


26 


10 


4.4 




TSBYG 


72 


28 d 


27 


10 


4.4 


Poised with citric acid 


TSB 


167 


28 d 


28 


10 


4.6 


Poised with lactic acid 


TSB 


167 


28 d 


29 


28 


4 


6 [5.12]% acetic acid 


BHI 


40 


62 d 


30 


28 


4 


9 [3.78]% lactic acid 


BHI 


40 


62 d 


31 


30 


4 


0.029% citric acid 


TSBYE 


51 


42 d 


32 


30 


4.5 


0.068 [0.043]% acetic acid 


TSBYE 


51 


42 d 


33 


30 


4.5 


0.043 [0.008]% lactic acid 


TSBYE 


51 


42 d 


Note: 


NS = Not stated. 









a These are responsible for the deviation of the data points from the growth boundary predicted by the 

model. 

b The following matrices refer to commonly used microbiological growth media: TSBYG; Tryptose- 

phosphate broth; Tryptose broth; TSBYE; TSB; Tryptic meat broth; BHI. 

c Time for which no growth was observed. 

d Concentration of acetic and lactic acids expressed as total, with undissociated in square brackets. 






2004 by Robin C. McKellar and Xuewen Lu 










1237_C05.fm Page 218 Wednesday, November 12, 2003 12:54 PM 







8 



7 - 



6 - 



5 - 



4 - 







+ 



11 



■+ 



19 



Growth 



No Growth + 



Lines: Model of McKellar and Lu (116), 
0.5% NaCI, no acetic acid 




90% growth probability 

50% growth probability 
10% growth probability 



10 



15 



20 



25 



30 



35 



40 



Temperature (C) 




FIGURE 5.3 The growth boundary predicted by the model of McKellar and Lu 1 1 6 that predicts 
the growth limits of Escherichia coli 0157:H7 as a function of temperature, pH, NaCI, sucrose, 
and acetic acid. This boundary is plotted alongside the data from the literature described in 
Table 5.4. 




Dynamic modeling has also been validated, 25 where predictions from FoodMi- 
croModel have been applied to the growth of L. monocytogenes and Salmonella in 
a range of foods incubated under constant as well as fluctuating temperatures. The 
authors found that generally the accuracy of prediction under the fluctuating tem- 
peratures was similar to the isothermal conditions, although inhibition by natural 
microflora did decrease the expected growth of L. monocytogenes in milk. Signif- 
icant deviation of predictions of the growth of bacteria growing as colonies when 
immobilized in gel occurred when predictions were made from isothermal growth 
in broth. 125 ' 126 

Validation of combined growth of the spoilage bacteria Pseudomonas, 
Shewanella putrefaciens, Brochothrix therm osphacta, and lactic acid bacteria was 
made in modified atmosphere packaged fish as a function of temperature and con- 
centration of carbon dioxide. 91 Combined models based on polynomial, Belehradek, 
and Arrhenius equations were developed and validated by comparison with experi- 
mental growth rates of these bacteria obtained on three Mediterranean fish species. 
Predictions of the models based on the Belehradek and Arrhenius equations were 
judged satisfactory overall. This approach has been modified 90 to determine a pro- 
cedure for modeling the shelf life of fish. Similarly, a quadratic response surface 
model has been used to describe the maximum specific growth rate of Y. enterocolit- 
ica. The model predicted growth rates as a function of refrigeration temperature and 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 219 Wednesday, November 12, 2003 12:54 PM 








TABLE 5.4 

Literature Values Used in the Validation of the Growth Boundary Model 

Shown in Figure 5.3 



Data 


Temp. 












Obs. 


Ref. 


(°C) 


pH 


Other Hurdles 3 




Matrix b 


Ref. 


Time' 








£. co// — Growth Data 








1 


8.2 


5.7 






Ground mutton 


83 




2 


10 


4 


0.5% NaCl 




TSB 


116 




3 


10 


5.5 


Poised with lactic acid 




TSBYE 


52 




4 


10 


5.5 


Poised with citric acid 




TSBYE 


52 




5 


10 


5.5 


0.5% NaCl 




BHI 


37 




6 


10 


5.5 


5% NaCl 




BHI 


41 










E. coli — No Growth Data 








7 


4 


4 






TSBYE 


52 


21 d 


8 


5 


3.6 






"Condiments" 


183 


7d 


9 


5 


4.8 


a w = 0.99 

w 




TSB 


156 


48 h 


10 


5 


6.5 


5% NaCl 




BHI 


41 


12 d 


11 


5 


«7.2 






Cucumber slices 


1 


10 d 


12 


8 


5.5 


0.5% NaCl 




BHI 


37 


10 d 


13 


10 


3.5 


0.5% NaCl 




TSB 


116 


72 h 


14 


10 


4 


Poised with acetic acid 




TSBYE 


52 


21 d 


15 


10 


4.5 


5% NaCl 




BHI 


41 


12 d 


16 


10 


5 


Poised with lactic acid 




TSBYE 


52 


21 d 


17 


10 


5 


Poised with citric acid 




TSBYE 


52 


21 d 


18 


12 


5.5 


30% sucrose, a w = 0.972 




BHI 


36 


24 h 


19 


12 


«7 






Shredded carrot 


1 


10 d 


20 


15 


4 


0.5% NaCl 




TSB 


116 


72 h 


21 


30 


5.1 


0.1 [0.03]% acetic acid as 


vinegar 41 


Nutrient agar 


68 


4d 


22 


37 


4.5 


Poised with lactic acid 




TSBYE 


74 


14 d 



Note: NS = Not stated. 

a These are responsible for the deviation of the data points from the growth boundary predicted by the 

model. 

b The following matrices refer to commonly used microbiological growth media: TSBYE; TSB; BHI; 

Nutrient agar. 

c Time for which no growth was observed. 

d Concentration of acetic acid expressed as total, with undissociated in square brackets. 



modified atmosphere and comparisons of the model predictions were made with 
growth rates obtained in seafood deliberately inoculated with Y. enterocolitica} 1 

Validations of the growth of L. monocytogenes in tryptose phosphate broth and 
in chicken and in beef have been made as a function of changing the pH and sodium 
chloride concentration. 133 Predictions of the growth of L. monocytogenes were then 
made using either a square root model 148 or a response surface polynomial model. 42 
The square root model predicted growth rates at between and 25°C with a 






2004 by Robin C. McKellar and Xuewen Lu 










1237_C05.fm Page 220 Wednesday, November 12, 2003 12:54 PM 








coefficient of determination of between 98.36 and 99.63%. The response surface 
polynomial model, however, predicted generation times at 5 to 25°C with between 
and 17.4% difference between the observed and expected generation times in 
broth. Of greater significance in terms of validation in food here are the large 
differences observed in the generation time at pH 5.6 and 8°C (25.5 h) and the 
generation time predicted by the Pathogen Modeling Program (PMP) in these con- 
ditions in tryptose phosphate broth (5.3 h). The PMP is a web-based package 
developed in the U.S. that contains secondary models of the effects of environmental 
factors (mainly pH, concentration of NaCl, and temperature) on the survival, growth, 
and inactivation of major food-borne pathogenic bacteria in broth. A divergence 
from predicted values was also shown at temperatures between and 3.5°C in the 
square root model. 

Predictions of the growth of Bacillus cereus from PMP were validated for its 
growth from spores in boiled rice. 114 An analysis of variance showed that there was 
no statistically significant difference between the observed and measured growth 
rates in boiled rice and predictions made from PMP. Modeled predictions were fail- 
safe for generation time and exponential growth rate at all temperatures. Although 
the model was fail-safe for lag phase duration at 20 and 30°C, it was not at 15°C. 

Modeling the growth of filamentous fungi is rare. The growth of three strains 
of heat-resistant fungi, as influenced by water activity adjusted using sucrose was 
modeled using the Baranyi model 13 to fit the changing colony diameter. 185 Modeling 
the growth of filamentous fungi has also been done using a model derived from the 
cardinal model family. The model was successfully fitted on data sets from a range 
of filamentous fungi whose growth was affected by a range of humectants including 
sodium chloride, glucose/fructose as a mixture, and glycerol and at different pH 
values. Further cardinal values were extracted from the literature and the model was 
used to predict the evolution of the radial growth of Penicillium rocqueforti and 
Paecilomyces variotii. 162 

In spite of the effort expended to develop and validate models, it is rare to find 
a model developed in broth that accurately predicts behavior in food systems. Models 
tend to fail-safe, and provide somewhat conservative predictions. 23 ' 54 ' 7381 ' 100 ' 114129 ' 170 ' 172 
Indeed, the use of faster-growing strains has been suggested to provide a margin of 
safety. 123 ' 131 ' 194 Although many validations of models show that there is a fail-safe 
tendency and hence a margin of safety in growth prediction, some manufacturers 
of foods find that the error is unacceptable and the margin of safety provided by 
such models may well be more conservative than is desirable for many food appli- 
cations. There are, however, examples of situations where the model makes what 
are clearly unsafe predictions, and these usually involve an overestimation of the 
extent of lag time. 69 ' 188 - 190 

An alternative approach is to develop models directly in food products. This is 
not possible in many cases, due to the requirement of appropriate facilities for 
incorporating pathogens into the process under carefully controlled conditions. In 
spite of this limitation, models have been developed for growth of L. monocytogenes 
on vacuum-packed cooked meats 63 and liver pate; 69 inactivation of Salmonella typh- 
imurium in reduced calorie mayonnaise; 121 inactivation of Enterobacteriaceae and 
Clostridia 18 and growth of Lactobacillus spp. in dry fermented sausage; 19 growth of 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 221 Wednesday, November 12, 2003 12:54 PM 







Clostridium botulinum in processed cheese; 169 thermal inactivation of L. 
monocytogenes 145 and Enterococcus faecium 161 during high-speed short-time pas- 
teurization; and the thermal inactivation of E. faecium during cooking of Bologna 
sausage. 208 These models generally provide good estimates of the behavior of food- 
borne pathogens in food processes. However, it is questionable if effort should be 
expended developing models specific for all food processes. Improved validation 
techniques for models derived in broth or other model systems would appear to have 
more general applicability. 

It has been suggested that models should only be regarded as first estimates of 
the behavior of pathogens, and that additional studies with products giving poor 
predictions should be undertaken. 195 Inclusion of additional data into models will 
often improve their predictive ability 207 ; however, it is important that users of these 
models take great care in their use, and ensure that predictions are carefully validated 
in any product of concern. 




REFERENCES 

1. Abdulraouf, U.M., L.R. Beuchat, and M.S. Ammar. 1993. Survival and growth of 
Escherichia coli 0157-H7 on salad vegetables. Appl. Environ. Microbiol. 
59:1999-2006. 

2. Abiss, J.S. 1983. Injury and resuscitation of microbes with reference to food micro- 
biology. Irish J. Food Sci. Technol. 7:69-81. 

3. Ahamad, N. and E.H. Marth. 1989. Behavior of Listeria monocytogenes ail, 13, 21, 
and 35 °C in tryptose broth acidified with acetic, citric, or lactic- acid. J. Food Prot. 
52:688-695. 

4. Alexandrou, O., C.D.W. Blackburn, and M.R. Adams. 1995. Capacitance measure- 
ment to assess acid-induced injury to S. enteritidis PT4. Int. J. Food Microbiol. 
27(l):27-36. 

5. Andrews, W.H. 1986. Resuscitation of injured Salmonella spp. and coliforms from 
foods. J. Food Prot. 49:62-75. 

6. Augustin, J.C., A. Brouillaud-Delattre, L. Rosso, and V. Carlier. 2000. Significance 
of inoculum size in the lag time of Listeria monocytogenes. Appl. Environ. Microbiol. 
66:1706-1710. 

7. Augustin, J.C. and V. Carlier. 2000. Mathematical modelling of the growth rate and 
lag time for Listeria monocytogenes. Int. J. Food Microbiol. 56:29-51. 

8. Augustin, J.C. and V. Carlier. 2000. Modelling the growth rate of Listeria monocy- 
togenes with a multiplicative type model including interactions between environmen- 
tal factors. Int. J. Food Microbiol. 56:53-70. 

9. Augustin, J.C, V. Carlier, and J. Rozier. 1998. Mathematical modelling of the heat 
resistance of Listeria monocytogenes. J. Appl. Microbiol. 84:185-191. 

10. Augustin, J.C, L. Rosso, and V. Carlier. 2000. A model describing the effect of 
temperature history on lag time for Listeria monocytogenes. Int. J. Food Microbiol. 
57:169-181. 

11. Baird- Parker, A.C. 1980. Organic acids. In Microbial Ecology of Foods, Vol. 1: 
Factors Affecting Life and Death of Micro -Organisms. Academic Press, New York, 
pp. 126-135. International Commission on Microbiological Specifications for 
Foods. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 222 Wednesday, November 12, 2003 12:54 PM 








12. Baird- Parker, A.C. and B. Freame. 1967. Combined effect of water activity, pH and 
temperature on the growth of Clostridium botulinum from spore and vegetative cell 
inocula. J. Appl. Bacteriol. 30:420-429. 

13. Baranyi, J. and T.A. Roberts. 1994. A dynamic approach to predicting bacterial growth 
in food. Int. J. Food Microbiol, 23:277-294. 

14. Baranyi, J. and T.A. Roberts. 1995. Mathematics of predictive food microbiology. 
Int. J. Food Microbiol. 26:199-218. 

15. Barrell, R.A.E. 1988. The survival and recovery of Salmonella typhimurium phage 
type U285 in frozen meats and tryptone soya yeast extract broth. Int. J. Food Micro- 
biol. 6:309-316. 

16. Begot, C, I. Lebert, and A. Lebert. 1997. Variability of the response of 66 Listeria 
monocytogenes and Listeria innocua strains to different growth conditions. Food 
Microbiol. 14:403-412. 

17. Bellara, S.R., P.J. Fryer, CM. McFarlane, C.R. Thomas, P.M. Hocking, and B.M. 
Mackey. 1999. Visualization and modelling of the thermal inactivation of bacteria in 
a model food. Appl. Environ. Microbiol. 65:3095-3099. 

18. Bello, J. and M.A. Sanchezfuertes. 1995. Application of a mathematical model for 
the inhibition of Enterobacteriaceae and Clostridia during a sausage curing process. 
/. FoodProt. 58:1345-1350. 

19. Bello, J. and M.A. Sanchezfuertes. 1995. Application of a mathematical model to 
describe the behaviour of the Lactobacillus spp. during the ripening of a Spanish dry 
fermented sausage (chorizo). Int. J. Food Microbiol. 27:215-227. 

20. Benito, A., G. Ventoura, M. Casadei, T. Robinson, and B. Mackey. 1999. Variation 
in resistance of natural isolates of Escherichia coliO\51 to high hydrostatic pressure, 
mild heat, and other stresses. Appl. Environ. Microbiol. 65:1564-1569. 

21. Bennik, M.H.J., E.J. Smid, F.M. Rombouts, and L.G.M. Gorris. 1995. Growth of 
psychrotrophic foodborne pathogens in a solid surface model system under the influ- 
ence of carbon dioxide and oxygen. Food Microbiol. 12:509-519. 

22. Bhaduri, S., C.O. Turnerjones, R.L. Buchanan, and J.G. Phillips. 1994. Response 
surface model of the effect of pH, sodium chloride and sodium nitrite on growth of 
Yersinia ente roc otitic a at low temperatures. Int. J. Food Microbiol. 23:333-343. 

23. Blackburn, CD., L.M. Curtis, L. Humpheson, C Billon, and P.J. McClure. 1997. 
Development of thermal inactivation models for Salmonella enteritidis and Escher- 
ichia coli 0157:H7 with temperature, pH and NaCl as controlling factors. Int. J. Food 
Microbiol. 38:31-44. 

24. Bouttefroy, A., M. Linder, and J.B. Milliere. 2000. Predictive models of the combined 
effects of curvaticin 13, NaCl and on the behaviour of Listeria monocytogenes ATCC 
15313 in broth. J. Appl. Microbiol. 88; 919-929. 

25. Bovill, R., J. Bew, N. Cook, M. D'Agostino, N. Wilkinson, and J. Baranyi. 2000. 
Predictions of growth for Listeria monocytogenes and Salmonella during fluctuating 
temperature. Int. J. Food Microbiol. 59:157-165. 

26. Bovill, R.A., J. A. Shallcross, and B.M. Mackey. 1994. Comparison of the fluorescent 
redox dye 5-cyano-2,3-ditolyltetrazolium chloride with /7-iodonitrotetrazolium violet 
to detect metabolic activity in heat-stressed Listeria monocytogenes cells. J. Appl. 
Bacteriol. 77(4): 35 3-35 8. 

27. Brocklehurst, T.F. 1994. Delicatessen salads, and chilled prepared fruit and vegetable 
products. In Evaluation of Shelf Life of Foods — Principles and Practice, Man, D. and 
Jones, A. (Eds.). Chapman & Hall, Glasgow, pp. 87-126. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 223 Wednesday, November 12, 2003 12:54 PM 








28. Brocklehurst, T.R, G.A. Mitchell, Y.P. Ridge, R. Seale, and A.C. Smith. 1995. The 
effect of transient temperatures on the growth of Salmonella typhimurium LT2 in 
gelatin gel. Int. J. Food Microbiol. 27:45-60. 

29. Brocklehurst, T.R, G.A. Mitchell, and A.C. Smith. 1997. A model experimental gel- 
surface for the growth of bacteria on foods. Food Microbiol. 14:303-311. 

30. Brocklehurst, T.R, M.L. Parker, P.A. Gunning, H.P. Coleman, and M.M. Robins. 
1995. Growth of food-borne pathogenic bacteria in oil- in- water emulsions. II. Effect 
of emulsion structure on growth parameters and form of growth. /. Appl. Bacteriol. 
78:609-615. 

31 . Brocklehurst, T.R, M.L. Parker, P.A. Gunning, and M.M. Robins. 1993. Microbiology 
of emulsions: physicochemical aspects. Lipid Technol. July/ August: 83-88. 

32. Brocklehurst, T.R and P.D.G. Wilson. 2000. The role of lipids in controlling microbial 
growth. Grasas y Aceitas 51:66-73. 

33. Brouillaud-Delatre, A., M. Make, C. Collette, C. Mattei, and C. Lahellec. 1997. 
Predictive microbiology of dairy products: influence of biological factors affecting 
growth of Listeria monocytogenes. J. AOAC Int. 80:913-919. 

34. Brul, S. and P. Coote. 1999. Preservative agents in foods: mode of action and microbial 
resistance mechanisms. Int. J. Food Microbiol. 50:1-17. 

35. Buchanan, R.L. and L.K. Bagi. 1997. Microbial competition: effect of culture con- 
ditions on the suppression of Listeria monocytogenes Scott A by Carnobacterium 
piscicola. J. Food Prot. 60:254-261. 

36. Buchanan, R.L. and L.K. Bagi. 1997. Effect of water activity and humectant identity 
on the growth kinetics of Escherichia coli 0157:H7. Food Microbiol. 14:413-423. 

37. Buchanan, R.L., L.K. Bagi, R.V. Goins, and J.G. Phillips. 1993. Response-surface 
models for the growth-kinetics of Escherichia coli 0157H7. Food Microbiol. 
10:303-315. 

38. Buchanan, R.L. and M.H. Golden. 1998. Interactions between pH and malic acid 
concentration on the inactivation of Listeria monocytogenes. J. Food Safety 18:37-48. 

39. Buchanan, R.L., M.H. Golden, and J.G. Phillips. 1997. Expanded models for the non- 
thermal inactivation of Listeria monocytogenes. J. Appl. Microbiol. 82:567-577. 

40. Buchanan, R.L., M.H. Golden, and R.C. Whiting. 1993. Differentiation of the effects 
of pH and lactic or acetic acid concentration on the kinetics of Listeria monocytogenes 
inactivation. J. Food Prot. 56:474-478 and 484. 

41 . Buchanan, R.L. and L.A. Klawitter. 1992. The effect of incubation- temperature, initial 
pH, and sodium chloride on the growth- kinetics of Escherichia coli 0157-H7. Food 
Microbiol. 9:185-196. 

42. Buchanan, R.L., H.G. Stahl, and R.C. Whiting. 1989. Effects and interactions of 
temperature, pH, atmosphere, sodium chloride and sodium nitrite on the growth of 
Listeria monocytogenes. J. Food Prot. 52:844-851. 

43. Buchanan, R.L. and R.C. Whiting. 1996. Risk assessment and predictive microbiol- 
ogy. /. Food Prot. 31-36. 

44. Cabo, M.L., M.A. Murado, M.P. Gonzalez, and L. Pastoriza. 2000. Dose-response 
relationships. A model for describing interactions, and its application to the combined 
effect of nisin and lactic acid on Leuconostoc mesenteroides. J. Appl. Microbiol. 
88:756-763. 

45. Casadei, M.A., R. Ingram, E. Hitchings, J. Archer, and J.E. Gaze. 2001. Heat resis- 
tance of Bacillus cereus, Salmonella typhimurium and Lactobacillus delbrueckii in 
relation to pH and ethanol. Int. J. Food Microbiol. 63:125-134. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 224 Wednesday, November 12, 2003 12:54 PM 








46. Chea, F.P., Y.H. Chen, T.J. Montville, and D.W. Schaffner. 2000. Modeling the 
germination kinetics of Clostridium botulinum 56A spores as affected by temperature, 
pH, and sodium chloride. /. Food Prot. 63:1071-1079. 

47. Cheroutre-Vialette, M. and A. Lebert. 2000. Growth of Listeria monocytogenes as a 
function of dynamic environment at 10°C and accuracy of growth predictions with 
available models. Food Microbiol. 17:83-92. 

48. Chhabra, A.T., W.H. Carter, R.H. Linton, and M.A. Cousin. 1999. A predictive model 
to determine the effects of pH, milk fat, and temperature on thermal inactivation of 
Listeria monocytogenes. J. Food Prot. 62:1143-1149. 

49. Christian, J.H.B. 1980. Reduced water activity. In Microbial Ecology of Foods, Vol. 
1: Factors Affecting Life and Death of Micro -Organisms. Academic Press, New York, 
pp. 70-91. 

50. Clayson, D.H.F. and R.M. Blood. 1957. Food perishability: the determination of the 
vulnerability of food surfaces to bacterial infection. J. Sci. Food Agri. 8:404-414. 

51. Conner, D.E., V.N. Scott, and D.T. Bernard. 1990. Growth, inhibition, and survival 
of Listeria monocytogenes as affected by acidic conditions. J. Food Prot. 53:652-655. 

52. Conner, D.E. and J.S. Kotrola. 1995. Growth and survival of Escherichia coli 0157- 
H7 under acidic conditions. Appl. Environ. Microbiol. 61:382-385. 

53. Cooper, A.L., A.C.R. Dean, and C. Hinshelwood. 1968. Factors affecting the growth 
of bacterial colonies on agar plates. Proc. Royal Soc. B. 171:175-199. 

54. Curtis, L.M., M. Patrick, and CD. Blackburn. 1995. Survival of Campylobacter jejuni 
in foods and comparison with a predictive model. Lett. Appl. Microbiol. 21:194-197. 

55. Daughtry, G.J., K.R. Davey, and K.D. King. 1997. Temperature dependence of growth 
kinetics of food bacteria. Food Microbiol. 14:21-30. 

56. Davis, M. J., P.J. Coote, and C.P. O'Byrne. 1996. Acid tolerance in Listeria mono- 
cytogenes: the adaptive acid tolerance response (ATR) and growth-phase-dependent 
acid resistance. Microbiology (UK) 142:2975-2982. 

57. Delignettemuller, M.L., L. Rosso, and J.P. Flandrois. 1995. Accuracy of microbial growth 
predictions with square root and polynomial models. Int. J. Food Microbiol. 27:139-146. 

58. Devlieghere, F., B. VanBelle, and J. Debevere. 1999. Shelf life of modified atmosphere 
packed cooked meat products: a predictive model. Int. J. Food Microbiol. 46:57-70. 

59. Dilworth, M. J., A.R. Glenn, W.N. Konings, I.R. Booth, R.K. Poole, T.A. Krulwich, 
R.J.Rowbury, J.B. Stock, J.L. Slonczewski, G.M. Cook, E. Padan, H. Kobayashi, 
G.N. Bennett, A. Matin, and V. Skulachev. 1999. Problems of adverse pH and bacterial 
strategies to combat it. Bacterial Response to pH, Novartis Foundation Symposium 
221. John Wiley & Sons, New York. 221:4-18. 

60. Dodd, C. 1990. Detection of microbial growth in food by cryosectioning and light 
microscopy. Food Sci. Tech. Today 4(3): 180-182. 

61. Dodd, C. and W.M. Waites. 1991. The use of toluidine blue for in situ detection of 
micro-organisms in foods. Lett. Appl. Microbiol. 13:220-223. 

62. Duffy, G., A. Ellison, W. Anderson, M.B. Cole, and G.S.A.B. Stewart. 1995. Use of 
bioluminescence to model the thermal inactivation of Salmonella typhimurium in the 
presence of a competitive microflora. Appl. Environ. Microbiol. 61:3463-3465. 

63. Duffy, L.L., P.B. Vanderlinde, and F.H. Grau. 1994. Growth of Listeria monocytogenes 
on vacuum-packed cooked meats: effects of pH, a w , nitrite and ascorbate. Int. J. Food 
Microbiol. 23:377-390. 

64. Eifert, J.D., C.R. Hackney, M.D. Pierson, S.E. Duncan, and W.N. Eigel. 1997. Acetic, 
lactic, and hydrochloric acid effects on Staphylococcus aureus 196E growth based 
on a predictive model. J. Food Sci. 62:174-178. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C05.fm Page 225 Wednesday, November 12, 2003 12:54 PM 








65. Eklund, T. 1983. The antimicrobial effects of dissociated and undissociated sorbic 
acid at different pH levels. /. Appl. Bacteriol. 54:383-389. 

66. Ellajosyula, K.R., S. Doores, E.W. Mills, R.A. Wilson, R.C. Anantheswaran, and S.J. 
Knabel. 1998. Destruction of Escherichia coli 0157:H7 and Salmonella typhimurium 
in Lebanon bologna by interaction of fermentation pH, heating temperature, and time. 
J. FoodProt. 61:152-157. 

67. Ellison, A., S.F. Perry, and G.S.A.B. Stewart. 1991. Bioluminescence as a real-time 
monitor of injury and recovery in Salmonella typhimurium. Int. J. Food Microbiol. 
12:323-332. 

68. Entani, E., M. Asai, S. Tsujihata, Y. Tsukamoto, and M. Ohta. 1998. Antibacterial 
action of vinegar against food-borne pathogenic bacteria including Escherichia coli 
0157:H7. J. Food Prot. 61:953-959. 

69. Farber, J.M., R.C. McKellar, and W.H. Ross. 1995. Modelling the effects of various 
parameters on the growth of Listeria monocytogenes on liver pate. Food Microbiol. 
12:447-453. 

70. Fernandez, A., M.J. Ocio, RS. Fernandez, and A. Martinez. 2001. Effect of heat 
activation and inactivation conditions on germination and thermal resistance param- 
eters of Bacillus cereus spores. Int. J. Food Microbiol. 63:257-264. 

71. Fernandez, P.S., S.M. George, C.C. Sills, and M.W. Peck. 1997. Predictive model of 
the effect of C0 2 , pH, temperature and NaCl on the growth of Listeria monocytogenes. 
Int. J. Food Microbiol. 37:37-45. 

72. George, S.M., B.M. Lund, and T.F. Brocklehurst. 1988. The effect of pH and tem- 
perature on initiation of growth of Listeria monocytogenes. Lett. Appl. Microbiol. 
6:153-156. 

73. Gill, CO., G.G. Greer, and B.D. Dilts. 1998. Predicting the growth of Escherichia 
coli on displayed pork. Food Microbiol. 15:235-242. 

74. Glass, K.A., J.M. Loeffelholz, J.P. Ford, and M.P. Doyle. 1992. Fate of Escherichia 
coli 0157/H7 as affected by pH or sodium chloride and in fermented, dry sausage. 
Appl. Environ. Microbiol. 58:2513-2516. 

75. Gomez, R.F. and Herrero, A. A. 1983. Chemical preservation of foods. In Food 
Microbiology, Rose, A.H. (Ed.). Academic Press, London, pp. 78-116. 

76. Graham, A.F., D.R. Mason, and M.W. Peck. 1996. A predictive model of the effect 
of temperature, pH and sodium chloride on growth from spores of non-proteolytic 
Clostridium botulinum. Int. J. Food Microbiol. 31:69-85. 

77. Hills, B.P., C.E. Manning, Y.P. Ridge, and T.F. Brocklehurst. 1996. NMR water 
relaxation, water activity and bacterial survival in porous media J. Sci. Food Agric. 
71:185-194. 

78. Hills, B.P., C.E. Manning, Y.P. Ridge, and T.F Brocklehurst. 1997. Water availability 
and the survival of Salmonella typhimurium in porous systems. Int. J. Food Microbiol. 
36:187-198. 

79. Hills, B.P. and K.M. Wright. 1994. A new model for bacterial growth in heterogeneous 
systems. /. Theor. Biol. 168:31-41. 

80. Hoffmans, CM., D.Y.C. Fung, and C.L. Kastner. 1997. Methods and resuscitation 
environments for the recovery of heat-injured Listeria monocytogenes: a review. 
/. Rapid Meth. Automat. Microbiol. 5(4):249-268. 

81. Houtsma, P.C., M.L. Kant-Muermans, EM. Rombouts, and M.H. Zwietering. 1996. 
Model for the combined effects of temperature, pH, and sodium lactate on growth 
rates of Listeria innocua in broth and Bologna-type sausages. Appl. Environ. Micro- 
biol. 62:1616-1622. 



2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 226 Wednesday, November 12, 2003 12:54 PM 








82. Hurst, A. 1977. Bacterial injury: a review. Can. J. Microbiol. 23:935-944. 

83. International Commission on Microbiological Specifications for Foods. 1996. Micro- 
organisms in Foods, Vol. 5: Microbiological Specifications of Food Pathogens. Blackie 
Academic and Professional, London. 

84. Johnson, S., E. Parsons, S. Stringer, T.F. Brocklehurst, C. Dodd, M. Morgan, and W. 
Waites. 1996. Multixenic growth of micro-organisms in food. Food Sci. Tech. Today 
12(l):53-56. 

85. Jeyamkondan, S., D.S. Jayas, and R.A. Holley. 2001. Microbial growth modeling 
with artificial neural networks. Int. J. Food Microbiol. 64:343-354. 

86. Juneja, V.K. and B.S. Eblen. 1999. Predictive thermal inactivation model for Listeria 
monocytogenes with temperature, pH, NaCl, and sodium pyrophosphate as controlling 
factors. J. Food Prot. 62:986-993. 

87. Juneja, V.K., B.S. Marmer, and B.S. Eblen. 1999. Predictive model for the combined 
effect of temperature, pH, sodium chloride, and sodium pyrophosphate on the heat 
resistance of Escherichia coli 0157:H7. J. Food Safety 19: 147—160. 

88. Katsaras, K. and L. Leistner. 199 1 . Distribution and development of bacterial colonies 
in fermented sausages. Biofouling 5:1145-124. 

89. Koutsoumanis, K., K. Lambropoulou, and G.J.E. Nychas. 1999. A predictive model 
for the non-thermal inactivation of Salmonella enteritidis in a food model system 
supplemented with a natural antimicrobial. Int. J. Food Microbiol. 49:63-74. 

90. Koutsoumanis, K. and G.J.E. Nychas. 2000. Application of a systematic experimental 
procedure to develop a microbial model for rapid fish shelf life predictions. Int. 
J. Food Microbiol. 60:171-184. 

91. Koutsoumanis, K.P., P.S. Taoukis, E.H. Drosinos, and G.J.E. Nychas. 2000. Appli- 
cability of an Arrhenius model for the combined effect of temperature and C0 2 
packaging on the spoilage microflora of fish. Appl. Environ. Microbiol. 66:3528-3534. 

92. Kroll, R.G. and R.A. Patchett. 1992. Induced acid tolerance in Listeria monocytoge- 
nes. Lett. Appl. Microbiol. 14:224-227. 

93. Labuza, T.P. and Bell, L.N. 1984. Moisture Sorption. Practical Isotherm Measurement 
and Use, 2nd ed. American Association of Cereal Chemists, St. Paul, MN. 

94. Lebert, I., V. Robles-Olvera, and A. Lebert. 2000. Application of polynomial models 
to predict growth of mixed cultures of Pseudomonas spp. and Listeria in meat. Int. 
J. Food Microbiol. 61:27-39. 

95. Lee, I.S., J.L. Slonczewski, and J.W. Foster. 1994. A low-pH-inducible, stationary-phase 
acid tolerance response in Salmonella typhimurium. J. Bacteriol. 176:1422-1426. 

96. Leguerinel, I. and P. Mafart. 2001. Modelling the influence of pH and organic acid 
types on thermal inactivation of Bacillus cereus spores. Int. J. Food Microbiol. 63:29-34. 

97. Leistner, L. 1992. Food preservation by combined methods. Food Res. Int. 
25:151-158. 

98. Leistner, L. and L.G.M. Gorris. 1995. Food preservation by hurdle technology. Trends 
Food Sci. Tech. 6:41-46. 

99. Le Marc, Y., V. Huchet, CM. Bourgeois, J.P Guyonnet, P. Mafart, and D. Thuault. 
2002. Modelling the growth kinetics of Listeria as a function of temperature, pH and 
organic acid concentration. Int. J. Food Microbiol. 73:219-237 . 

100. Little, C.L., M.R. Adams, W.A. Anderson, and M.B. Cole. 1994. Application of a 
log-logistic model to describe the survival of Yersinia enterocolitica at sub-optimal 
pH and temperature. Int. J. Food Microbiol. 22:63-71. 

101. Llaudes, M.K., L.H. Zhao, S. Duffy, and D.W. Schaffner. 2001. Simulation and 
modelling of the effect of small inoculum size on time to spoilage by Bacillus 
stearothermophilus. Food Microbiol. 18:395-405. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 227 Wednesday, November 12, 2003 12:54 PM 








102. Mackey, B.M. and CM. Derrick. 1982. The effects of sub-lethal injury by heating, 
freezing, drying and gamma-radiation on the duration of the lag phase of Salmonella 
typhimurium. J. Appl. Bacteriol. 53:243-251. 

103. Mackey, B.M., E. Boogard, CM. Hayes, and J. Baranyi. 1994. Recovery of heat- 
injured Listeria monocytogenes. Int. J. Food Microbiol. 22(4): 227-237. 

104. Malakar, P., T.F. Brocklehurst, A.R. Mackie, RD.G. Wilson, M.H. Zwietering, and 
K. Van't Riet. 2000. Microgradients in bacterial colonies: use of fluorescent ratio 
imaging, a non- invasive technique. Int. J. Food Microbiol. 56:71-80. 

105. Malakar, P.K., D.E. Martens, M.H. Zwietering, C. Beal, and K. Van't Riet. 1999. 
Modelling the interactions between Lactobacillus curvatus and Enterobacter cloacae. 
II. Mixed cultures and shelf life predictions. Int. J. Food Microbiol. 51:67-79. 

106. Manas, P., R. Pagan, I. Leguerinel, S. Condon, P. Mafart, and F. Sala. 2001. Effect 
of sodium chloride concentration on the heat resistance and recovery of Salmonella 
typhimurium. Int. J. Food Microbiol. 63:209-216. 

107. Martens, D.E., C. Beal, P.K. Malakar, M.H. Zwietering, and K. Van't Riet. 1999. 
Modelling the interactions between Lactobacillus curvatus and Enterobacter cloacae. 
I. Individual growth kinetics. Int. J. Food Microbiol. 51:53-65. 

108. Marth, E.H. 1998. Extended shelf life refrigerated foods: microbiological quality and 
safety. Food Tech. 52:57-62. 

109. Mattila, T. and A.J. Frost. 1988. The growth of potential food poisoning organisms 
on chicken and pork muscle surfaces. J. Appl. Bacteriol. 65:455-461. 

110. Mattila, T. and A.J. Frost. 1988. Colonization of beef and chicken muscle surfaces 
by Escherichia coli. Food Microbiol. 5:219-230. 

111. McClure, P.J., A.L. Beaumont, J.P. Sutherland, and T.A. Roberts. 1997. Predictive 
modelling of growth of Listeria monocytogenes: the effects on growth of NaCl, pH, 
storage temperature and NaN02. Int. J. Food Microbiol. 34:221-232. 

112. McClure, P.J., M.B. Cole, and K.W. Davies. 1994. An example of the stages in the 
development of a predictive mathematical model for microbial growth: the effects of 
NaCl, pH and temperature on the growth of Aeromonas hydrophila. Int. J. Food 
Microbiol. 23:359-375. 

113. McDonald, K. and D.W. Sun. 1999. Predictive food microbiology for the meat 
industry: a review. Int. J. Food Microbiol. 52:1-27. 

114. McElroy, D.M., L.A. Jaykus, and P.M. Foegeding. 2000. Validation and analysis of 
modeled predictions of growth of Bacillus cereus spores in boiled rice. J. Food Prot. 
63:268-272. 

115. McKay, A.L. and A.C. Peters. 1995. The effect of sodium chloride concentration and 
pH on the growth of Salmonella typhimurium colonies on solid medium. J. Appl. 
Bacteriol. 79:353-359. 

116. McKellar, R.C. and X.W. Lu. 2001. A probability of growth model for Escherichia 
coli 0157:H7 as a function of temperature, pH, acetic acid, and salt. J. Food Prot. 
64:1922-1928. 

117. McKellar, R.C, G. Butler, and K. Stanich. 1997. Modelling the influence of temper- 
ature on the recovery of Listeria monocytogenes from heat injury. Food Microbiol. 
14:617-625. 

118. McKellar, R.C, R. Moir, and M. Kaleb. 1994. Factors influencing the survival and 
growth of Listeria monocytogenes on the surface of Canadian retail wieners. J. Food 
Prot. 57:387-392. 

119. McMeekin, T.A., J.N. Olley, T. Ross, and D.A. Ratkowsky. 1993. Predictive Micro- 
biology: Theory and Application. Wiley, New York. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 228 Wednesday, November 12, 2003 12:54 PM 








120. McMeekin, T.A., K. Presser, D. Ratkowsky, T. Ross, M. Salter, and S. Tienungoon. 
2000. Quantifying the hurdle concept by modelling the bacterial growth/no growth 
interface. Int. J. Food Microbiol. 55:93-98. 

121. Membre, J.M., V. Majchrzak, and I. Jolly. 1997. Effects of temperature, pH, glucose, 
and citric acid on the inactivation of Salmonella typhimurium in reduced calorie 
mayonnaise. J. Food Prot. 60:1497-1501. 

122. Miles, D.W. and T. Ross. 1999. Identifying and quantifying risks in the food produc- 
tion chain. Food Aust. 51:298-303. 

123. Miles, D.W., T. Ross, J. Olley, and T.A. McMeekin. 1997. Development and evalu- 
ation of a predictive model for the effect of temperature and water activity on the 
growth rate of Vibrio parahaemolyticus. Int. J. Food Microbiol. 38:133-142. 

124. Mitchell, A.J. and J.W.T. Wimpenny. 1997. The effects of agar concentration on the 
growth and morphology of submerged colonies of motile and non-motile bacteria. 
/. Appl. Bacteriol. 83:76-84 

125. Mitchell, G.A., T.F. Brocklehurst, R. Parker, and A.C. Smith. 1994. The effect of 
transient temperatures on the growth of Salmonella typhimurium LT2. I. Cycling 
within the growth region. J. Appl. Bacteriol. 77:113-119. 

126. Mitchell, G.A., T.F. Brocklehurst, R. Parker, and A.C. Smith. 1995. The effect of 
transient temperatures on the growth of Salmonella typhimurium LT2. II. Excursions 
outside the growth region. /. Appl. Bacteriol. 79:128-134. 

127. Mossel, D.A.A., J.E.L. Corry, C.B. Struijk, and R.M. Baird. 1995. Essentials of the 
Microbiology of Foods: A Textbook for Advanced Studies. Wiley, Chichester, UK. 

128. Morinigo, M.A., R. Cornax, D. Castro, E. Martinez-Manzanares, and J.J. Borrego. 
1990. Viability of Salmonella spp. and indicator micro-organisms in seawater using 
membrane diffusion chambers. Antonie van Leeuwenhoek 57(2): 109—1 18. 

129. Murphy, P.M., M.C. Rea, and D. Harrington. 1996. Development of a predictive 
model for growth of Listeria monocytogenes in a skim milk medium and validation 
studies in a range of dairy products. J. Appl. Bacteriol. 80:557-564. 

130. Nerbrink, E., E. Borch, H. Blom, and T. Nesbakken. 1999. A model based on 
absorbance data on the growth rate of Listeria monocytogenes and including the 
effects of pH, NaCl, Na- lactate and Na-acetate. Int. J. Food Microbiol. 47:99-109. 

131. Neumeyer, K., T. Ross, and T.A. McMeekin. 1997. Development of a predictive 
model to describe the effects of temperature and water activity on the growth of 
spoilage pseudomonads. Int. J. Food Microbiol. 38:45-54. 

132. Nicolai, B.M., J.F. van Impe, B. Verlinden, T. Martens, J. Vandewalle, and J. De 
Baerdemaeker. 1993. Predictive modelling of surface growth of lactic acid bacteria 
in vacuum-packed meat. Food Microbiol. 10:229-238. 

133. Nyati, H. 2000. Survival characteristics and the applicability of predictive mathemat- 
ical modelling to Listeria monocytogenes growth in sous vide products. Int. J. Food 
Microbiol. 56:123-132. 

134. O'Driscoll, B., C.G.M. Gahan, and C. Hill. 1996. Adaptive acid tolerance response 
in Listeria monocytogenes: isolation of an acid- tolerant mutant which demonstrates 
increased virulence. Appl. Environ. Microbiol. 62:1693-1698. 

135. Oscar, T.P. 1999. Response surface models for effects of temperature and previous 
growth sodium chloride on growth kinetics of Salmonella typhimurium on cooked 
chicken breast. J. Food Prot. 62:1470-1474. 

136. Oscar, T.P. 1999. Response surface models for effects of temperature and previous 
temperature on lag time and specific growth rate of Salmonella typhimurium on 
cooked ground chicken breast. J. Food Prot. 62:1111-1114. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 229 Wednesday, November 12, 2003 12:54 PM 








137. Oscar, T.P. 1999. Response surface models for effects of temperature, pH and previous 
growth pH on growth kinetics of Salmonella typhimurium in brain heart infusion 
broth. /. Food Prot. 62:106-111. 

138. Parente, E., M.A. Giglio, A. Ricciardi, and F. Clementi. 1998. The combined effect 
of nisin, leucocin F10, pH, NaCl and EDTA on the survival of Listeria monocytogenes 
in broth. Int. J. Food Microbiol. 40:65-75. 

139. Parker, M.L., T.F. Brocklehurst, P.A. Gunning, H.P Coleman, and M.M. Robins. 
1995. Growth of food-borne pathogenic bacteria in oil-in- water emulsions. I. Methods 
for investigating the form of growth of bacteria in model oil-in- water emulsions and 
dairy cream. J. Appl. Bacteriol. 78:601-608. 

140. Parker, M.L., P.A. Gunning, A.C. Macedo, EX. Malcata, and T.F. Brocklehurst. 1998. 
The micro structure and distribution of micro-organisms within mature Serra cheese. 
/. Appl. Microbiol. 84:523-530. 

141. Peters, A.C., J.W.T. Wimpenny, and J.P. Coombs. 1987. Oxygen profiles in, and in 
the agar beneath, colonies of Bacillus cereus, Staphylococcus albus and Escherichia 
coli. J. Gen. Microbiol. 133:1257-1263. 

142. Pin, C. and J. Baranyi. 1998. Predictive models as means to quantify the interactions 
of spoilage organisms. Int. J. Food Microbiol. 41:59-72. 

143. Pin, C, J. Baranyi, and G. de Fernando. 2000. Predictive model for the growth of 
Yersinia enterocolitica under modified atmospheres. /. Appl. Microbiol. 88:521-530. 

144. Pin, C, J.P. Sutherland, and J. Baranyi. 1999. Validating predictive models of food 
spoilage organisms. J. Appl. Microbiol. 87:491-499. 

145. Piyasena, P., S. Liou, and R.C. McKellar. 1998. Predictive modelling of inactivation 
of Listeria spp. in bovine milk during HTST pasteurization. Int. J. Food Microbiol. 
39:167-173. 

146. Potts, M. 1994. Desiccation tolerance of prokaryotes. Microbiol. Rev. 58:755-805. 

147. Presser, K.A., T. Ross, and D.A. Ratkowsky. 1998. Modelling the growth limits 
(growth no growth interface) of Escherichia coli as a function of temperature, pH, 
lactic acid concentration, and water activity. Appl. Environ. Microbiol. 64:1773-1779. 

148. Ratkowsky, D.A., R.K. Lowry, T.A. McMeekin, A.N. Stokes, and R.E. Chandle. 1983. 
Model for bacterial culture-growth rate throughout the entire biokinetic temperature- 
range. J. Bacteriol. 154(3): 1222-1226. 

149. Ratkowsky, D.A. and T. Ross. 1995. Modelling the bacterial growth/no growth inter- 
face. Lett. Appl. Microbiol. 20:29-33. 

150. Ray, B. 1979. Methods to detect stressed micro-organisms. J. Food Prot. 42:346-355. 

151. Riordan, D.C., G. Duffy, J.J. Sheridan, B.S. Eblen, R.C. Whiting, I.S. Blair, and D.A. 
Mcdowell. 1998. Survival of Escherichia coli 0157:H7 during the manufacture of 
pepperoni. J. Food Prot. 61:1 46— 151. 

152. Robins, M.M. and P.D.G. Wilson. 1994. Food structure and microbial growth. Trends 
Food Sci. Tech. 5:289-293. 

153. Robins, M.M., T.F. Brocklehurst, and P.D.G. Wilson. 1994. Food structure and the 
growth of pathogenic bacteria. Food Technol. Int. (Eur.) 31-36. 

154. Robinson, R.A. and R.H. Stokes. 1959. Electrolyte Solutions. Butterworths, London. 

155. Robinson, T.P., J.W.T. Wimpenny, and R.G. Earnshaw. 1991. pH gradients through 
colonies of Bacillus cereus and the surrounding agar. J. Gen. Microbiol. 
137:2885-2889. 

156. Rocelle, M., S. Clavero, and L.R. Beuchat. 1996. Survival of Escherichia coli 
0157:H7 in broth and processed salami as influenced by pH, water activity, and 
temperature and suitability of media for its recovery. Appl. Environ. Microbiol. 
62:8-2740. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 230 Wednesday, November 12, 2003 12:54 PM 








157. Ross, T. 1996. Indices for performance evaluation of predictive models in food 
microbiology. J. Appl. Bacteriol. 81:501-508. 

158. Ross, T., P. Dalgaard, and S. Tienungoon. 2000. Predictive modelling of the growth 
and survival of Listeria in fishery products. Int. J. Food Microbiol. 62:231-245. 

159. Ross, T. and J. Olley. 1997. Problems and solutions in the application of predictive 
microbiology. In Seafood Safety, Processing, and Biotechnology, Shahidi, R, Jones, 
Y., and Kitts, D.D. (Eds.). Technomic Publishing Company Inc., Lancaster, PA, pp. 
101-118. 

160. Ross, T., T.A. McMeekin, and J. Baranyi. 1999. Predictive microbiology and food 
safety. In Encyclopedia of Food Microbiology, Robinson, R.K., Batt, C.A., and Patel, 
P. (Eds.). Academic Press, New York, pp. 1699-1710. 

161. Ross, W.H., H. Couture, A. Hughes, T. Gleeson, and R.C. McKellar. 1998. A non- 
linear mixed effects model for the destruction of Enterococcus faecium in a high- 
temperature short-time pasteurizer. Food Microbiol. 15:567-575. 

162. Rosso, L. and T.P Robinson. 2001. A cardinal model to describe the effect of water 
activity on the growth of moulds. Int. J. Food Microbiol. 63:265-273. 

163. Schaffner, D.W. and T.P. Labuza. 1997. Predictive microbiology: where are we, and 
where are we going? Food Tech. 51:95-99. 

164. Skandamis, P.N. and G.J.E. Nychas. 2000. Development and evaluation of a model 
predicting the survival of Escherichia coli 0157:H7 NCTC 12900 in homemade 
eggplant salad at various temperatures, pHs, and oregano essential oil concentrations. 
Appl. Environ. Microbiol. 66:1646-1653. 

165. Skandamis, P., E. Tsigarida, and G-J.E. Nychas. 2000. Ecophysiological attributes of 
Salmonella typhimurium in liquid culture and within a gelatin gel with or without 
the addition of oregano essential oil. World J. Microbiol. Biotech. 16(1):31— 35. 

166. Sofos, J.N. and F.F. Busta. 1981. Antimicrobial activity of sorbate. /. Food Prot. 
44:614-622, 647. 

167. Sorrells, K.M., D.C. Enigl, and J.R. Hatfield. 1989. Effect of pH, acidulant, time, 
and temperature on the growth and survival of Listeria monocytogenes. J. Food Prot. 
52:571-573. 

168. Stecchini, M.L., M. Del Torre, S. Donda, and E. Maltini. 2000. Growth of Bacillus 
cereus on solid media as affected by agar, sodium chloride, and potassium sorbate. 
/. Food Prot. 63:926-929. 

169. Steeg, PET. and H.G.A.M. Cuppers. 1995. Growth of proteolytic Clostridium botu- 
linum in process cheese products. II. Predictive modeling. /. Food Prot. 
58:1100-1108. 

170. Sutherland, J.P, A. Aherne, and A.L. Beaumont. 1996. Preparation and validation of 
a growth model for Bacillus cereus: the effects of temperature, pH, sodium chloride, 
and carbon dioxide. Int. J. Food Microbiol. 30:359-372. 

171. Sutherland, J.P. and A.J. Bayliss. 1994. Predictive modelling of growth of Yersinia 
enterocolitica: the effects of temperature, pH and sodium chloride. Int. J. Food 
Microbiol. 21:197-215. 

172. Sutherland, J.P, A.J. Bayliss, and D.S. Braxton. 1995. Predictive modelling of growth 
of Escherichia coli 0157:H7: the effects of temperature, pH and sodium chloride. 
Int. J. Food Microbiol. 25:29-49. 

173. Sutherland, J.P, A.J. Bayliss, D.S. Braxton, and A.L. Beaumont. 1997. Predictive 
modelling of Escherichia coli 0157:H7: inclusion of carbon dioxide as a fourth factor 
in a pre-existing model. Int. J. Food Microbiol. 37:113-120. 

174. te Giffel, M.C. and M.H. Zwietering. 1999. Validation of predictive models describing 
the growth of Listeria monocytogenes. Int. J. Food Microbiol. 46:135-149. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 231 Wednesday, November 12, 2003 12:54 PM 








175. ter Steeg, P.F., F.H. Pieterman, and J.C. Hellemons. 1995. Effects of air/nitrogen, 
temperature and pH on energy-dependent growth and survival of Listeria innocua in 
continuous culture and water-in-oil emulsions. Food Microbiol. 12:471-485. 

176. Thomas, L.V. and J.W.T. Wimpenny. 1993. Method for investigation of competition 
between bacteria as a function of three environmental factors varied simultaneously. 
Appl. Environ. Microbiol. 59:1991-1997. 

177. Thomas L.V., J.W.T. Wimpenny, and G.C. Barker. 1997. Spatial interactions between 
subsurface bacterial colonies in a model system: a territory model describing the 
inhibition of Listeria monocytogenes by a nisin-producing lactic acid bacterium. 
Microbiology (UK) 143(Part 8): 2575-2582. 

178. Thomas, L.V., J.W.T. Wimpenny, and J.G. Davis. 1993. Effect of three preservatives 
on the growth of Bacillus cereus, verocytotoxigenic Escherichia coli and Staphylo- 
coccus aureus on plates with gradients of pH and sodium chloride concentration. Int. 
J. Food Microbiol. 17:289-301. 

179. Thomas, L.V., J.W.T. Wimpenny, and A.C. Peters. 1991. An investigation of four 
variables on the growth of Salmonella typhimurium using two types of gradient gel 
plates. Int. J. Food Microbiol. 14:261-275. 

180. Thomas, L.V., J.W.T. Wimpenny, and A.C. Peters. 1992. Testing multiple variables 
on the growth of a mixed inoculum of Salmonella strains using gradient plates. Int. 
J. Food Microbiol. 15:165-175. 

181. Tienungoon, S., D.A. Ratkowsky, T.A. McMeekin, and T. Ross. 2000. Growth limits 
of Listeria monocytogenes as a function of temperature, pH, NaCl, and lactic acid. 
Appl. Environ. Microbiol. 66:4979-4987. 

182. Troller, J. A. 1983. Effect of low moisture environments on the microbial stability of 
foods. In Food Microbiology, Rose, A.H. (Ed.). Academic Press, London, pp. 173-198. 

183. Tsai, Y.W. and S.C. Ingham. 1997. Survival of Escherichia coli 0157:H7 and Sal- 
monella spp. in acidic condiments. /. Food Prot. 60:751-755. 

184. Tuynenburg Muys, G. 1971. Microbial safety in emulsions. Process Biochem. 
6(6):25-28. 

185. Valik, L. and E. Pieckova. 2001. Growth modelling of heat-resistant fungi: the effect 
of water activity. Int. J. Food Microbiol. 63:11-17. 

186. Verrips, C.T. and J. Zaalberg. 1980. The intrinsic microbial stability of water-in-oil 
emulsions. I. Theory. Eur. J. Appl. Microbiol. Biotech. 10:187-196. 

187. Verrips, C.T., D. Smid, and A. Kerkhof. 1980. The intrinsic microbial stability of 
water-in-oil emulsions. II. Experimental. Eur. J. Appl. Microbiol. Biotec h. 10:7 '3—85. 

188. Walls, I. and V.N. Scott. 1996. Validation of predictive mathematical models describ- 
ing the growth of Escherichia coli 0157:H7 in raw ground beef. /. Food Prot. 
59:1331-1335. 

189. Walls, I. and V.N. Scott. 1997. Validation of predictive mathematical models describ- 
ing the growth of Listeria monocytogenes. J. Food Prot. 60:1142-1145. 

190. Walls, I., V.N. Scott, and D.T. Bernard. 1996. Validation of predictive mathematical 
models describing growth of Staphylococcus aureus. J. Food Prot. 59:11-15. 

191. Walker, S.J., P. Archer, and J.G. Banks. 1990. Growth of Listeria monocytogenes at 
refrigeration temperatures. /. Appl. Bacteriol. 68:157-162. 

192. Walker, S.L., T.F. Brocklehurst, and J.W.T. Wimpenny. 1997. The effects of growth 
dynamics upon pH gradient formation within and around subsurface colonies of 
Salmonella typhimurium. J. Appl. Microbiol. 82:610-614. 

193. Walker, S.L., T.F. Brocklehurst, and J.W.T. Wimpenny. 1998. Adenylates and adeny- 
late-energy charge in submerged and planktonic cultures of Salmonella enteritidis 
and Salmonella typhimurium. Int. J. Food Microbiol. 44:107-113. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C05.fm Page 232 Wednesday, November 12, 2003 12:54 PM 








194. Whiting, R.C. 1997. Microbial database building: what have we learned? Food Tech. 
51:82-109. 

195. Whiting, R.C. and M.O. Masana. 1994. Listeria monocytogenes survival model val- 
idated in simulated uncooked-fermented meat products for effects of nitrite and pH. 
/. Food Sci. 59:760-762. 

196. Wijtzes, T., KM. Rombouts, M.L.T. Kant-Muermans, K. Van't Riet, and M.H. Zwi- 
etering. 2001. Development and validation of a combined temperature, water activity, 
pH model for bacterial growth rate of Lactobacillus curvatus. Int. J. Food Microbiol. 
63:57-64. 

197. Wilson, P.D.G., T.F. Brocklehurst, S. Arino, D. Thuault, M. Jakobsen, M. Lange, J. 
Farkas, J.W.T. Wimpenny, and J.F. Van Impe. 2002. Modelling microbial growth in 
structured foods: towards a unified approach. Int. J. Food Microbiol. 73:275-289. 

198. Wilson, P.D.G., D.R. Wilson, and C.R. Waspe. 2000. Weak acids: disassociation in 
complex buffering systems and partitioning into oils. J. Sci. FoodAgric. 80:471-476. 

199. Wimpenny, J.W.T. 1979. The growth and form of bacterial colonies. /. Gen. Microbiol. 
114:483-486. 

200. Wimpenny, J.W.T. and J.P. Coombs. 1983. Penetration of oxygen into bacterial 
colonies. J. Gen. Microbiol. 129:1239-1242. 

201. Wimpenny, J.W.T., L. Leistner, L.V. Thomas, A.J. Mitchell, K. Katsaras, and P. Peetz. 
1995. Submerged bacterial colonies within food and model systems: their growth, 
distribution and interactions. Int. J. Food Microbiol. 28:299-315. 

202. Wimpenny, J.W.T. and M.W.A. Lewis. 1977. The growth and respiration of bacterial 
colonies. J. Gen. Microbiol. 103:9-18. 

203. Wimpenny, J.W.T. and P. Waters. 1984. Growth of microorganisms in gel-stabilized 
two-dimensional gradient systems. J. Gen. Microbiol. 130:2921-2926. 

204. Wimpenny, J.W.T. and P. Waters. 1987. The use of gel- stabilized gradient plates to 
map the responses of microorganisms to three or four factors varied simultaneously. 
FEMS Microbiol. Lett. 40:263-267. 

205. Windholz, M. 1983. The Merck Index. Merck and Co Inc., Rahway, NJ. 

206. Yu, L.S. and D.Y.C. Fung. 1993. Five-tube most-probable-number method using the 
Fung-Yu tube for enumeration of Listeria monocytogenes in restructured meat pro- 
ducts during refrigerated storage. Int. J. Food Microbiol. 18(2):97-106. 

207. Zaika, L.L., J.G. Phillips, J.S. Fanelli, and O.J. Scullen. 1998. Revised model for 
aerobic growth of Shigella flexneri to extend the validity of predictions at temperatures 
between 10 and 19°C. Int. J. Food Microbiol. 41:9-19. 

208. Zanoni, B., C. Peri, C. Garzaroli, and S. Pierucci. 1997. A dynamic mathematical 
model of the thermal inactivation of Enterococcus faecium during bologna sausage 
cooking. Food Sci. Techno I. Lebensm. Wiss. 30:727-734. 

209. Zhao, L., T.J. Montville, and D.W. Schaffner. 2000. Inoculum size of Clostridium 
botulinum 56A spores influences time-to-detection and percent growth-positive sam- 
ples. J. Food Sci. 65:1369-1375. 

210. Zhao, L., T.J. Montville, and D.W. Schaffner. 2002. Time to detection, percent- 
growth-positive and maximum growth rate models for Clostridium botulinum 56A 
at multiple temperatures. Int. J. Food Microbiol. 77:187-197. 




2004 by Robin C. McKellar and Xuewen Lu