Skip to main content

Full text of "كتب في علم الميكروبيولوجي"

See other formats



1237_C03.fm Page 63 Wednesday, November 12, 2003 12:40 PM 






~V 




O Secondary Models 

Thomas Ross and Paw Dalgaard 




CONTENTS 

3.1 Introduction 

3.1.1 Philosophy, Terminology, and Methodology 

3.2 Secondary Models for Growth Rate and Lag Time 

3.2.1 Square-Root-Type Models 

3.2.1.1 Temperature 

3.2.1.2 Water Activity 

3.2.1.3 pH 

3.2.1.4 Other Factors 

3.2.2 The Gamma Concept 

3.2.2.1 Expanding Existing Models 

3.2.3 Cardinal Parameter Models 

3.2.3.1 Secondary Lag Time Models and the Concept 
of Relative Lag Time 

3.2.4 Secondary Models Based on the Arrhenius Equation 

3.2.4.1 The Arrhenius Equation 

3.2.4.2 Mechanistic Modifications of the Arrhenius Model 

3.2.4.3 Empirical Modifications of the Arrhenius Model 

3.2.4.4 Application of the Simple Arrhenius Model 

3.2.5 Polynomial and Constrained Linear Polynomial Models 

3.2.6 Artificial Neural Networks 

3.3 Secondary Models for Inactivation 

3.4 Probability Models 

3.4.1 Introduction 

3.4.2 Probability Models 

3.4.2.1 Logistic Regression 

3.4.2.2 Confounding Factors 

3.4.3 Growth/No Growth Interface Models 

3.4.3.1 Deterministic Approaches 

3.4.3.2 Logistic Regression 

3.4.3.3 Relationship to the Minimum Convex Polyhedron 
Approach 

3.4.3.4 Artificial Neural Networks 

3.4.3.5 Evaluation of Goodness of Fit and Comparison 
of Models 




2004 by Robin C. McKellar and Xuewen Lu 













~V 






1237_C03.fm Page 64 Wednesday, November 12, 2003 12:40 PM 






3.4.4 Experimental Methods and Design Considerations 

3.4.4.1 Measuring Both Growth and Inactivation 

3.4.4.2 Inoculum Size 

3.4.4.3 Are There Absolute Limits to Microbial Growth? 

3.4.4.4 Experimental Design 

3.4.4.5 Conclusion 

Appendix A3. 1 — Characterization of Environmental Parameters Affecting 

Microbial Kinetics in Foods 

A 3 . 1 . 1 Temperature 

A3 . 1 . 2 Storage Atmosphere 

A3. 1 .3 Salt, Water-Phase Salt, and Water Activity 

A3. 1.4 pH 

A3. 1 .5 Added Preservatives Including Organic Acids, Nitrate, 
and Spices 

A3. 1.6 Smoke Components 

A3. 1 .7 Other Environmental Parameters 
References 



3.1 INTRODUCTION 

Changes in populations of microorganisms in foods over time (i.e., "microbial 
kinetics") are governed by storage conditions ("extrinsic" factors) and product 
characteristics ("intrinsic" factors). Collectively these have been termed "environ- 
mental parameters." They may represent simple situations, e.g., where the storage 
temperature is the only important factor influencing microbial kinetics, but in many 
foods the environmental parameters that influence microbial kinetics are complex 
and dynamic and include the combined effects of extrinsic factors such as temper- 
ature and storage atmosphere; intrinsic factors such as water activity, pH, naturally 
occurring organic acids, and added preservatives; and interactions between groups 
of microorganisms. 

Consistent with the widely accepted terminology introduced by Whiting and 
Buchanan (1993), we term those models that describe the response of microorgan- 
isms to a single set of conditions over time as "primary" models (see Chapter 2). 
Models that describe the effect of environmental conditions, e.g., physical, chemical, 
and biotic features, on the values of the parameters of a primary model are termed 
"secondary" models. 

Knowledge of the environmental parameters that most in uence growth of 
microorganisms in foods is essential for the development, as well as for the practical 
use, of predictive microbiology models. Secondary models that do not include all 
the environmental parameters important in a food are said to be "incomplete" (Ross, 
Baranyi, McMeekin, 2000) and require expansion (or simple "calibration" if those 
factors are constant) to accommodate their effect on microbial kinetics. The envi- 
ronmental parameters that are important for particular foods, however, are not always 
known. In those situations the systematic approach of predictive microbiology can 
help to elucidate the microbial ecology of the product. 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 65 Wednesday, November 12, 2003 12:40 PM 




~V 









In this chapter we consider a range of types of secondary models including 
those that model the probability that a predicted kinetic response will occur. The 
chapter includes descriptions and comparison of models, as considerations for 
development of robust, secondary models. Appendix A3. 1 details methods to mea- 
sure environmental factors of importance — an essential element of the application 
of predictive microbiology. 

3.1.1 Philosophy, Terminology, and Methodology 

The history of predictive microbiology, including the philosophical motivations of 
Roberts and Jarvis (1983), who first proposed the concept, w as traced by Ross and 
McMeekin (1994). From a purely pragmatic perspective, predictive microbiology 
aims to collect and make accessible computerized data on the behavior of microbial 
populations in response to defined environmental conditions, but mathematical mod- 
eling also provides a useful and rigorous framework for the hypothetico-deductive 
scientific process. To develop a consistent framework that enables us to understand 
and predict the microbial ecology of foods it is desirable to integrate the patterns 
of microbial behavior revealed in predictive modeling studies with knowledge of 
the physiology of microorganisms and physical and chemical processes and phe- 
nomena that occur in foods and food processes (Ross, Baranyi, McMeekin, 2000). 

Various types and categorizations of models are recognized. Empirical models 
are, essentially, pragmatic and simply describe a set of data in a convenient mathe- 
matical relationship with no consideration of underlying phenomena. Mechanistic 
models are built up from theoretical bases and, if they are correctly formulated, can 
allow the response to be interpreted in terms of known physical, chemical, and 
biological phenomena. An advantage of mechanistic approaches is that they tend to 
provide a better foundation for subsequent development and expansion of models; 
i.e., taken to their logical extreme, models for specific situations would simply be 
special, or reduced, cases of a much larger and holistic model that describes, quan- 
titatively, the microbial ecology of foods. The process of developing models that are 
able to be integrated readily with other models so as to describe more complex 
phenomena has been termed "nesting" or "embedding." A fuller explanation of the 
bene ts of that approach w as provided by Baranyi and Roberts (1995). 

In one sense, a model is the mathematical expression of a hypothesis. If this 
approach is adopted, it follows that the parameters in such models might be readily 
interpretable properties of the system under study, and that the mathematical form of 
the model would enable interpretation of the interactions between those factors. 
Interpretability of model parameters is a feature highly valued by many authors in the 
predictive microbiology literature (e.g., Augustin and Carlier, 2000a,b; Rosso et al., 
1993; Wijtzes et al., 1995). Although the development of predictive microbiology has 
seen the embedding of more and more mechanistic elements, or at least models whose 
structure and parameterization reflects known or hypothesized underlying phenomena, 
in practice many models currently available in predictive microbiology are not purely 
empirical, and none are purely mechanistic (Ross, Baranyi, McMeekin, 2000). 

Another, often cited, advantage of mechanistic models is that if they are built on 
sound theory they are more likely to facilitate prediction by extrapolation. Conversely, 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 66 Wednesday, November 12, 2003 12:40 PM 




~V 









as none of the models in use in predictive microbiology can be considered to be 
mechanistic, they can only be used to make predictions by interpolation. (Determi- 
nation of the interpolation region encompassed by a model is discussed in Sections 
3.2.5 and 3.4.3.4.) It is perhaps ironic, then, that 20 years of experience in predictive 
microbiology has not demonstrated the practical usefulness of mechanistic models 
that have been proposed to date (see Section 3.2.4). In general, even with good quality 
data the mechanistic models do not provide better fit and are us ually harder to work 
with than quasi-mechanistic or empirical models currently used. 

Predictive microbiology is a specific application of the field of mathematical 
modeling and, as such, the same rules of modeling as are applied in those other 
disciplines are relevant to the development of predictive food microbiology models. 
These have been discussed by various authors (Draper and Smith, 1981; McMeekin 
et al., 1993; Ratkowsky, 1993), and an overview is presented in Table 3.1. 

Experimental methods and design considerations relevant to kinetic models were 
discussed in detail in McMeekin et al. (1993; Chapter 2), Davies (1993), and Legan 
et al. (2002) and are also discussed in Chapter 1 . Two points that we feel are necessary 
to reiterate are the limitations of the central composite design in predictive micro- 
biology studies, and consideration of spoilage domains when growth of spoilage 
microorganisms is studied. Legan et al. (2002) accentuated the importance of exper- 
imental design in growth modeling studies stating: 

in other disciplines, such as engineering, central composite designs are commonly used 
for developing response surface models. For microbiological modeling, however, these 
designs have serious limitations and should be avoided. Central composite designs 
concentrate treatments in the centre of the design space and have fewer treatments in 
the extreme regions where biological systems tend to exhibit much greater variability. 

Microbial food spoilage is dynamic and in some cases relatively small changes in 
environmental parameters cause a complete shift in the micro or a responsible for 
product spoilage. Thus, to avoid modeling growth of spoilage microorganisms under 
conditions where they have no in uence on quality , a product-oriented approach that 
includes determination of the spoilage domain of specific micoor ganisms is often 
required (Dalgaard, 2002). 

We will not comment further on methodology appropriate to development of kinetic 
models, other than to say that to develop reliable secondary models an understanding 
of microbial physiology and its interaction with food environments and storage and 
processing conditions must be borne in mind in the design of experiments including 
the preparation of cultures and interpretation of measurements of population changes. 
This issue is particularly explored and exempli ed in Section 3 .4.4.4 concerning exper- 
imental considerations relevant to the development of growth limits models. 

3.2 SECONDARY MODELS FOR GROWTH RATE AND 
LAG TIME 

Implicit in the appropriate development and use of secondary models in predictive 
food microbiology is the ability to characterize foods, and the environment they 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 61 Wednesday, November 12, 2003 12:40 PM 





~V 






TABLE 3.1 

Some Considerations in the Selection of Models 



Subject 

Parameter estimation 
properties 

Stochastic assumption 



Parameter 
interpretability 



Parsimony 



Interpolation region 



Correct qualitative 
features 



"Extendibility" 
(embedding, nesting) 



Reasons 

Relates to the procedure and reliability of estimating the model parameters. 
In general, models should have parameters that are independent, identically 
distributed, normal or "iidn" (see, e.g., Ratkowsky, 1993) 

The form of the model, and choice of response variables, should be such 
that the difference between prediction and observations (or some 
mathematical transformation of them) is normally distributed, and that the 
magnitude of the error is independent of the magnitude of the response 
modeled. If not, the fitting can be dominated by some data, at the expense 
of other data 

As noted in the text, it is useful if the parameters have biological/ 
physical/chemical interpretations that can be readily related to the 
independent and dependent variables. This can simplify the process of 
model creation and also aid in understanding of the model (This may be 
less important than the behavior and performance of the model.) 

Models should have no more parameters than are required to describe the 
underlying behavior studied. Too many parameters can lead to a model that 
ts the error in the data, i.e., generates a model that is specific to a particular 
set of observations. Nonparsimonious models have poor predictive ability 

No models in predictive microbiology can be considered to be mechanistic 
and predictions can be made by interpolation only. Thus, the interpolation 
region de nes the useful range of applicability of the model. The 
interpolation region is affected by not only the range of individual variables, 
but also the experimental design (see Section 3.2.5) 

In mathematical terms, these are the analytical properties of the model 
function. They include convexity, monotonity, locations of extreme, and 
zero values. If biological considerations prescribe any of these, the model 
should reflect those properties accurately 

When a model is developed further (such as to include more or dynamically 
changing environmental factors) the new, more complex model should 
embody the old, simpler model as a special case 



Source: Modi ed from Ross, T, Baranyi, J., and McMeekin, T A. In Encyclopaedia of Food Microbiology, 
Robinson, R., Bart, C.A., and Patel, P. (Eds.), Academic Press, London, 2000, pp. 1699-1710. 






present to contaminating microorganisms, in terms of those biotic and abiotic ele- 
ments that affect the dynamics of the microbial population of interest. Methods to 
characterize the physicochemical environment, including temperature, gaseous 
atmosphere, salt and/or water activity, pH and organic acids, spices, smoke, and 
other components, are discussed in detail in Appendix A3.1. These topics are also 
considered in Chapter 5, including discussion of the in uence o f other organisms 
and heterogeneity in the environment. Another important element is the ability to 
characterize temporal changes in the environment. Techniques for modeling micro- 
bial population dynamics under time -varying conditions are considered in Chapter 7. 

2004 by Robin C. McKellar and Xuewen Lu 











~V 






1237_C03.fm Page 68 Wednesday, November 12, 2003 12:40 PM 






Within predictive microbiology the development and application of secondary 
models for growth rates and lag times have been extensively reviewed (Buchanan, 
1993b; Davey, 1999; Farber, 1986 ; ICMSF, 1996a,b; McDonald and Sun, 1999; 
McMeekin et al, 1993; Ross, 1999a,b; Ross and McMeekin, 1994; Skinner et al., 
1994; Whiting, 1995). This section describes types of secondary growth rate and 
lag time models that are currently available, but with particular focus on more 
recent developments, and also includes a detailed tabulation of models available 
for specific microorganisms. 

3.2.1 Square-Root-Type Models 

3.2.1.1 Temperature 

As discussed later (Section 3.2.4), in many cases the classical Arrhenius equation 
is inappropriate to describe the effect of suboptimal temperature on growth rates of 
microorganisms because the (apparent) activation energy (E a ) itself is temperature 
dependent. To overcome this problem Ratkowsky et al. (1982) suggested a simple 
empirical model (Equation 3.1). When this model was fitted to experimental growth 
rates the data were square -root transformed to stabilize their variance and this simple 
model and its numerous expansions are named square-root-type, Ratkowsky -type, 
or Beleradek-type models (McMeekin et al., 1993). These models, and the closely 
related cardinal parameter models (see Section 3.2.3), are probably the most impor- 
tant group of the secondary models within predictive microbiology. 




Jii =b-(T-T . ) (3.1) 

V " max v mm y v y 

where b is a constant and T is the temperature. The parameter T min , a theoretical 
minimum temperature for growth, is the intercept between the model and the tem- 
perature axis (Figure 3.1). T min is a model parameter and its value can be 5 to 10°C 
lower than the lowest temperature at which growth is actually observed. This inter- 
pretation differs from that embodied in the cardinal parameter models, as discussed 
in Section 3.2.3 and Chapter 4). 

From growth rates measured at several different constant temperatures the 
values of b and T min in Equation 3.1 can be determined by classical model tting 
techniques (see Chapter 4). Recently it was suggested that b and T min could be 
estimated from a single, optimally designed, experiment where growth resulting 
from a temperature profile is recorded (Bernaerts et al., 2000). These authors 
concluded that such an optimal, dynamic, one-step experiment would reduce the 
experimental work required to develop a model signi cantly and would have sub- 
stantial potential within predictive microbiology. So far this technique has not found 
wider use within predictive microbiology and its ability to estimate model param- 
eters accurately remains to be con rmed for different microorganisms and environ- 
mental parameters. 

Ratkowsky et al. (1983) expanded Equation 3.1 to include the entire biokinetic 
range of growth temperatures (Equation 3.2, Figure 3.1). From this model the optimal 



2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 69 Wednesday, November 12, 2003 12:40 PM 








Temperature (°C) 




FIGURE 3.1 Simulation of Equation 3.1 (solid line) and Equation 3.2 (dashed line), b = 
0.025 h°- 5 /°C, T ]nui = -8°C, c = 0.30°C- 1 ,and T max = 40°C. 

growth temperature can be determined by solving the following equation: c x (T opt 
- r min ) = exp[c x (r opt - r min )] - 1 (McMeekin et al., 1993). 




Ai 



M max =b-(T- TV ) • (1 - exp(c(r - T ))) 



(3.2) 



where b and c are constants, T is the temperature, r min the theoretical minimum 
temperature below which no growth is possible, and T max is the theoretical maximum 
temperature beyond which growth is not possible. 

While Ratkowsky et al. (1982, 1983) settled for an exponent of 2, the original 
Beleradek models had a variable exponent value. Dantigny (1998) and Dantigny and 
Molin (2000) used the concepts of dimensionless growth rate variables (effectively 
the same as the gamma factor concept; see Section 3.2.3) to explore the most 
appropriate value of the exponent for bacterial growth rate data using Beleradek- 
type models. They reported a correlation between the estimate of T min and the 
exponent value used and found that when T min and the exponent were simultaneously 
fitted by nonlinear re gression, thermophiles had lower fitted e xponent values than 
did mesophiles or psychrotrophic organisms. They reported that the use of the square- 
root model leads to an underestimation of the minimum temperature for growth 
when the exponent value is significantly less than 2. 

3.2.1.2 Water Activity 

McMeekin et al. (1987) found that growth responses of Staphylococcus xylosus 
followed Equation 3.1 at different values of water activity. T min was constant and 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 70 Wednesday, November 12, 2003 12:40 PM 







thus independent of water activity and Equation 3.3 was suggested to describe the 
combined effect of temperature and water activity (McMeekin et al., 1987). 



M =b-(T-T . )-a -a . (3.3) 

" max v mm ' -\l w w mm v ' 



where b and T mm are as previously de ned, a w is the water activity, and a w min is the 
theoretical minimum water activity below which growth is not possible. 

Later, Miles et al. (1 997) suggested that Equation 3.4 be used to take into account 
the effect of the entire biokinetic ranges of both temperature and water activity. 




M =b-(T-T . )-Q-exp(c(T-T )))■ (a -a . )(1 - exp(d(a -a ))) 

"max v rmn- / v .rv \ max /7/ \f v w wmin /v -^ v v w wmax 777 

(3.4) 



where b, c, T, T min , T max , a w and a w min are as previously defined, d is a fitted constant, 
and a w max is a theoretical maximum water activity beyond which growth is not 
possible. 

Most food-related microorganisms grow at water activities very close to 1.000 
and in those cases the expanded water activity term (i.e., containing a w max ) in 
Equation 3.4 is not needed to predict growth in foods. However, some microorgan- 
isms, e.g., several marine bacteria, have a substantial requirement for minerals. To 
model growth responses of these microorganisms, the inhibitory effect of high water 
activities, i.e., low salt concentrations, must be taken into account. For the human 
pathogen Vibrio parahae mo lyticus, a wmax has been determined to be 0.998. Some 
seafood spoilage bacteria are more inhibited by high water activity; e.g., growth of 
Halobacterium salinarium was only observed at a w values below 0.9 (Chandler and 
McMeekin, 1989; Doe and Heruwati, 1988; Miles et al., 1997). 

3.2.1.3 pH 

Vox Yersinia enterocolitica, Adams et al. (1991) found that growth responses followed 
Equation 3.1 at different values of pH. Again, T mm was constant and Equation 3.5 
was suggested. 




m =b-(T-T )JpH-pH . (3.5) 

V " max v min y \j -* ■* min v y 

where pH mm is the theoretical minimum pH below which growth is not possible and 
other parameters are as previously defined. 

On the basis of the observation of Cole et al. (1990) that growth rate was 
proportional to hydrogen ion concentration, Presser et al. (1997) introduced the 
following quasi-mechanistic term to describe the effect of pH on bacterial growth: 

2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 71 Wednesday, November 12, 2003 12:40 PM 








P m ax = Pop,x(l-10^'"" p// ) (3.6a) 



By analogy, another term was introduced for superoptimal (i.e., alkaline) pH con- 
ditions, leading to the following model for the entire biokinetic pH range: 



M max = IV x (1 " 1 pHmia ~ pH ) x (1 " 1 pH ~ pH ™ ) (3 .6b) 



The validity of that term was evaluated against an extensive data set for Escherichia 
coli growth, including variables of temperature, water activity, and lactic acid con- 
centration for a range of acid and alkaline environmental pH levels (see Equation 3.10). 
Wijtzes et al. (1995, 2001) continued the development of square -root-type mod- 
els and suggested Equation 3.7 for growth responses of Lactobacillus curvatus at 
different temperatures, a w values, and pH 

V=b-(a w -a wm J-(pH-pH im J-(pH-pH m J-(T-T mn f (3.7) 

3.2.1.4 Other Factors 

Equation 3.8 was suggested to model the effect of carbon dioxide-enriched (%C0 2 ) 
atmospheres on growth of the specific spoilage organism Photobacterium phospho- 
reum on sh (Dalgaard, 1995; Dalgaard et al., 1997). Later, similar but square -root- 
transformed terms were used to model the effect of C0 2 and sodium lactate (NaL) 
on growth of Lactobacillus sake and Listeria monocytogenes at a constant pH 
(Equation 3.9; Devlieghere et al., 1998, 2000a,b, 2001). 

(%ca -%ca) 




M =b(T-T . ) x -^ — (3.8) 

"max v min / (\/r^r\ v ' 

2 max 



"V •"*■ max V min 



>max = b 

•(T-T . ) 

V mm ' 



mm 



a w~ a ww* (3.9) 

• [CO, -CO, 

-v 2 max 2 

• J NaL -NaL 

v max 



As noted above, a more comprehensive square-root-type model that includes the 
effects of temperature, pH, water activity, and lactic acid has been suggested and 
developed in a series of publications (Presser et al., 1997; Ross, 1993a,b; Salter et al., 
1998; Tienungoon, 1988) and has been applied to Listeria monocytogenes and Escher- 
ichia coli growth rates. It was presented in its most complete form in Ross et al. (2003): 

2004 by Robin C. McKellar and Xuewen Lu 













~V 



1237_C03.fm Page 72 Wednesday, November 12, 2003 12:40 PM 






U = C 

r* max 



x(T-T mm )x(l-exp(dx(T-T))) 



max 



x . fa ^a ~ x (1 - expfe x (a -a ))) 

^7 w wmin v r\&\w wmax /// 

Xa/i-IO^""^ 
x A /i_io p// "- p// °'« 



(3.10) 



x 1- 



x 1- 



L4C 



1/ ■ x(l + 10 p// " M ") 

rain v J 



LAC 



D . x(l + 10 

mm v 



/* fl -pff 



mm 



) 




where c, d, and g are fitted parameters, LAC is the lactic acid concentration (mM), 
^min the minimum concentration (iriM) of undissociated lactic acid that prevents 
growth when all other factors are optimal, D inin the minimum concentration (mM) of 
dissociated lactic acid that prevents growth when all other factors are optimal, pK a is 
the pH for which concentrations of undissociated and dissociated lactic acid are equal, 
reported to be 3.86 (Budavari, 1989), and all other terms are as previously defined. 

One of the advantages of the square -root-type models, and the cardinal param- 
eters models, is that their form enables them to be readily simplified into models 
for special cases; e.g., in Equation 3.10, if one factor is held constant then the terms 
involving that factor simply reduce to constants. 

An example is a model developed for Listeria monocytogenes (Ross et al, in 
press; WHO/FAO, in press), in which the superoptimal water activity term is not 
relevant, and in which a term for the effect of nitrite on L. monocytogenes growth 
rate was also included. That novel term was based on analysis of the predictions of 
the Pathogen Modeling Program (Buchanan, 1993a; www.arserrc.gov/mfs/patho- 
gen.htm). The fitted model is shown in Equation 3.11. 




M 



max 



= 0.1626 
x(T- 0.60) x (1 - exp(0. 1 29 x (T - 5 1 .0))) 



xJ(a -0.925) 



w 



xJl-lO^-p* 



x 1- 




LAC 



4.55x(l + 10 pH " 386 ) 



( 



493 - NIT x 



1 + 



v 



(6.5 -pH) 
2 



\ 



/ 



\ 



/ 



(3.11) 



\ 



'493 



/ 



where NIT is the concentration of nitrite and all other terms are as previously defined. 

As shown in Figure 3.2, Equation 3.10 and Equation 3.11 represent a new 

generation of square -root-type models where the level of lactic acid in uences the 





2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 73 Wednesday, November 12, 2003 12:40 PM 








range of pH values for which growth is theoretically observed, reflecting the known 
interaction between pH and undissociated lactic acid, and also the individual growth 
rate suppressing effects of hydrogen ion concentration and undissociated lactic acid 
concentration. This was not the case for the environmental parameters included in 
Equation 3.1 to Equation 3.9. In those models, each term expressed how an envi- 
ronmental factor reduced the growth rate of a microorganism. However, for those 
models the expected multidimensional growth space was not influenced by levels 
of the different environmental parameters. This limitation of predictive models for 
growth rate has been recognized and has led, in part, to the development of growth/no 
growth models (discussed in Section 3.2.3 and Section 3.4). To make accurate 
predictions, a model can include terms to force the predicted growth rate to zero 
(Augustin and Carlier, 2000b; Le Marc et al., 2002). Alternatively, the probability 
of growth under the test conditions can first be assessed using a growth boundary 
model. If growth is possible, a growth rate model in combination with a lag time 
model can be used to estimate the extent of growth (Ross et al., in press; WHO/FAO, 
in press). 

It is also notable that the pH and lactic acid terms in Equation 3. 10 are effectively 
gamma-model type terms (see Section 3.2.2), in which the effect of the level of 
growth rate inhibitor is scaled between and 1, where 1 represents no inhibition, 
i.e., the optimal level of that environmental factor. In the case of lactic acid, the 
optimal level would be 0, while for pH the optimum is ~7. This illustrates the close 
relationship between square-root-type models, and those that embody the gamma 
concept, such as the cardinal parameter models. 

3.2.2 The Gamma Concept 

The concept of dimensionless growth factors, now known as the gamma (y) concept, 
was introduced in predictive microbiology by Zwietering et al. (1992). Later, minor 
changes and new developments were added (Wijtzes et al., 1998, 2001; Zwietering, 
1999; Zwietering et al., 1996). 

The gamma (y) concept relies on: 

1. The observation (e.g., Adams et al., 1991; McMeekin et al., 1987) that 
many factors that affect microbial growth rate act independently, and that 
the effect of each measurable factor on growth rate can be represented by 
a discrete term that is multiplied by terms for the effect of all other growth 
rate affecting factors, i.e.: 

(i, = /temperature) x/a w ) x/(pH) x /(organic acid) 
xXother^ x/(other 2 ) x ... ../(otherj 

2. That the effect on growth rate of any factor can be expressed as a fraction 
of the maximum growth rate (i.e., the rate when that environmental factor 
is at the optimum level) 



2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 74 Wednesday, November 12, 2003 12:40 PM 







a) 



in 
6 



>< 

to 

E 

zi. 




4.0 4.5 5.0 5.5 



6.0 6.5 7.0 7.5 8.0 8.5 9.0 
PH 



b) 



in 

6 




X 

TO 
5 



1.0-1 
0.9- 
0.8 
0.7 
0.6 
0.5 
0.4 
0.3 
0.2 
0.1 




• * 

// 



i — M- 





l i 
4.0 4.5 5.0 5.5 



i i i i 

6.0 6.5 7.0 7.5 



i i i 

8.0 8.5 9.0 



PH 

FIGURE 3.2 Simulation of Equation 3.7 (a) and Equation 3.10 (b) at a fixed temperature 
and water activity. pH inin is 4.0 and pH^^ is 9.0. For Equation 3.10, U ir]hl = 10 mM and D max 
= 1000 mM The concentrations of lactic acid (LAC) depicted are mM (dashed line), 50 
mM (dotted line), and 100 mM (dash-dotted line). 



Under completely optimal conditions each microorganism has a reproducible 
maximum growth rate, notwithstanding the potential effect of strain variability. As 
any environmental factor becomes suboptimal the growth rate declines in a predict- 
able manner, and the extent of that inhibition can be related to the optimum growth 
rate by calculating the relative rate at the test condition compared to that at the 
optimum. Thus, under the gamma concept approach, the cumulative effect of many 
factors poised at suboptimal levels can be estimated from the product of the relative 
inhibition of growth rate due to each factor, as indicated by Equation 3.12. The 
relative inhibitory effect of a specific environmental variable is described by a growth 
factor "gamma" (y), a dimensionless measure that has a value between and 1 (e.g., 
Equation 3.13 to Equation 3.15). 



2004 by Robin C. McKellar and Xuewen Lu 













~V 



1237_C03.fm Page 75 Wednesday, November 12, 2003 12:40 PM 






The relative inhibitory effect can be determined from the "distance" between 
the optimal level of the factor and the minimum (or maximum) level that completely 
inhibits growth by recourse to a predictive model. In the gamma model approach, 
the reference growth rate is |i max , so that reference levels of temperature, water 
activity, etc. are those that are the optimum for growth rate, usually represented as 
T opt , # w opt , pH opt , etc. The combined effect of several environmental factors is then 
determined by multiplication of their respective y factors (Equation 3.16). 



Y = 



Growth rate at actual environmental conditions 
Growth rate at optimal environmental conditions 

P m ax(7>„pH,etc.) 



M 



max opt 



(3.12) 



y(T) = 



T-T 

mm 



\ 2 



T -T - 

y opt mm j 



(3.13) 




YK) = 



a — a 

w w mm 



1 — a 



w mm 



y(pH) = 



{pH-pH^ydpH-pH) 



(P H n „, ~ P H mm ) • (P H n™ - P H o,J 



opt 



opt 



(3.14) 




(3.15) 



P,»x = P 



max opt 



V(T)-y(aJ-y(pH) 



(3.16) 



The effect of environmental parameters like carbon dioxide, sodium lactate, and 
nitrite has also been included in square-root-type models (see, e.g., Equation 3.8 to 
Equation 3.11). The absence of these inhibitory substances is optimal for growth 
and therefore the calculation of y factors requires information only about the lowest 
concentration of each substance that prevents growth (or, similarly, the maximum 
level that can be tolerated before growth ceases) analogous to minimum inhibitory 
concentrations (MICs). 



y(co 2 ) = 



' /<>co 2max -%co 2 v - 



V 



%co, _ - %co, t 

2 max 2 opt 



r %co 2wK -%co 2 v 



/ 



V 



%co 



2 max 



J 



(3.17) 



3.2.2.1 Expanding Existing Models 

Given that there is a finite number of models (see Table 3.5 and Table 3.6), and that 
few models include factors of relevance to all foods, some workers have attempted 
to integrate terms for specific variables from one model into another to suit a specific 





2004 by Robin C. McKellar and Xuewen Lu 











~V 






1237_C03.fm Page 76 Wednesday, November 12, 2003 12:40 PM 






food and the conditions of interest. Because of the assumption of independent action 
of growth rate inhibitors, the dimensionless y factors can, in principle, be readily 
exchanged between existing models and, at the time of writing, this is increasingly 
being done. Values of parameters like \i max , [\ ovP T nmv T opt , a wmm , pH mm , pH opt , pH max , 
and %C0 2max from which gamma factors can be derived are known for a considerable 
number of food-related pathogenic microorganisms. The approach was possibly 
taken to its logical conclusion by Augustin and Carlier (2000a,b) who collated, and 
integrated into a single model, literature data and observations for more than 15 
factors in foods that affect the growth rate of L. monocytogenes. 

For spoilage bacteria from chilled foods, growth kinetics at low temperatures 
are often well characterized but values of (J max , |i opt , r opt , pH opt , and pH max are fre- 
quently unknown or have not been determined accurately. This is the case, for 
example, for the specific spoilage organisms Photobacterium phosphoreum, 
Shewanella piitrefaciens, and Brochothrix thermosphacta. In this situation the clas- 
sical gamma concept cannot be used to develop a secondary model. However, when 
a simple square-root-type model including the effect of temperature and, e.g., C0 2 , 
has been developed for chilled product stored at a known pH (pH ref ) and water 
activity (a w ref ) then these models can be expanded at suboptimal growth conditions 
by addition of y-like factors, as shown in Equation 3.18 (Dalgaard et al., 2003). 



Umax = b 

• (T-T . ) 

v mm / 



nun 




^C0 lmax -%C0 2 )l%CQ lmm (3.18) 

Ma - a . ) I (a , — a . ) 

A/ v w w mm / v w ref w mm y 



(pH-pH m J/(pH ref - P H m J 

Clearly, this approach should be used with some caution because the assumption 
of independent action has not been tested for all environmental factor combinations. 
Thus, the range of applicability of the expanded model should be evaluated, e.g., 
by comparison with data from challenge tests or naturally contaminated products 
(Gimenez and Dalgaard, in press). (Section 3.2.5 discusses the expansion of existing 
polynomial models.) 

3.2.3 Cardinal Parameter Models 

Cardinal parameter models (CPMs) were introduced to predictive microbiology in 
1993 and have become an important group of empirical secondary models (Augustin 
and Carlier, 2000a,b; Le Marc et al., 2002; Messens et al., 2002; Pouillot et al., 
2003; Rosso, 1995, 1999; Rosso et al., 1993, 1995; Rosso and Robinson, 2001). 
The basic idea behind CPMs is to use model parameters that have a biological or 
graphical interpretation. When models are fitted to experimental data by nonlinear 
regression (see Chapter 4), this has the obvious advantage that appropriate starting 
values are easy to determine. General CPMs rely on the assumption that the inhib- 
itory effect of environmental factors is multiplicative, an assumption that was 

2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 77 Wednesday, November 12, 2003 12:40 PM 







formalized in the gamma (y) concept discussed above (Section 3.2.2). Thus, general 
CPMs consist of a discrete term for each environmental factor, with each term 
expressed as the growth rate relative to that when that factor is optimal; i.e., each 
term has a numerical value between and 1 . At optimal growth conditions all terms 
have a value of 1 and thus (i max is equal to |i. t (Equation 3.19). 

Equation 3. 19 to Equation 3.21 show a CPM that includes the effect of temper- 
ature (T), water activity (a w ), pH, inhibitory substances (c z ) and qualitative factors 
(k) on (i max (Augustin and Carlier, 2000a). This extensive CPM was developed from 
available literature data from many studies for growth of Listeria monocytogenes. 
The inhibitory substances included (1) undissociated acetic acid, lactic acid, and 
citric acid, (2) Na-benzoate, K -sorb ate, and the undissociated form of sodium nitrite, 
and (3) glycerol monolaurin, butylated hydroxy anisole, butylated hydroxy toluene, 
fer^-butylhydroquinone, C0 2 , caffeine, and phenol. In addition, the effect of com- 
petitive growth of microorganisms and the inhibitory effect due to specific types of 
foods were included in the model as qualitative factors. 



n p 

M IM x = K P , ■CM 1 {T)-CM i {aJ-CM l (pH)-Y[y(c l )-\\ k j ( 319 ) 



i=\ 



7=1 



CM, 



a 




0. 



(X-X )-(X-X . ) 

v may / V mm J 



n 



max 



mm 



.71-1 



(x -x . ) -[(X t -x . yoc-x t )-(x t -x) 

V nnt mm-' L\ nnt mm-' \ nnt / V nnt max-' 



opt 



mm 



opt 



mm 



opt 



opt 



«n-\)-X t +X -«■!)] 

vv y opt mm /J 



0. 



X < x„. 



mm 



(3.20) 




X . <x<x 

mm max 



x>x 



max 



Yfe) = 



(i-c./Mic.y 
o. 



c <MIC. 

i i 

c >MIC. 



(3.21) 



where X is temperature, water activity, or pH. X min and X max are, respectively, the 
values of X t below and above which no growth occurs, X opt is the value at which 
Umax i s equal to its optimal value (i opt . MIC t is the minimal inhibitory concentration 
of specific compounds above which no growth occurs. 

Within predictive microbiology various CPMs were developed during the 1 990s 
and in the same period different cardinal parameter temperature models were inde- 
pendently developed in other fields, e.g., to predict the effect of temperature on 
growth rates (r) of crops (Equation 3 .22; Yan and Hunt, 1 999; Yin and Wallace, 1 995). 



( 



r = 7i 



max 



T-T 

mm 



\ 



T -T 

V opt mm i 



( 



T -T 

max 



\ 



T -T 

v max opt J 



T —T 

max opt 

T -T ■ 
opt mm 



(3.22) 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 78 Wednesday, November 12, 2003 12:40 PM 






~V 




TABLE 3.2 

Parameter Values in Square-Root Type (Sqrt) and Cardinal Parameter 

Models (CPM) 



min 



opt 



^o 



Pt 



Organism 



Sqrt CPM Sqrt CPM Sqrt CPM Sqrt CPM 



Reference 



Escherichia coli 2.9 4.9 41.0 41.3 49.2 47.5 2.3 2.3 Rosso et al. (1993) 

Salmonella Typhimurium 3.8 5.7 39.8 40.0 51.1 49.3 1.7 1.7 Oscar (2002) 



pH min 



PH 



c i 



pH ma 



R 



opt 



Sqrt CPM Sqrt CPM Sqrt CPM Sqrt CPM 

Listeria monocytogenes 4.2 4.6 7.0 7.1 9.8 9.4 1.0 0.95 Rosso et al. (1995) 




In several ways CPMs resemble square-root models and responses of the two 
types of models can be practically identical, e.g., for the effect of temperature, water 
activity, and pH (Oscar, 2002, Rosso et al., 1993, 1995). Parameters in the two types 
of models are typically named T imn , T max , a w min , a w max , pH mm , and pH max . However, 
these model parameters are not de ned in entirely the same way for CPMs and 
square-root-type models. In fact, when identical data are fitted to the two types of 
models square -root-type models estimate lower T irdn , # w Illin , and pH min values and 
higher r max , # wmax , and pH max values (Table 3.2; see also Chapter 4). 

r mm values estimated by CPMs and square-root-type models often differ by 
~2°C as shown in Table 3.2. Table 3.3 shows that a 2°C difference of a r min value 
has a pronounced effect on (J max values predicted by both a square -root-type model 
and a CPM. Thus, parameter values estimated by using one of these types of models 




TABLE 3.3 

Effect of 7" min Values (-1°C and +1°C) on \i max Values Predicted by a 

Square-Root and a Cardinal Parameter Model at 4, 8, and 12°C 



Square-Root Model 3 



Cardinal Parameter Model' 



Temperature 
(°C) 


rhnax 


(h" 1 ) 


/o 


H"max 


(h" 1 ) 


o/ 

/o 


T = -1 °C 

'mm ' *- 


7min = +1°C 


Difference 


T = -1 °C 

'mm ' *~ 


^mi„=+1°C 


Difference 


4 


0.0216 


0.0078 


64 


0.0216 


0.0087 


60 


8 


0.0700 


0.0424 


40 


0.0718 


0.0485 


33 


12 


0.1461 


0.1046 


28 


0.1549 


0.1237 


20 



a The model of Ratkowsky et al. (1983) used with values of the model parameters b and c selected to 

obtain a T opt value ~37°C and a U opt value of- 1.0 h 1 . T max was 45.0°C. 

b The model of Rosso et al. (1993) used with T opt of 37°C, u opt 1.0 h 1 , and T max 45.0°C. 





2004 by Robin C. McKellar and Xuewen Lu 











~V 



1237_C03.fm Page 79 Wednesday, November 12, 2003 12:40 PM 







1 I 

-10 -5 



i 




t 1 1 1 1 1 1 1 1 r 

5 10 15 20 25 30 35 40 45 50 



Temperature (°C) 

FIGURE 3.3 Simulation of the model \i max = u opt x CM 2 (7), with CM 2 (T) given by Equation 
3.23 and with 7^ of-6°C, T x of 1 °C, T c of 12°C, 7 0Pt of 37°C, and T max of 45°C. u opt was 1 .0 lr 1 . 




cannot be used with the other type of model. This situation is similar to the 
estimation of |i max values by some primary growth models. Modi ed Gompertz 
models (Gibson et al., 1987; Zwietering et al., 1990), e.g., overestimate |i. max by 
-15% (Dalgaard et al., 1994; Membre et al., 1999; Whiting and Cygnarowicz- 
Provost, 1992) and their growth rate values should not be used together with the 
exponential, the logistic, or other Richards family of growth models that rely on 
accurate |i max values. 

Classical CPMs (Equation 3.19 and Equation 3.20) as well as square-root-type 
models describe a straight line relation between suboptimal temperatures and 



|i max (Figure 3.1 and Figure 4.4 [Chapter 4]). It has been reported by Bajard et al. 
(1996) that a different, biphasic, relationship can be observed for some strains of 
Listeria monocytogenes. More recently, Le Marc et al. (2002) observed a biphasic 
relationship for a strain of Listeria innocua. Le Marc et al. (2002) suggested an 
expanded CPM (Equation 3.23) to simulate this type of growth response (Figure 
3.3). In Equation 3.23, r c is the change temperature and T x corresponds to the T 1Xihl 
value in a classical cardinal temperature model (Rosso et al., 1993). McMeekin et 
al. (1993), however, cautioned against the interpretation of apparently continuously 
curved relationships as the combination of two linear responses, and provided a 
simple illustration of the effect. It should also be noted that other workers (e.g., 
Nichols et al., 2002a,b; see also Chapter 4) have not observed the "curvature" in 
the low temperature region of L. monocytogenes growth. 





2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 80 Friday, November 14, 2003 9:34 AM 






CM (T) = 



(T-T.) 2 (T-T ) 

v l 7 v max 7 



(T -T)-[(T -T)-(T-T f )-(T -T )(T . + T -2T)] 

v opt l 7 LV opt l 7 v opt-' v opt max 7 v opt 1 7J 

(T -T) 2 (T-T ) 

v c l 7 v max 7 



,T<T<T 



in ax 



- / J* _J- \ 2 

min 



<r„ - r,) ■ [(r w - r,) -(T e - r ,) - ( r , - r„) • (r , + r, - 2 ■ r c )] 



opt 



opt 



opt 



opt 



T-T 

\ c min J 



mm 



, T . <T<T 

(3.23) 



As stated above, general CPMs rely on the assumption that different environ- 
mental parameters have independent and thereby multiplicative effects on (i max 
(Equation 3.19). The successful use of many general CPMs and square-root-type 
models has shown this assumption to be reasonable for wide ranges of environmental 
conditions. However, numerous studies have shown that the growth range of a 
microorganism to one environmental condition is affected by other environmental 
factors (see Section 3.4). This suggests that the predictive accuracy of general CPMs 
can be improved by taking into account interactions between environmental param- 
eters, particularly where one factor is sufficiently stringent that it reduces the growth 
range of the organism in other environmental "dimensions." 

Various approaches have been suggested to describe growth limits under the 
influence of multiple variables (see Section 3.4.4). Two such approaches have been 
suggested for direct incorporation in CPMs and are discussed briefly here. Augustin 
and Carlier (2000b) developed a global secondary model for L. monocytogenes, 
including terms for interactions that prevented growth. Absolute minimal cardinal 
values X^ [n were estimated by assuming that all inhibitory substances were absent. 
Similarly, absolute minimal inhibitory concentrations MIC® were estimated for 
optimal concentrations of other environmental parameters (X = X opt ). Then, interac- 
tion between environmental parameters was taken into account by modifying each 
of the X° in values (Equation 3.24) and the MIC® values, depending on levels of 
other environmental parameters. After calculation of appropriate r min , a w min , pH min , 
and MIC f values, growth rates were then predicted by using Equation 3.19 to 
Equation 3.21. 




r 



a 



n 



X ■ =X -(X -X°. ) 

mm opt x opt min 7 



1- 



C. 



\ 



( 



V 



?=1 



MIC. 



' / 



Y -Y 

opt 

-0 



\ 



y - r u 

. opt min j 



r v^ 1/3 

Z -Z 

opt 

z -z°. 

V opt mill j 



(3.24) 



with X, 7, and Z being temperature, pH, or water activity. 

A different approach was used by Le Marc et al. (2002) to model the interactive 
effects of temperature, pH, and concentration of undissociated organic acids (HA) 
on growth of Listeria innocua. Cardinal parameter values were kept constant and 
the space of environmental factors was divided into (1) the independent effect space 
(^ = 1), (2) the interaction space (0 < ^ < 1), and (3) the no growth space (^ = 0) 
(Equation 3.25 to Equation 3.27). 




2004 by Robin C. McKellar and Xuewen Lu 











~V 



1237_C03.fm Page 81 Wednesday, November 12, 2003 12:40 PM 





M max = K B , ■ CM i CO ■ CM, {pH) ■ x{\HA\) ■ %(T, P H, [HA]) 



(3.25) 



Z,(<tfT,pH,ULA)) = 



1 y<0 

2(1 - v) 9 < y < 1 
V|/>1 



(3.26) 



¥ = 



I 



<P, 



2n (i -v 



(3.27) 






with (p r =(W CM 2CO)-> <? pH = {\-CM l (pH))\im& ^ UndissocjatedLacticAcid(JJLA) 
= 1 - (ULA/MIC ULA )) and where e i are the environmental factors. For calculation of 
CM 2 (T) and CM^pH), see Equation 3.20. Le Marc et al. (2002) selected a value of 
0.5, which was used for 9. 

The performance of the two approaches to model interaction between environ- 
mental parameters is considered in greater detail in Section 3.4.4. As shown above, 
CPMs that take into account the effect of interaction between environmental param- 
eters are relatively complicated models. Thus, these models are not fully in agree- 
ment with the originally cardinal parameter modeling approach, i.e., that CPM uses 
only simple biological meaningful parameters that microbiologists are familiar with 
and that are easy to use by biologists (Rosso et al., 1993), and raises questions about 
whether those models are the most parsimonious forms available. 

The model suggested by Augustin and Carlier (2000b) predicts the effect of 
interaction between temperature, pH, and lactic acid concentration on growth of 
Listeria monocytogenes to be more pronounced than the effect predicted for Listeria 
innocua by the model of Le Marc et al. (2002). For example, the Augustin and 
Carlier (2000b) model predicts no growth of Listeria monocytogenes at 8°C, pH 
6.0, and with 200 mM of lactic acid, whereas at this condition the model of Le Marc 
et al. (2002) predicts growth and also that there is no interactive effect of the 
environmental factors (t > = 1). Recently, Gimenez and Dalgaard (in press) found the 
model of Augustin and Carlier (2000b) to substantially underestimate growth of 
Listeria monocytogenes in cold-smoked salmon. This could indicate that the model 
is in fact overestimating the importance of the interaction between at least some sets 
of environmental factors. 

In a similar vein Ratkowsky and Ross (1995), recognizing the relationship 
between absolute limits for each environmental factor and their relationship to the 
parameters of square-root-type models and CPMs, experimented with the use of a 
kinetic model as the basis of a growth boundary model using linear logistic regres- 
sion. This approach is discussed later (see Section 3.4.3.2). 

The classical CPMs, in particular those including the effect of temperature, water 
activity, or pH, are now popular and used for many purposes within predictive 
microbiology (see Table 3.5 and Table 3.6). As one example a cardinal temperature 
and pH model has been combined with classical models of microbial kinetics, i.e., 

2004 by Robin C. McKellar and Xuewen Lu 












~V 






1237_C03.fm Page 82 Wednesday, November 12, 2003 12:40 PM 






models that rely on yield factors and maintenance constants. In this way, production 
of curvacin A by Lactobacillus curvatus LTH 1174 growing in MRS broth was 
successfully modeled between 20 and 38°C and at pH values from 4.8 to 7.0 
(Messens et al., 2002). Other examples include the use for CPMs to predict the 
radial growth rate of molds on solidi ed laboratory media (Panagou et al., 2003; 
Rosso and Robinson, 2001; Sautour et al., 2001). The ability of these models to 
predict growth in foods deserves further study. 

For practical use of secondary predictive models it is important to know the 
precision of the predicted responses. With CPMs it has been suggested to determine 
cardinal parameters values for a number of different strains within each of the 
microbial species of interest (Membre et al., 2002). In this way a measure of intra - 
species variability can be obtained. As an example, variability in the pH min value for 
10 strains of E. coli was ±0.20 corresponding to approximately four times the 
experimental error (Membre et al., 2002). More recently Pouillot et al. (2003) 
suggested the use of a CPM together with a Bayesian procedure for parameter 
estimation. This approach includes the use of hyperparameters and allows uncer- 
tainty (due to imperfect knowledge or data) and true variability (e.g., due to differ- 
ence between strains) to be determined separately (see also Chapter 4). The approach 
seems most interesting and de nitely deserves to be studied further for different 
secondary predictive models. 

3.2.3.1 Secondary Lag Time Models and the Concept 
of Relative Lag Time 

When exponentially growing microorganisms are transferred from one environment 
into another, similar environment, growth usually continues without delay, i.e., a lag 
time is rarely observed. However, when the two environments differ, a lag time is 
often observed. Similarly, when microorganisms in the lag or stationary phases are 
transferred into identical or new environmental conditions a lag time may continue 
or result, respectively. Depending on the physiological state of the microorganisms, 
the magnitude of the shift in the environmental conditions, and the new environmental 
conditions themselves, the duration of the lag time may range from to in nity. 

Development of secondary lag time models is complicated by the fact that lag 
time is in uenced not only by the actual environmental conditions but also by 
previous environmental conditions and the physiological status of the cell, i.e., the 
growth phase of microorganisms at the time of transfer between environments and 
their "enzymatic readiness" to exploit the specific carbon and energy resources 
within the new environment. Within predictive microbiology, two main approaches 
have been used for development of secondary lag time models: (1) models where 
lag time and growth rate are modeled independently and (2) models where lag time 
is assumed proportional to the generation time. The latter group of models typically 
rely on the assumption that microorganisms need to perform a given amount of work 
to adapt to a new environment and that the rate at which this work can be done 
depends on the growth rate potential of the organism in the new environment 
(Robinson et al., 1998). 




2004 by Robin C. McKellar and Xuewen Lu 











~V 






1237_C03.fm Page 83 Wednesday, November 12, 2003 12:40 PM 






In the former approach, lag times or lag rates (i.e., reciprocal of lag time) are 
typically log-transformed to stabilize the variance of these data. Frequently, poly- 
nomial models (see Section 3.2.5) or artificial neural networks (see Section 3.2.6) 
have been used to develop independent secondary lag time models (Table 3.5). To 
model the effect of temperature downshifts, temperature upshifts, and physiological 
status of cells (e.g., exponential phase, stationary phase, starved, frozen, dried), 
separate polynomial models have been used for the different physiological conditions 
(Whiting and Bagi, 2002). When square-root-type and Arrhenius-type models are 
used for lag time modeling, lag rates are modeled or reciprocal forms of the growth 
rate models are used (see Section 3.2.1 and Section 3.2.4; Table 3.5 and Table 3.6). 

Zwietering et al. (1994), e.g., used a square-root model (Equation 3.2) with 
identical values of the parameters T min , c, and T msx to model lag time and growth 
rate — only the value of b differed between the two models. Specific secondary lag 
time models for particular environmental parameters have also been suggested, e.g., 
a hyperbola model for the effect of temperature (Equation 3.28; Oscar, 2002; Zwi- 
etering et al., 1994): 



X = 



P 



T-q 



(3.28) 



where X is the lag time, T the temperature, p the rate of change of lag time as a 
function of temperature, q the temperature at which lag time is in nite, and m is an 
exponent to be estimated. 

Baranyi and Roberts (1994), Smith (1985), and McMeekin et al. (1993) have 
observed that lag times for identical inocula introduced to (at least some) envi- 
ronmental conditions are inversely proportional to growth rates and thus propor- 
tional to generation times (T gen ). This generalization has limits, however, as dis- 
cussed further below and probably is most relevant to changes in environmental 
temperature . For example, Zwietering et al. (1994) showed that for the effect of 
temperature on Lactobacillus plantarum the product of |i max and lag time (k) was 
constant and had an average value close to 2. In these situations secondary lag 
time models can be derived directly from a growth rate model by using the simple 
concept of relative lag time (RLT; Equation 3.29) in common use but first defined 
by Mellefont and Ross (2003). Clearly, RLT reflects the physiological status of 
microorganisms introduced into a new environment as well as the difference 
between their actual and their previous environments, and can be interpreted as 
the amount of work the cell has to do to change its physiology (e.g., enzymes, 
membrane composition, number of ribo somes) to be able to grow at |i max in that 
new environment. 

Baranyi and Roberts (1994) suggested a primary model to estimate lag times 
from microbial growth curves and this model allowed determination of the para- 
meters h , q , and a all of which reflect the physiological state of microorganisms 
and, thereby, their readiness to grow in a given environment (Equation 3.29; Chapter 
2). It can be seen that the parameter RLT is directly proportional to h . 




2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 84 Wednesday, November 12, 2003 12:40 PM 







^ = rlt x = **>t.W2) 




gen r 1- max 

(3.29) 

( 1 "\ 
X-ii =RLT-\n(2) = h = In 1 + — =-ln(oc ) 

"max v y o v o J 



where all parameters have meanings as indicated earlier. 

Experimental methods to determine the physiological status of low levels of 
microorganisms in foods remain to be developed. Thus, for the time being these 
parameters have mainly theoretical importance. 

The RLT concept is practically very useful for development of secondary lag 
time models, but it should be used with caution. Delignette-Muller (1998) eval- 
uated data from nine studies where the effect of temperature, pH, NaCl, and 
NaN0 2 on lag time and generation time on different food-borne microorganisms 
had been modeled independently. In four of the nine studies, RLT was constant 
and an independent lag time model was not needed. However, primarily pH and 
NaCl in uenced RLT in the remaining studies. On the basis of large amounts of 
experimental data, Ross (1999a) showed the distribution of RLT of B. stearother- 
mophilus, Clostridium perfringens, E. coli, L. monocytogenes, Salmonella, and 
S. aureus included peaks in the range 3 to 6 under a very wide range of experi- 
mental conditions. These distributions were similar to those presented by Augus- 
tin and Carlier (2000a), who observed a median RLT of 3.09 for L. monocytogenes 
(n = 1176). Using extreme environmental shifts, and severely growth-limiting 
outgrowth conditions, the hypothesis that RLT values have an upper limit was 
tested (Mellefont et al., 2003, in press). It was found that most RLTs were in the 
range 4 to 6, and that RLTs greater than 8 could not be induced within the 
experimental system employed. These observations suggest that while lag time 
is apparently highly variable, RLT is more uniform and reproducible. Distribu- 
tions of RLT can be used in stochastic modeling studies, for example, microbial 
food safety risk assessments, where they could be used as plausible default 
assumptions if specific lag time information was not available. This approach can 
also simplify the growth modeling process because use of the RLT as a variable 
enables the effects of growth rate and lag to be predicted by a single growth rate 
model, as explained above. 

The RLT concept implies that X is at a minimum value (A, min ) when the growth 
rate is optimal (|i opt ). This relation has been used together with CPMs to obtain 
simple secondary lag time models (Equation 3.30 and Equation 3.31; Augustin 
and Carlier, 2000a; Le Marc et al., 2002; Pouillot et al., 2003; Rosso, 1995, 
1999a,b). 



X . -LI , 

X= — opt (3.30) 

" max 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 85 Wednesday, November 12, 2003 12:40 PM 








X = ™ (3.31) 

CM 2 (J) ■ CM 2 (0" CM (pH) 

For RLT models to be used in practice it must be known if, and to what extent, 
abrupt or smooth shifts in environmental parameters like temperature, pH, and water 
activity in uence RLT. 

Data presented by Rosso (1999a,b) suggested that the effect of shifts in temper- 
ature and pH on growth of E. coli during fermentation of yoghurt was appropriately 
predicted by a CPM that relied on assumption of a constant RLT. Augustin et al. 
(2000) suggested a model to take into account the effect of growth phase and 
temperature history of L. monocytogenes on its RLT. For temperature downshifts 
the RLT increased from ~0 for a temperature shift of 0-5 °C to ~2 for a downshift 
of 30-3 5°C. To model the effect of temperature downshifts and upshifts on RLT of 
L. monocytogenes, Delignette-Muller et al. (2003) recently used the data of Whiting 
and Bagi (2002) and suggested simple biphasic linear models. Separate models were 
used for inoculum with different physiological states. For E. coli, Mellefont and 
Ross (2003) found a similar effect of temperature downshifts whereas temperature 
upshifts had no systematic effect on RLT. For abrupt downshifts and upshifts in 
water activity the data of Mellefont et al. (2003) suggest that simple biphasic linear 
models, with different slopes for down- and upshifts, may be appropriate to predict 
RLT of both Gram -negative and Gram-positive food-borne bacteria. The universality 
of these responses remains unclear. For example, RLTs of S. aureus and L. mono- 
cytogenes were largely unaffected by abrupt osmotic shifts over a wide range of salt 
concentrations, whereas RLT of Gram-negative cells was strongly affected. More 
research is required before models that are as reliable as existing growth rate models 
can be developed for lag time, or RLT. 

3.2.4 Secondary Models Based on the Arrhenius Equation 
3.2.4.1 The Arrhenius Equation 

The empirical Arrhenius-van't Hoff relationship: 

rate = A exp(AE a I RT) (3.32) 

or its mechanistic interpretation and modi cation due to Eyring (1935) based in 
absolute reaction-rate theory: 

rate = KTexpiAH* I RT) (3.33) 

where the parameters may be interpreted as follows: A is a constant related to the 
number of collisions between reactants per unit time, E a the activation energy, R the 
gas constant (8.314 J/K/mol), T the temperature in Kelvin, K is similar to A but 
includes steric and entropic effects, and A// J is the enthalpy difference between the 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 86 Wednesday, November 12, 2003 12:40 PM 







transition state complex and the reactants, are well established in chemistry to 
describe the effects of temperature on the rate of chemical reactions. Taking the 
logarithm of both sides of Equation 3.32: 

In (rate) = ln(A) x AE/RT 

and reparameterizing the equation becomes: 



In (rate) = A' + 



( 



V 



AE 
R 



\ 



J 



x 



r i a 



\Tj 



Thus, if ln(rate) is plotted against 






the resulting plot is a straight line over 



temperature ranges relevant to microbial growth and allows estimation of the "acti- 
vation energy" of the reaction, as shown in Figure 3.4. The activation energy can 
be used to characterize the reaction. 

Temperature (°C) 

76.84 66.84 56.84 46.84 36.84 26.84 16.84 6.84 -3.16 -13.16 -23.16 
6 A 1 1 1 1 ! 1 1 ; 1 1 1- 






0.00286 



0.00306 0.00326 0.00346 0.00366 

1/(Temperature [K]) 



0.00386 



FIGURE 3.4 Diagram showing the effect of temperature on reaction rate predicted using the 
Arrhenius model (Equation 3.33; solid line) and the effect of temperature on microbial growth 
rate (dashed line) for a representative mesophilic organism. The "activation energy" is esti- 
mated from the slope of the solid line, multiplied by the universal gas constant. Over a narrow 
range of temperatures, the microbial growth rate follows the Arrhenius model prediction 
(Equation 3.29). This range has been termed the "normal physiological range" (NPR). At 
temperatures above or below the NPR, microbial growth rate deviates markedly from that 
predicted by the Arrhenius model. 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 87 Wednesday, November 12, 2003 12:40 PM 




~V 









It has been argued that because all life processes are the result of chemical 
reactions, the growth rate of organisms that cannot achieve thermal homeostasis 
should also be described by Arrhenius kinetics. Within a narrow range of temperature 
this is true. In practice, however, when microbial growth rate data for the full 
biokinetic temperature range are presented as an Arrhenius plot, the data are far 
from linear as shown in Figure 3.4, and con rmed by numerous studies (Heitzer et 
al., 1991; McMeekin et al., 1993; Schoolfield et al., 1981). 

A range of secondary models, based on adherence to the reaction kinetics 
described by the Arrhenius model, but including terms to account for the observed 
deviations, have been proposed. These models fall into two groups: 

1. Those based on putative mechanistic modifications of the Arrhenius 
models 

2. Those based on empirical modifications 

3.2.4.2 Mechanistic Modifications of the Arrhenius Model 

Models in this category include those of Johnson and Lewin (1946) to describe the 
high-temperature growth of bacteria, Hultin (1955) to describe rates of enzymatic 
catalysis in the low temperature region, Sharpe and DeMichele (1977) who syn- 
thesized these two equations to produce a model for the temperature dependence 
of bacterial growth rate in the entire biokinetic region, the model of Schoolfield et 
al. (1981), which is a reparameterization of the Sharpe and DeMichele model to 
overcome difficulties in tting by nonlinear regression, and the models of 
McMeekin et al. (1993) and Ross (1993a, 1999b). The latter models incorporate 
contemporary knowledge of the thermodynamics of protein folding to overcome 
failures in the Schoolfield et al. model related to unrealistic parameter estimate 
(Ratkowsky et al, 1991). 

The above models were originally developed to provide an interpretation of 
microbial growth rates or enzyme-catalyzed reaction rates, in response to tempera- 
ture but their mechanistic basis makes them attractive for use as secondary models. 

This class of secondary models have previously been reviewed (McMeekin et 
al., 1993; Ratkowsky et al., 1991; Ross, 1999b; Ross and McMeekin, 1994). In 
summary, all of the models are based on the assumption that there is a single, 
enzyme-catalyzed, rate-limiting reaction in any microorganism. This reaction is 
characterized by an activation energy, which governs the rate of reaction in response 
to temperature, according to Arrhenius kinetics. Enzymes are proteins, however, and 
are themselves subject to the effects of temperature. The functional activity of 
enzymes is dependent upon their shape, or conformation, but they are flexible — 
the flexibility being required to achieve their catalytic function. Because temperature 
affects the bonds in the molecule, if the temperature changes too much, the confor- 
mation becomes so disrupted that denaturation takes place, both at high and low 
temperatures. These denaturation events are reversible, but at high temperatures if 
the temperature increases suf ciently, irreversible denaturation takes place (Ross, 
1999b). Thus, these models include terms to model the probability, as a function of 
temperature, that the enzyme is in its metabolically active conformation and use this 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 88 Wednesday, November 12, 2003 12:40 PM 







estimate to modify the predictions of the Arrhenius model. Equation 3.34 to Equation 
3.36 are examples of this form of model. 
Model of Hmshelwood (1946): 



rate = A, cx V (-EJRT)-A 2 cx V (-E ahigh /RT 



(3.34) 



where R,T,A, and E Q have the same meaning as above. E aUgh is the activation energy 
of the high-temperature denaturation of the rate-limiting enzyme. 
Model of Schoolfield et al. (1981): 



T 



(25) 



298 



exp- 



HJ I 



R 



1 



\ 



,298 T j 



K 



1 + exp 




T 



\ 



J 



+ exp 




T 



\ 



J 



(3.35) 




where T is the absolute temperature, R is the universal gas constant, and, for modeling 
bacterial growth, the other parameters have been interpreted as follows: K is the 
response (e.g., generation) time, p a scaling factor equal to the response rate (l/K) 
at 25°C, H A the activation energy of the rate-controlling reaction, H L the activation 
energy of denaturation of the growth-rate-controlling enzyme at low temperatures, 
H H the activation energy of denaturation of the growth-rate-controlling enzyme at 
high temperatures, T V2 the lower temperature at which half of the growth-rate- 

Li 

controlling enzyme is denatured, and T l/2 is the higher temperature at which half 
of the growth-rate -controlling enzyme is denatured. 
Model of Ross (1999b): 




rate = 



CT exp(A/f ! / RT) 

1 + exp(-n(AH * -TAS* +AC [(T -T+ H )- T\n(T I T * g )]) / RT) 



(3.36) 



where C is a parameter whose value must be estimated, AH J the activation enthalpy 
of the reaction catalyzed by the enzyme controlling the overall reaction rate, AC p 
the difference in heat capacity (per mole amino acid residue) between the native 
(catalytically active) and denatured state of the enzyme, T H * the temperature (K) at 
which the AC p contribution to enthalpy is 0, T 5 * the temperature (K) at which the 
AC p contribution to entropy is 0, AH * the value of enthalpy at T H * per mole amino 
acid residue, AS* the value of entropy at T s * per mole amino acid residue, T the 
temperature (K), R the gas constant (8.314 J/K/mol), and n is the number of amino 
acid residues in the enzyme. 

Equation 3.34 to Equation 3.36 include the simple Arrhenius model in the 
numerator of the equation. The denominators in Equation 3.35 and Equation 3.36, 
however, model the probability that the enzyme is in its active conformation. When 

2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 89 Wednesday, November 12, 2003 12:40 PM 







4- 



0) 

o 

i- 
U) 

Q) 

+■» 

en 



1 ■ 







260 



! 

i 
i 


i : 

i \ 

i 

i I 

i • 

i 1 
i ; 
' / 

/ 

\ /© / 

/ A 



■■ 4 



■■ 3 



o 

+■» 

(0 

c 

£ 

o 

c 
a> 
Q 



2 © 





■■ 1 



270 



280 290 300 

Temperature (K) 



310 




320 



(0 

> 




FIGURE 3.5 Diagram showing the interaction of terms in mechanistic models for microbial 
growth rate response to temperature. Curve 1 (dashed line) is the predicted growth rate in the 
absence of master enzyme denaturation, i.e., Arrhenius kinetics as modeled by the numerator 
of Equation 3.36. Curve 2 (dot-dash line) is the inverse of the probability of the "master 
enzyme" being in the active conformation, i.e., the denominator of Equation 3.36. Curve 3 
(solid line) is the overall predicted rate from the model, i.e., the quotient of values in Curve 
1 divided by values in Curve 2. 

that probability is high, the denominator takes values close to 1 , so that the overall 
rate is close to that predicted by the Arrhenius equation in the numerator. When the 
probability is lower, the value of the denominator increases, so that the observed 
rate is lower than that predicted by the numerator alone. These relationships are 
shown in Figure 3.5 for Equation 3.36, presented as rate vs. temperature for clarity 
of interpretation. 

In practice, few of these types of models have been routinely applied in predictive 
microbiology, possibly because the models are highly nonlinear, and initial parameter 
estimates are dif cult to determine. Furthermore, it is currently not possible to 
independently measure the values of the parameters of the model because the putative 
master reaction has not been identified, and the concept that a single reaction is rate 
limiting under all environmental conditions seems improbable (Daughtry et al., 1 997; 
Ross, 1999b). Finally, several workers (Heitzer et al., 1991; Ratkowsky, D.A., 
personal communication, 2003; Ross, 1993b) demonstrated that even with good 
quality data, square-root-type models provide an equally good fit as those "mecha- 
nistic" models, and are usually easier to work with. Examples of their use include 
Broughall et al. (1983) and Broughall and Brown (1984) who used the Schoolfield 
model, but also extended it to model the effect of water activity and pH, by replacing 
some terms in the model with polynomial expressions in a w and pH. Adair et al. 
(1989) used a reparameterized form of the Schoolfield et al. model, which was 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 90 Wednesday, November 12, 2003 12:40 PM 







essentially another form of the Johnson and Lewin (1946) model. Recent studies 
(Ratkowsky et al., unpublished) have con rmed that Equation 3.36 does describe 
bacterial temperature-growth rate curves well for a wide range of species, and, in 
contrast with earlier models, that the estimated parameter values are realistic and 
consistent with the theoretical bases of the model. 

3.2.4.3 Empirical Modifications of the Arrhenius Model 

A second class of Arrhenius-based models for growth rate and reciprocal of lag time 
have been presented by Davey and coworkers. Davey (1989) introduced an Arrhe- 
nius-type model for the effects of temperature and water activity, which is linear 
and thus allows for explicit solution of the optimum parameter values. This model 
has the form: 



c a 



In (rate) = C + -J- + -f + Cm + C A a* (3.37a) 




T T 

where T is temperature (K), a w has its usual meaning, and C , C 1? C 2 , C 3 , C 4 are 
coef cients to be determined. 

Davey (1989) reported that the model described well seven data sets from the 
literature and subsequently demonstrated the ability of the model to also describe the 
reciprocal of lag phase duration (Davey, 1991). Davey (1994) fitted a variation of the 
model to the data of Adams et al. (1991) for Yersinia enterocolitica growth. The model 
included terms for temperature and pH, and is analogous to Equation 3.37a: 




C C 



In (rate) = C + -± + -± + C 3 pH + C 4 pH 2 (3.37b) 

where T is temperature (K), pH has its usual meaning, and C , C l5 C 2 , C 3 , C 4 are 
coef cients to be determined. 

On the basis of these observations, Davey (1994) extended his earlier proposed 
general model structure for linear Arrhenius models (Davey, 1989) to account for the 
effect of multiple environmental factors affecting growth rate to the following form: 



j 
In (rate) = C + ]T (C 2 ,._^ +C 2 2 ,r) (3.37c) 

1=1 

where j environmental factors, V, act in combination to affect the growth of the 
modeled organism, and C , C 1? C 2 , ... , C. are coefficients to be determined. 

This general form was applied by Davey and Daughtry (1995) to data of Gibson 
et al. (1988) for Salmonella growth in response to temperature, NaCl, and pH. Thus, 
their equation had the form: 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 91 Wednesday, November 12, 2003 12:40 PM 




~v 









c c 

ln(rate) = C + -^ + -|- + C 3 S+ C 4 S 2 + C 5 /?# + C 6 /?# 2 (3.37d) 

where S is salt concentration (% w/v). 

While the above model forms are empirical, they also recognize implicitly the 
temperature dependence of microbial growth rates. Daughtry et al. (1997) invoked 
chemical reaction rate theory to develop an alternative mechanistic model based on 
the Arrhenius equation. Those workers cited Levenspiel (1972) as stating that cur- 
vature in Arrhenius plots can arise if there are two, or more, reactions that "compete" 
to limit the reaction rate and dominate under different conditions so that the overall 
effect of temperature is the synthesis of the individual activation energies for the 
rate-limiting reactions at different temperatures. Daughtry et al. (1997) considered 
that bacterial growth was likely to be such a system. 

By assuming that the "heat of reaction" (equivalent to the activation energy or 
activation enthalpy in the above discussion) is a polynomial function of temperature, 
the following modi ed Arrhenius model was developed: 

C 
ln(rate) = C + -^ + C 2 lnT (3.38) 

This model fitted experimental data as well as the temperature-only form of Equation 
3.37a. 

The "linear Arrhenius" or "Davey" models have been used to model growth of 
molds on solid microbiological media (Molina and Giannuzzi, 1999; Panagou et al., 
2003). Panagou et al. (2003) preferred cardinal parameter and gamma-concept-type 
models (see Sections 3.2.2 and 3.2.3) over the Davey model because of their inter- 
pretable parameter values. Davey models have also been applied to UV and thermal 
inactivation and data describing the combined effects of pH and water activity on 
thermal inactivation, including vitamin denaturation (see Section 3.3), but they have 
not been widely adopted by other workers. McMeekin et al. ( 1 993) and Davey (200 1 ) 
identi ed a close correlation between estimates of coef cients C 1 and C 2 , and C 3 
and C 4 , of Equation 3.37a, suggesting that the model was overparameterized. 

3.2.4.4 Application of the Simple Arrhenius Model 

For the entire biokinetic temperature range, growth rates of microorganisms are 
described less appropriately by the Arrhenius-type equations (Equation 3.34 to 
Equation 3.36) than by square-root-type and cardinal parameter models (see Section 
3.2.4.2; Rosso et al., 1993; Zwietering et al., 1991). However, Arrhenius-type models 
remain useful as secondary kinetic models for less extensive ranges of storage 
temperatures (Table 3.5 and Table 3.6). Koutsoumanis and Nychas (2000) used 
Equation 3.32 to model the effect of temperatures between and 15°C on n max and 
reciprocal lag time of naturally occurring pseudomonads growing aerobically on a 
type of Mediterranean sh. Koutsoumanis et al. (2000) also expanded the classical 



2004 by Robin C. McKellar and Xuewen Lu 












~V 



1237_C03.fm Page 92 Wednesday, November 12, 2003 12:40 PM 






Arrhenius model to take into account the combined effect of temperature and C0 2 
on growth rates of spoilage bacteria in modified atmosphere packed fresh fish 
(Equation 3.39). 



Infix ) = 

^r max / 



E 

a_ 

R 



t 



x 



1 



V re f 



J. 
T 



\ 



+ ln(a ,-dL, x%CO.) 



CO, 



(3.39) 



/ 



where T, E a , and R have their usual meaning, %C0 2 is the equilibrium concentration 
of C0 2 in the headspace gas, d co is a constant expressing the effect of C0 2 on 
Mmax an d T Kf and |i Tef are temperature and maximum specific growth rate, respectively, 
at 273 K and %C0 2 . The term including d co in Equation 3.39, describing C0 2 
inhibition of growth rate, was previously suggested by Kalina (1993). 

The simple Arrhenius model has also been used to calculate relative rates of 
spoilage (RRS) (Equation 3.37). RRS for a food product is de ned as the shelf life 
(determined by sensory evaluation) at a reference temperature (r ref ) divided by the 
shelf life observed at the actual storage temperature (Equation 3.40). 




RRS = 



Shelf life at T 



ref 



Shelf life at T 



= exp 



E 



( 



A 



R 



x 



1 1 



\ 



V 



T T 



ref 



J 



(3.40) 




where T, E a , and R have their usual meaning and T ief is a reference temperature at 
which the shelf life is known. 

RRS models are interesting because they enable shelf life to be predicted at 
different temperatures and for products where the specific spoilage organisms or the 
type of reaction responsible for spoilage are not known. For an unusually tempera- 
ture-sensitive modified atmosphere packed shrimp product (E a > 100 kJ/mol), Equa- 
tion 3.40 described the effect of temperature (0 to 25 °C) on shelf life more appro- 
priately than a similarly formulated RRS model relying on the square-root model 
(Equation 3.1). However, a simple exponential RRS model was as useful as Equation 
3.40. That the Arrhenius and exponential RRS models performed better than the 
square-root model was due to the fact that different groups of microorganisms were 
responsible for spoilage at low and high storage temperatures, respectively (Dalgaard 
and Jorgensen, 2000). This situation is common and a reason why entirely empirical 
RRS models can be more appropriate for shelf-life prediction than kinetic models 
relying on growth of known spoilage microorganisms. In fact, kinetic models for 
growth of spoilage bacteria are generally useful only for shelf-life prediction within 
the spoilage domain of a specific microorganism (Dalgaard, 2002). 



3.2.5 Polynomial and Constrained Linear 
Polynomial Models 

Of the types of secondary models applied within predictive microbiology polyno- 
mial models are probably the most common. As shown in Table 3.5 and Table 3.6, 





2004 by Robin C. McKellar and Xuewen Lu 











~V 






1237_C03.fm Page 93 Wednesday, November 12, 2003 12:40 PM 






the effect of many different environmental parameters (e.g., temperature, NaCl/a^ 
pH, nitrite, C0 2 , organic acids, and natural antimicrobials) has been described by 
these linear models. Polynomial models were extensively used during the 1990s 
and they remain widely applied although square -root-type and CPMs are becoming 
increasingly popular (Table 3.5). Polynomial models are attractive, rst, because 
they are relatively easy to fit to experimental data by multiple linear regression, 
which is available in most statistical packages. Second, polynomial models allow 
virtually any of the environmental parameters and their interactions to be taken 
into account. Thus, application of polynomial models is a simple way to summarize 
information from a data set. Once the coefficients in a polynomial model have been 
estimated, the information is easy to use particularly if the model is included in 
application software. In fact, the application software packages Pathogen Modeling 
Program and Food MicroModel rely primarily on the use of polynomial models 
(www.arserrc.gov/mfs/pathogen.htm; Buchanan, 1993a; McClure et al., 1994a). 

To illustrate the use of polynomial models a quadratic equation used by McClure 
et al. (1993) is shown below (Equation 3.41): 

\ny =p x + p 2 x { + p 3 x 2 + p 4 x 3 + p A x l x 1 + p 6 x^x 3 

(3.41) 

+p 7 x 2 x 3 + p s x 7 ~ + p 9 x\ + p l0 x 7 - + e 



"max' 



where In y denotes the natural logarithm of the modeled growth responses (y = (J, 
lag time or maximum population density [MPD], or the modified Gompertz model 
parameters B or AT): p, (i= 1, ..., 10) are the coefficients to be estimated; x 1 is the 
temperature (°C); x 2 is the pH; x 3 is NaCl (% w/v); e is a random error supposed to 
have zero mean and constant variance. 

As shown by Equation 3.41 the same polynomial equation can be used to model 
different microbial growth responses. Actually, many studies have modeled the effect 
of environmental conditions on specific parameters in primary growth models, par- 
ticularly B, M, and C in the modified Gompertz model. Measures of lag time, growth 
rate, or time for, e.g., a 1000-fold increase in the cell concentration are then calculated 
at specific environmental conditions from the predicted value of B, M, and C (Bucha- 
nan and Phillips, 2000; Eifert et al., 1997; Skinner et al., 1994; Zaika et al., 1998). 
Growth responses to be modeled are typically In- or log 10 -transformed (Equation 
3.41) and it is common practice to transform the growth response without trans- 
forming the model. 

However, polynomial models have properties that limit their usefulness as 
secondary predictive models. Polynomials include many coefficients that have no 
biological interpretation. As an example, Equation 3.41 uses 10 coefficients to model 
the effect of three environmental parameters. With four environmental parameters, 
polynomials with 15 coefficients are frequently used. The high number of coeffi- 
cients and their lack of biological interpretation make it dif cult to compare poly- 
nomial models with other secondary predictive models. The important information 
included in, e.g., the T min parameter of a square -root -type model, is not provided 
by a polynomial model. 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 94 Wednesday, November 12, 2003 12:40 PM 




~V 









Higher order polynomial models, e.g., cubic or quadratic models have been 
criticized for being too flexible and for attempting to model, rather than eliminate, 
experimental error (Chapter 4; Baranyi et al., 1996; Sutherland et al., 1996). Because 
of the very flexible nature of higher order polynomial models they should not be 
used as secondary models within predictive microbiology unless very high quality 
experimental data are available and support the application of these models. Fur- 
thermore, because quadratic polynomial models are highly flexible they should only 
be used to provide predictions by interpolation. Baranyi et al. (1996) pointed out 
that the interpolation region of a polynomial model is the minimum convex poly- 
hedron (MCP) defined by the ranges of the environmental parameters used to develop 
the model, i.e., the experimental design. These authors also stressed that the inter- 
polation region (Figure 3.10) can be substantially smaller than the rectangular 
parallelepiped whose sides are given by the endpoints of the ranges of environmental 
parameters, termed the "nominal variable space" (Baranyi et al., 1996). 

Determination of the interpolation region of a polynomial model is not self- 
evident and requires information about ranges of the environmental parameters used 
to develop the model. Pin et al. (2000) suggested a method to determine if a specific 
environmental condition is inside or outside the interpolation region of a particular 
polynomial model. This method relies on the iterative algorithm used by the Solver 
add-in of Microsoft Excel and thus is readily accessible to many users. However, 
we believe for it to become widely used the calculation of interpolation regions 
should be included in dedicated predictive modeling application software. 

To overcome the problem that quadratic polynomial models can be too exible, 
and therefore in some situations provide predictions that are not logical, the appli- 
cation of constrained polynomial models was recently suggested (Geeraerd et al., 
in press). With this approach, the basic idea is to combine a priori information about 
the effect of environmental parameters on growth responses with classical polyno- 
mial models. For example, at suboptimal conditions it was assumed that the growth 
rate should always increase for increasing temperature and a w values and decrease 
for increasing C0 2 levels. Thus, the partial derivative of the model with respect to 
temperature and a w should always be positive whereas the partial derivative of the 
model with respect to C0 2 should always be negative. Coef cients of the polynomial 
model were then fitted with the constraints obeyed at all edges of the experimental 
design. The constrained polynomial model was fitted by the usual process of min- 
imizing the sum of squared errors and the tting was carried out using the Optimi- 
zation Toolbox within the MatLab software (Geeraerd et al., in press). As compared 
to classical polynomial models, constrained polynomial models have the clear advan- 
tage of being more robust but the clear disadvantage of being substantially more 
dif cult to t. Simpli cation of the tting process seems necessary before con- 
strained polynomial models nd wide application in predictive microbiology. 

Masana and Baranyi (2000a) described methods for integration of new data into 
existing polynomial models, pointing out that the interpolation region of the newly 
developed model can be unexpectedly small and also presenting methods for quan- 
tifying the increased risk of inadvertent extrapolation (Baranyi et al., 1996). Poly- 
nomial models feature many cross-product terms, making the addition of new terms 
much more complex than with models embodying the gamma concept (Section 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 95 Wednesday, November 12, 2003 12:40 PM 




~V 








3.2.2). Nonetheless, when expanding a model by the addition of data for a new 
variable, Masana and Baranyi (2000b) demonstrated that the original model can be 
retained as a special case of the expanded model, by holding the terms of the original 
model, i.e., those that do not contain the new variable, as constants during the model 
tting process for the expanded model. 

3.2.6 Artificial Neural Networks 

Arti cial neural networks (ANNs) are algorithms that can be used to perform 
complex statistical modeling between a set of predictor variables and response 
variables. Their particular advantage is that they have the potential to approximate 
underlying relationships of any complexity between those variables. They have been 
used to generate secondary models for microbial growth rates and lag times (Garcia - 
Gimeno et al., 2002, 2003; Geeraerd et al., 1998a; Jeyamkondan et al., 2001; Lou 
and Nakai, 2001; Najjar et al., 1997), growth under uctuating environmental con- 
ditions (Cheroutre-Vialette and Lebert, 2000; Geeraerd et al., 1998a), microbial 
inactivation (Geeraerd et al., 1998b), and have been proposed as an alternative to 
logistic regression modeling techniques (Tu, 1996). Their potential to replace logistic 
regression for growth limits modeling (see Section 3.4) has also been described 
(Hajmeer and Basheer, 2002, 2003 a,b) in which context they have been termed 
"probabilistic neural networks" (PNNs). 

Hajmeer et al. (1997) and Hajmeer and Basheer (2002, 2003a) describe the 
principles of ANNs and related technologies in the context of predictive microbiol- 
ogy, and numerous texts are dedicated to the subject but the following is largely 
drawn from the succinct and lucid description of Tu (1996). 

Arti cial neural networks were conceived decades ago by researchers attempting 
to reproduce the function of the human brain, i.e., its ability to learn and remember, 
but it was only in the 1980s that the "back-propagation" technique was rediscovered, 
enabling such computational systems to "learn" mathematical relationships between 
input and output variables. 

Neural networks are effectively a series of mathematical relationships between 
predictor variables ("input nodes"), a series of hidden "nodes," and an output variable 
("output node") (Figure 3.6). Each input node is related to each hidden node, and 
each hidden node is related to the output node, by some mathematical function. 
Each input is given a weight during the "training" routines, the value of each hidden 
node being the sum of a weighted linear combination of the input node values. In 
addition, bias values can be added to the weighted values of the inputs. These are 
analogous to the intercept in regression equations, while the weights are analogous 
to coef cients of the independent variables. The output node receives a weighted 
input from each of the hidden nodes, to which is often applied a logistic transfor- 
mation or other function (the "activation function") to determine the overall output. 

A set of input and corresponding output values is presented to the network, the 
error is evaluated, and the weights are then adjusted to minimize the difference 
between the predicted output and that which was observed. This process of adjust- 
ment of weights is the back-propagation step and involves algorithms based on 
complex equations. Input data are continuously presented to the neural network until 

2004 by Robin C. McKellar and Xuewen Lu 











INPUT LAYER 
and NODES 



Temperature S X 
► ( T* ) 



pH 



■► f pH* ) 





HIDDEN LAYER 
and NODES 



*w 



W 



(T, HN1) 



rafT 



W 



(aw,HN1 




OUTPUT LAYER 





>® 







ro 

I 
O 

o 

3' 

cro 
o 

< 

a 



a 
w 

a- 

pa 

c 

< 

o 

3 

cr 
o 



o 
o 



o 





FIGURE 3.6 Diagram of an imaginary arti cial neural network that might be used in predictive microbiology. The output is the response of the population 
of microorganisms to variations in the temperature, pH, and water activity of their growth medium. (The diagram is fully explained in the text.) 




2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 97 Wednesday, November 12, 2003 12:40 PM 




~V 









the overall error has been minimized, a process analogous to the iterative routines 
employed in nonlinear regression software. Optimal training algorithms can, at this 
time, only be determined empirically. Additionally, when using ANNs other elements 
of the modeling require experimentation, including the number of training cycles 
(too many can reduce predictive performance), the number of nodes in the hidden 
layer, and the ideal learning rate (the magnitude of change in the weights for each 
training case). 

In Figure 3.6, the input, hidden, and output layers are shown, as well as the 
connections between them. Nodes are represented by circles. The W^ terms indicate 
the weight applied to the inputs to hidden nodes. (Not all weights are represented 
in the diagram.) The hidden nodes have a transformation applied to them, e.g., a 
logistic function represented by the functions hi, hi, etc. Thus, in the example: 

hi = l/(l+exp(Bias 1 + W (T> HN1) x T* + W w HN1) x pH* + W (aw> HN1) x a w *)) 

and, similarly: 

Output =1/(1+ exp(B2 + W (hl) + W (h2) )) 

Suf cient data are required so that a subset of data can be used to train the ANN, 
while the remainder is used to test the predictive ability of the ANN. One complete 
cycle of the training data set is called an "epoch" and the duration of the training 
is often described as the number of epochs required to minimize the error in the 
training set. 

Tu (1996) compared the advantages and disadvantages of the ANN approach to 
those of traditional statistical regression modeling, as summarized in Table 3.4. 
Evaluation of the approach as applied to predictive models for microbial growth is 
presented below, and in relation to growth limits models in Section 3.4.2. Further 
comment is provided in Chapter 4, Section 4.4.3. 

The use of ANN in predictive growth modeling remains relatively little devel- 
oped, and direct comparison of the performance of different ANN techniques is still 
lacking. To describe growth curves, Schepers et al. (2000) concluded ANN was less 
appropriate than classical nonlinear sigmoidal growth models. Cheroutre-Vi alette 
and Lebert (2000), however, found a recurrent (i.e., back-propagation) ANN suitable 
to predict growth of Listeria monocytogenes under constant and uctuating pH and 
NaCl conditions. As shown in Table 3.5 and Table 3.6, several secondary ANN 
models have been developed including models for Aeromonas hydrophila, Brocho- 
thrix thermosphaca, Escherichia coli, lactic acid bacteria, Listeria monocytogenes, 
and Shigella exneri. These secondary ANN models have been compared with 
polynomial, square -root-type, and cardinal parameter models. The comparisons 
showed ANN typically fitted experimental data better and in most cases provided 
slightly more accurate predictions. Thus, in general, ANN may provide slightly 
improved predictions. Commercial neural network software is available and devel- 
opment of ANNs has become relatively easy. However, ANN is a data-driven 
approach and this could be a drawback because a secondary model that can be 
written as an equation with coef cients and parameters is not produced. The 

2004 by Robin C. McKellar and Xuewen Lu 












~V 



1237_C03.fm Page 98 Wednesday, November 12, 2003 12:40 PM 







TABLE 3.4 

Advantages and Disadvantages of Neural Network Approaches to Modeling 



Disadvantages 

Neural networks are a "black box" and have limited 
ability to specifically identify possible causal 
relationships between predictor and response 
variables 

Neural network models may be more dif cult to 
use in the field 

Neural network modeling requires greater 
computational resources 

Neural network models are prone to over tting 



Advantages 

Neural networks require less formal statistical 
training to develop 



Neural network models can implicitly detect 

complex nonlinear relationships between 

predictor and response variables 
Neural network models have the ability to detect 

all possible interactions between predictor 

variables 
Neural networks can be developed using multiple 

different training algorithms 
Neural network model development is empirical, 

and many methodological issues remain to be 

resolved 



Source: After Tu, J.V. J. Clin. Epidemiol, 11, 1225-1231, 1996. 



incorporation of classical secondary models in user-friendly application software 
has been essential for the usefulness of predictive microbiology in industry, teaching, 
and research. It remains to be demonstrated whether successful ANN models can, 
in a similar way, be communicated to and conveniently applied by wide groups of 
users within predictive microbiology. 




3.3 SECONDARY MODELS FOR INACTIVATION 

There are relatively few models that consider the effects of multiple environmental 
factors on the rate of death of microorganisms, and these are discussed in Chapter 
2 and Chapter 5. Some available inactivation models are also summarized in Table 
3.5 and Table 3.6. 



3.4 PROBABILITY MODELS 
3.4.1 Introduction 

Models to predict the likelihood, as a function of intrinsic and extrinsic factors, that 
growth of a microorganism of concern could occur in a food were rst explored in 
the 1970s. Those models were concerned with prediction of the probability of 
formation of staphylococcal enterotoxin or botulinum toxin within a specified period 
of time under de ned conditions of storage and product composition (Genigeorgis, 
1981; Gibson et al., 1987). Phenomena that have been modeled using this approach 





2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 99 Wednesday, November 12, 2003 12:40 PM 






TABLE 3.5 

Examples of Secondary Models for Growth of Pathogenic and 

Indicator Microorganisms 






Type of 




Microorganisms and 


Secondary 


Response 


References 


Model 


Variables 


Aeromonas hydrophila 






Palumbo et al. 


Polynomial 


GT, b lag 


(1992) a 






McClure et al. 


Polynomial 


GT, lag 


(1994b) 






Palumbo et al. 


Polynomial 


GT, lag 


(1996) a 






Devlieghere et al. 


Square-root, 


Hmax> ^g 


(2000a) 


polynomial 




Jeyamkondan et al. 


ANN 


GT, lag 


(2001) 






Aspergillus spp. 






Pitt (1995) 


Kinetic with 


Growth and 




yield factors 


a atoxin 
formation 


Molina and 


Arrhenius 


Colony growth 


Giannuzzi (1999) 






Rosso and 


CPM 


Colony growth 


Robinson (2001) 






Sautour et al. (2001) 


CPM 


Colony growth 


Bacillus cereus 






Benedict et al. 


Polynomial 


GT, lag 


(1993) a 






Sutherland et al. 


Polynomial 


GT, lag 


(1996) c 






Zwietering et al. 


Gamma 


r*max 


(1996) 






Chorin et al. (1997) 


Polynomial 


Growth rate, 
lag 


Singaglia et al. 


Polynomial 


Spore 


(2002) 




germination 


Clostridium botulinum 






Baker and 


Polynomial 


Time to toxin 


Genigeorgis 




formation 


(1990) a 






Graham et al. 


Polynomial 


GT, time to 


(1996) c 




toxin 



Independent Variables and Ranges 



r(5-42°C); NaCl (0.5-4.5%); pH 

(5.0-7.3); Na-nitrite (0-200 ppm); 

anaerobic 
T (3-20°C); NaCl (0.5-4.5%); pH 

(4.6-7.0); aerobic 
T (5-42°C); NaCl (0.5-4.5%); pH 

(5.0-7.3); Na-nitrite (0-200 ppm); 

aerobic 
r(4-12°C); a w (0.974-0.992); C0 2 

(0-2403 ppm); pH 6.12; nitrite (22 

ppm) 
Data from McClure et al (1994b) 



T; # w ;pH; and colony size: limits not 
specified in manuscript 

T (25-36°C); propionic acid 

(129-516 ppm) 
T(25, 30, 37°C); a w (0.83-0.99); pH 

(6.5); humectant: glucose/fructose 
T(25°C); a w (0.88-0.99) 

T (5-42°C); NaCl (0.5-5.0%); pH 

(4.5-7.5); Na-nitrite (0-200 ppm); 

aerobic 
r(10-30°C); NaCl (0.5-10.5%); pH 

(4.5-7.0); CO, (10-80%) 
T (10-30°C); a w (0.95-1.00); pH 

(4.9-6.6) 
T (7-30°C); a w (0.95-0.991); pH 

(5-7.5); humectant: glycerol 
T (20-40°C); a w (0.94-0.99); pH 

(4.5-6.5) 

T (4-30°C); initial spore cone. (-2 to 
+4 log cfu/g); initial aerobic plate 
count (-2 to +3 log cfu/g) 

T (4-30°C); NaCl (1.0-5.0%); pH 
(5.0-7.3) 






2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 100 Wednesday, November 12, 2003 12:40 PM 







TABLE 3.5 (Continued) 

Examples of Secondary Models for Growth of Pathogenic and 

Indicator Microorganisms 






Type of 




Microorganisms and 


Secondary 


Response 


References 


Model 


Variables 


Whiting and 


Polynomial 


Time to 


Oriente (1997) a 




turbidity 


Chea et al. (2000) 


Polynomial 


Spore 
germination 


Fernandez et al. 


Polynomial 


Time to 


(2001) 




turbidity 


Clostridium perfringens 






Juneja et al. (1996) a 


Polynomial 


GT, lag 


Escherichia coli 






Buchanan and Bagi 


Polynomial 


GT, lag 


(1994) a 






Sutherland et al. 


Polynomial 


GT, lag 


(1995, 1997) c 






Rasch (2002) 


Polynomial 


Growth rate 


Ross et al. (2003) 


Square-root 


GT 


Skandamins et al. 


Vitalistic 


Time to decline 


(2002) 




in cell 
concentration 


Garcia-Gimeno et 


ANN 


Growth rate, 


al. (2003) 




lag time 


Whiting and Golden 


Polynomial 


Time to decline 


(2003) 




in cell 
concentration 


Listeria 






monocytogenes 6 






Buchanan et al. 


Polynomial 


Time to decline 


(1997) 




in cell 
concentration 


Razavilar and 


Polynomial 


Probability of 


Genigeorgis 




growth 


(1998) 






Cheroutre-Vialette 


ANN 


Absorbance at 


and Lebert (2000) 




600 nm 



Independent Variables and Ranges 

T (4-28°C); NaCl (0-4%); pH (5-7); 

initial spore cone. (1-5 log cfu/g) 
T (15-30°C); NaCl (0.5-4.0%); pH 

(5.5-6.5) 
T (5-12°C); NaCl (0.5-2.5%); pH 

(5.5, 6.5); C0 2 (0-90%) 

T (12-42°C); NaCl (0-3%); pH 
(5.5-7.0.); Na-pyrophosphate 
(0-3%) 

T (5-42°C); NaCl (0.5-5.0%); pH 

(4.5-8.5); Na-nitrite (0-200 ppm); 

aerobic and anaerobic 
T (10-30°C); NaCl (0.5-6.5%); pH 

(4.0-7.0); Na-nitrite (0-200 ppm); 

aerobic 
T (10-30°C); NaCl (0.5-3.0%); pH 

(4.5-6.5); reuterin (0-4 AU/ml) 
T (7.6-47.4°C); a w (0.95 1-0.999); 

pH (4.02-8.28); lactic acid (0-500 

mM) 
T (0-15°C); pH (4.0-5.0); oregano 

essential oil (0.0-2.1%) 

T (9-2 1°C); NaCl (0-8%); pH 
(4.5-8.5); Na-nitrite (0-200 ppm) 

T (4-37°C); NaCl (0-15%); pH 
(3.5-7.0); Na-lactate (0-2%); Na- 
nitrite (0-75 ppm) 



T (4-42°C); NaCl (0.5-19%); pH 
(3.2-7.3); lactic acid (0-2%); Na- 
nitrite (0-200 ppm) 

T (4-30°C); NaCl (0.5-12.5%); 
methyl paraben (0-0.2%); pH (-5.9); 
K-sorbate (0.3%); Na-propionate 
(0.1%); Na-benzoate (0.1%) 

T (20°C); pH (5.6-9.5); NaCl 
(0-8%) 






2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 101 Wednesday, November 12, 2003 12:40 PM 








TABLE 3.5 (Continued) 

Examples of Secondary Models for Growth of Pathogenic and 

Indicator Microorganisms 





Type of 




Microorganisms and 


Secondary 


Response 


References 


Model 


Variables 


Bouttefroy et al. 


Polynomial 


Cell 


(2000) 




concentration 


Rodriguez et al. 


Arrhenius 


Hm ax 


(2000) 






Augustin and 


CPM 


Umax, ^g 


Carlier (2000a,b) 







Ross et al. (in press) Square-root 



Buchanan and 
Phillips (2000) 

Buchanan and 
Phillips (2000) 

Devlieghere et al. 
(2001) 

Le Marc et al. 

(2002) 



Polynomial 



Polynomial 



Square-root, 
polynomial 

CPM 



Seman et al. (2002) Polynomial 



K 



GT, lag 



GT, lag 



Mmax» la § 



Mmax> la § 



Growth rate 



Independent Variables and Ranges 

T (22°C); pH (5.0-8.2); NaCl 
(0-6%); curvaticin 13 (0-160 
AU/ml) 

T (4-20°C) 

T (-2.7 to -45.5^);^ 
(0.910-0.997); pH (4.55-9.61); 
acetic acid (0-20. 1 mM); lactic acid 
(0-5.4 mM); citric acid (0-1.6 
mM); Na-benzoate (0-0.7 mM); K- 
sorbate (0-5.1 mM); Na-nitrite 
(0-11.4 \\M); glycerol monolaurin 
(0-118.5 ppm); butylated 
hydroxyanisole (0-254 ppm); 
butylated hydroxytoluene (0-48.7 
ppm); terf-butylhydroquinone 
(0-1400 ppm); C0 2 (0-1.64 
proportion); caffeine (0-10.8 g/1); 
phenol (0-12.5 ppm) 

7(3-40 °C); a w (0.920-0.997); pH 
(4.0-7.8); lactic acid (0-450 mM); 
nitrite (0-150 ppm) 

T (4-37°C); pH (4.5-7.5); NaCl 
(0.5-10.5%); Na-nitrite (0-1000 
ppm); aerobic 

T (4-37°C); pH (4.5-8.0); NaCl 
(0.5-5.0%); Na-nitrite (0-1000 
ppm); anaerobic 

T(4-12°C); tf w (0.9622-0.9883); Na- 
lactate (0-3.0%); Na-nitrite (20 
ppm); pH (6.2) 

r(0.5-43°C); pH (4.5-9.4); acetic 
acid (16-64 mM); lactic acid 
(40-138 mM); propionic acid 
(18-55 mM) 

T (4°C); NaCl (0.8-3.6%); Na- 
diacetate (0.0-0.2%); K-lactate 
(0.15-5.6%); Na-erythrobate (317 
ppm); Na-nitrite (97 ppm); Na- 
tripolyphosphate (0.276%) 






2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 102 Wednesday, November 12, 2003 12:40 PM 







TABLE 3.5 (Continued) 

Examples of Secondary Models for Growth of Pathogenic and 

Indicator Microorganisms 



Microorganisms and 
References 



Type of 

Secondary 

Model 



Gimenez and Square-root 

Dalgaard (in press) 



Salmonella 

Gibson et al. (1988) c Polynomial 



Response 
Variables 



M n 



GT, lag 



Davey and 


Arrhenius 


Growth rate, 


Daughtry (1995) 




lag 


Koutsoumanis et al. 


Polynomial 


Mm ax 


(1998) 






Oscar (1999) 


Polynomial 


Umax' la § 


Oscar (2002) 


Square-root, 
CPM 


Umax, ^g 


Skandamins et al. 


Vitalistic 


Time to decline 


(2002) 




in cell 
concentration 


Shigella 






Zaika et al. (1994, 


Polynomial 


GT, lag 


1998) a 






Jeyamkondan et al. 


ANN 


GT, lag 


(2001) 






Staphylococcus aureus 






Ross and 


Square-root 


Growth rate 


McMeekin (1991) 






Buchanan et al. 


Polynomial 


GT, lag 


(1993) a 






Dengremont and 


Square-root 


r*max 


Membre (1997) 






Eifert et al. (1997) 


Polynomial 


Parameters in 
primary 
model 


Vibrio spp. 






Miles et al. (1997) 


Square-root 


GT 



Independent Variables and Ranges 

r(4-10°C); %WPS (2-6%); smoke 
components/phenol (3-10 ppm); 
pH (5.9-6.3); lactic acid (0-20,000 
ppm); interaction with lactic acid 
bacteria 

T (10-30°C); NaCl (0.5-4.5%); pH 

(5.6-6.8); aerobic 
Data from Gibson et al. (1988) 

T (22-42°C);pH (5.5-7.0); 

oleuropein (0-0.8%); aerobic 
T (15-40°C); pH (5.2-7.4); previous 

pH (5.7-8.6); aerobic 
T (8-48°C); aerobic 

T (5-20°C); pH (4.3-5.3); oregano 
essential oil (0.5-2.0%) 



T (10-37°C); NaCl (0.5-5.0%); pH 
(5.0-7.5); Na-nitrite (0-1000 ppm); 
aerobic 

Data from Zaika et al. (1994) 



T (5-35°C); a w (0.848-0.997) 

7 7 (12-45°C); NaCl (0.5-16.5%); pH 
(4.5-9.0); Na-nitrite (1-200 ppm); 
aerobic and anaerobic 

T (10-37°C); NaCl (0-10%); pH 
(5-8) 

T (12-28°C); NaCl (0.5-8.5%); pH 
(5.0-7.0); acidulants HC1, acetic 
acid, or lactic acid; aerobic 

T (8-45°C); a w (0.936-0.995); 
aerobic 






2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 103 Wednesday, November 12, 2003 12:40 PM 







TABLE 3.5 (Continued) 

Examples of Secondary Models for Growth of Pathogenic and 

Indicator Microorganisms 



Microorganisms and 
References 


Type of 

Secondary 

Model 


Response 
Variables 


Yersinia spp. 
Bhaduri et al. 
(1995) a 


Polynomial 


GT, lag 


Sutherland and 

Bayliss (1994) c 
Pin et al. (2000) 


Polynomial 
Polynomial 


GT, lag 

Umax' la § 


Wei et al. (2001) 


Square-root 


Umax, ^g 




Independent Variables and Ranges 

T (5-42°C); NaCl (0.5-5.0%); pH 

(4.5-8.5); Na-nitrite (0-200 ppm); 

aerobic 
T (5-30°C); NaCl (0.5-6.5%); pH 

(4.0-7.0); aerobic 
T (1-8°C); C0 2 (0-83%); 2 

(0-60%) 
r(4-34°C); air; vacuum; C0 2 100% 

a Models included in the Pathogen Modeling Program, which is available free of charge at 
www. arserrc . gov/mfs/PMP6_download.htm. 
b Generation time = ln(2)/(i max . 

c Model included in Food MicroModel. The values of model parameter are not included in the manu- 
script. 
d See Ross et al. (2000) for a list of Listeria monocytogenes growth models published prior to 2000. 




include germination of spores, population growth, survival, and toxin formation. 
These types of models became known as "probability" models. 

In the latter part of the 1 990s it seemed that the only way to manage the risk to 
consumers from certain pathogens was to ensure that the organism was never present 
in foods, or to ensure that it was not able to grow in foods that could become 
contaminated. The latter imperative led to the re-development of "growth/no-growth 
boundary," or "interface" modeling. 

This section is divided into three main parts. In the rst, Section 3.4.2, "tradi- 
tional" probability modeling is brie y discussed. Section 3.4.3 presents and discusses 
the newer growth/no-growth (G/NG) modeling approaches, while Section 3.4.4 
considers methodological issues relevant to probability and G/NG modeling. 

3.4.2 Probability Models 

Several reviews of probability modeling were presented in the early 1990s (Baker, 
1993; Baker and Gemgeorgis, 1993; Dodds, 1993; Lund, 1993; Ross and 
McMeekin, 1994; Whiting, 1995) but, possibly because of the relative paucity of 
new publications in this field since then, there has been no more recent dedicated 
review. Whiting and Oriente (1997) and Zhao et al. (2001), however, provide 
succinct updates. 



2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 104 Wednesday, November 12, 2003 12:40 PM 






TABLE 3.6 

Examples of Secondary Models for Growth of Spoilage Microorganisms 




Microorganisms and 


Type of 


Response 


References 


Model 


Variables 


Bacillus 






stearothermophilus 






Ng et al. (2002) 


Polynomial 


Growth rate 
GOL a 


Brochothrix 






thermosphacta 






McClure et al. 


Polynomial 


Mmax, la E 


(1993) 






Abdullah et al. 


Polynomial 


GT, lag, 


(1994) 




MPD 


Geeraerd et al. 


ANN 


Mmax> ^g 


(1998a) 






Pin and Baranyi 


Polynomial 


Umax* la § 


(1998) 






Koutsoumanis et 


Arrhenius 


Hmax 


al. (2000) 






Jeyamkondan et al. 


ANN 


GT, lag 


(2001) 






Chryseomonas spp. 






Membre and 


Square-root 


Mm ax 


Kubaczka(1998) 






Enterobacteriaceae 






Pin and Baranyi 


Polynomial 


Mmax> ^g 


(1998) 






Lactic acid bacteria 






Passosetal. (1993) 


Kinetic 


Mm ax 


Ganzle et al. 


Square-root 


Hmax 


(1998) 






Ganzle et al. 


CPM 


Mm ax 


(1998) 






Devlieghere et al. 


Square-root, 


Mmax> ^g 


(2000a,b) 


polynomial 




Lou and Nakai 


ANN 


Mmax> ^g 


(2001) 






Wijtzes et al. 


Square-root 


Mm ax 


(2001) 







Independent Variables and Ranges 



T (45-60°C); NaCl (0-1.5%); pH (5.5-7.0) 



r(l-30°C); NaCl (0.5-8.0%); pH (5.6-6.8); 

aerobic 
T (-2 to -10°C); C0 2 (2-40%); diameter of 

meat particles (2-10 mm) 
Data from McClure et al. (1993) 

T (2-1 1°C); pH (5.2-6.4); aerobic 

T (0-20°C); C0 2 (0-100%) 

Data from McClure et al. (1993) 



r(1.3-10°C); aerobic 



T (2-1 1°C); pH (5.2-6.4); aerobic 



pH (3.8-6.0); lactic acid (0-30 mM); acetic 
acid (0-40 mM); NaCl (0-9%); cucumber 
juice 

T (3-41°C); aerobic 

pH (4.2-6.7); ionic strength (0.0-1.97); 

acetate (0-0.2 mM); aerobic 
r(4-12°C); a w (0.962-0.9883); C0 2 (0-1986 

ppm); Na-lactate (0.0-3.0 5); pH 6.2 

r(4-12°C); a w (0.9 62-0.9883); CO, (0-241 1 

ppm); pH 6.2 
Subset of data from Devlieghere et al. 

(2000a,b) 
T (3-30°C); a w (0.932-0.990); pH (5.0-7.5) 






2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 105 Wednesday, November 12, 2003 12:40 PM 







TABLE 3.6 (Continued) 

Examples of Secondary Models for Growth of Spoilage Microorganisms 



Microorganisms and 


Type of 


Response 


References 


Model 


Variables 


Connil et al. (2002) 


Polynomial 


Umax, ^g 


Messens et al. 


CPM 


Growth, 


(2002) 




bacteriocin 
production 


Leroy and De 


CPM 


Growth, 


Vuyst (2003) 




bacteriocin 
production 


Garcia-Gimeno et 


ANN 


Growth rate, 


al. (2002) 




lag 


Messens et al. 


CPM 


Growth, 


(2002) 




bacteriocin 
production 


Garcia-Gimeno et 


ANN 


Growth rate, 


al. (2002) 




lag 


Molds 






Gibson et al. 


Polynomial 


Colony 


(1994) 




growth 


Cuppers et al. 


Square-root, 


Colony 


(1997) 


CPM 


growth 


Valik et al. (1999) 


Polynomial 


Colony 
growth 


Batteyetal. (2001) 


Polynomial 


Probability 
of growth 


Panagou et al. 


Polynomial, 


Colony 


(2003) 


Arrhenius, 
CPM 


growth 


Photobacterium 






phosphoreum 






Dalgaard et al. 


Polynomial, 


Mm ax 


(1997) b 


square-root 




P seudomonas 






Membre and 


Polynomial 


Mmax> ^g 


Burlot (1994) 






Neumeyer et al. 


Square-root 


GT 


(1997) c 






Pin and Baranyi 


Polynomial 


Mmax> ^g 


(1998) 







Independent Variables and Ranges 

T (3-9°C); pH (2.5-6.5); glucose (0.2-0.6%) 
7 1 (20-38°C); pH (4.8-7.0) 



T 7 (20-37°C); pH (4.5-6.5) 



T (20, 28°C); NaCl (0-6%); pH (4-7) 



7 1 (20-38°C); pH (4.8-7.0) 



T (20, 28°C); NaCl (0-6%); pH (4-7) 



r(30°C); a w (0.810-0.995) 

T (5-37°C); NaCl (0-7%) 

T (25°C); a w (0.87-0.995); aerobic 

T (25°C); pH (2.8-3.8); titratable acidity 
(0.2-0.6%); sugar content (8-16°Brix); Na- 
benzoate (100-350 ppm); K-sorbate 
(100-350 ppm) 

T (20-40°C); NaCl (2-10%); pH (3.5-5.0) 



T (0-1 5°C); CQ 2 (0-100%) 



T (4-30°C); pH (6-8); NaCl (0-5%) 



T(0-30°C); a w (0.947-0.966) 



T (2-1 1°C); pH (5.2-6.4); aerobic 






2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 106 Wednesday, November 12, 2003 12:40 PM 







TABLE 3.6 (Continued) 

Examples of Secondary Models for Growth of Spoilage Microorganisms 




Microorganisms and 


Type of 


Response 




References 


Model 


Variables 


Independent Variables and Ranges 


Koutsoumanis et 


Arrhenius 


Mm ax 


T (0-20°C); C0 2 (0-100%) 


al. (2000) 








Koutsoumanis 


Square-root 


Mmax> ^g 


r(0-15°C) 


(2001) 








Rasmussen et al. 


Process risk 


GT 


Data from Neumeyer et al. (1997) 


(2002) 


model 






Shewanella spp. 








Dalgaard (1993) b 


Square-root 


Mm ax 


T (0-3 5 °C); aerobic, anaerobic 


Koutsoumanis et 


Arrhenius 


rnnax 


T (0-20°C); C0 2 (0-100%) 


al. (2000) 








Yeasts 








Deak and Beuchat 


Polynomial 


Changes in 


T (10-30°C); a w (0.93-0.99); pH (3.8-4.6 


(1994) 




conductance 


K-sorbate (0-0.06%) 


Passosetal. (1997) 


Kinetic with 


PfUW 


T (30°C); pH (3.2-5.9); lactic acid (0-55 




product 




mM); acetic acid (0-35 mM); NaCl (0-6% 




inhibition 




cucumber juice; aerobic and anaerobic 


Ganzle et al. 


Square-root 


Hm ax 


T (8-36°C); aerobic 


(1998) 








Ganzle et al. 


CPM 


Mm ax 


Ionic strength (0.0-3.2); acetic acid (0-90 


(1998) 






mM); aerobic 




a GOL = germination, outgrowth, and lag time. 

b Model included in the Seafood Spoilage Predictor (SSP) software available free of charge at 

www.dfu.min.dk/micro/ssp/. 

c Model included in the Food Spoilage Predictor (FSP) software. 



3.4.2.1 Logistic Regression 

Dodds (1993) explains that in relation to the hazard presented by Clostridium 
botulinum in foods, the detection of the toxin is often more important than growth 
and that while growth is continuous and fairly easily determined, the presence of 
detectable toxin was seen as an "all-or-none" response. This led workers to seek 
methods to predict the probability of production of detectable toxin levels in response 
to the independent variables. 

In probability models in predictive microbiology the data are usually that the 
response (e.g., growth, detectable toxin production) is observed under the experi- 
mental conditions, or that it is not. Responses such as detectable toxin production 
can be coded as either (response not observed) or 1 (response observed) or, if 
repeated observations have been made, as probability (between and 1). The prob- 
ability is related to potential predictor variables by some mathematical function 
using regression techniques. 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 107 Wednesday, November 12, 2003 12:40 PM 







Logistic regression is a widely used statistical modeling technique — and is the 
technique of choice — when the outcome of interest is dichotomous (i.e., has only 
two possible outcomes). It is widely used in medical research (e.g., Hosmer and 
Lemeshow, 1989). Because regression techniques do not exist for dichotomous data, 
the regression equation is usually related to the log odds, or logit, of the outcome 
of interest. This has the effect of transforming the response variable from a binary 
response to one that extends from -oo to +°o re ecting the possible ranges of the 
predictor variables, and has desirable mathematical features also (Hosmer and Leme- 
show, 1989). The logit function is de ned as: 

logit P = log(iV(l - P)) (3.42) 

where P is the probability of the outcome of interest. 

Logit P is commonly described as some function Y of the explanatory variables, i.e.: 



logit P = 7 (3.43) 



Equation 3.43 can be rearranged to: 



1/(1 + e~ Y ) = P 





or 



e Y /(\ +e Y ) = P 

where Y is the function describing the effects of the independent variables. 

The latter parameterizations appear in some of the earlier probability modeling 
literature. 

Zhao et al. (2001) assessed the performance of linear and logistic regression to 
model percentage data that are "bounded," and may be considered as rescaled 
probability values. It was con rmed that logistic regression provided a much more 
accurate description of percentage data than linear regression, which had the insur- 
mountable problem of predicting values outside the range of the data (i.e., less than 
0% or greater than 100%). 

3.4.2.2 Confounding Factors 

Probability modelers used logistic regression to de ne the probability that detectable 
toxins would be produced within a specified period of time and under specified 
product composition and storage conditions. Models were based on the idea that a 
product was safe/acceptable or that it was not. Nonetheless, the responses measured 
in "probability modeling" were related to a number of factors that were in turn 
related to the growth of the organism under study and, in some cases, also included 
elements of survival. This approach appears to have arisen from the ideas of Riemann 

2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 108 Wednesday, November 12, 2003 12:40 PM 




~V 









(1967) that the success of a preservation method with regard to C. botulinum is 
related to the probability that one spore will germinate and give rise to toxin in the 
nished product. In general, to assess the effect of preservation conditions on 
probability of toxin production, the probability of growth from a single cell is 
estimated as the number of spores able to initiate growth under the test conditions 
(usually determined by MPN [most probable number] methods) divided by the 
number originally inoculated (Lund, 1993). Often a series of increasingly dilute 
inocula are subjected to the test conditions to determine the minimum fraction able 
to initiate detectable growth under the test conditions. 

It might be expected that the probability of detection would increase with time. 
Indeed, Lindroth and Genigeorgis (1986) recognized that the probability of growth 
detection was also dependent upon the lag time of the inoculum, its initial density, 
and the duration of the study. They introduced a modi cation to the logit model to 
specifically model these effects. That model was subsequently used in a number of 
other studies (Baker et al., 1990; Ikawa and Genigeorgis, 1987). Whiting and Call 
(1993) criticized earlier models for probability of C. botulinum outgrowth and toxin 
production because they did not specifically monitor the time at which growth/toxin 
formation was rst detected, and specifically modeled the probability of formation 
of toxin as a function of time and storage conditions using the logistic function, i.e., 
the probability of detectable growth, when plotted as a function of time, is a sigmoid 
curve. That approach was further re ned (Whiting and Oriente, 1997 ; Whiting and 
Strobaugh, 1 998) by inclusion of the inoculum density as an independent variable 
in the model. 

Clearly, the probability of the responses in many of these traditional probability 
models is strongly related to the growth rate of the organism under the experimental 
conditions, leading Ross and McMeekin (1994) to conclude that the distinction that 
had traditionally been made between probability and kinetic models was an arti cial 
one. Similarly, Baker et al. (1990) noted that "The rate of P increase ... expresses 
the growth rate ... ." 

However, under some experimental conditions P does not always reach an 
asymptote of 1 . This is evident in the data of Whiting and colleagues (Whiting and 
Call, 1993; Whiting and Oriente, 1997; Whiting and Strobaugh, 1998), of Chea et 
al. (2000), and of Razavilar and Genigeorgis (1998). It had also been described 
earlier by Lund et al. (1987) who introduced to predictive microbiology a model 
that recognizes that under some conditions, no matter how long one waits, not all 
samples will show growth/toxi genesis. 

While the above studies considered spores, Razavilar and Genigeorgis (1998) 
applied a logistic regression approach to the probability of growth initiation within 
58 days of Listeria monocytogenes and other Listeria species in response to 
combinations of pH, salt, temperature, and methyl paraben, sodium propionate, 
sodium benzoate, and potassium sorbate (Table 3.5). Their results, also, suggested 
that under near-growth-limiting environmental conditions the asymptotic probabil- 
ity of growth (i.e., given in nite incubation time) was sometimes less than 1. 
Stewart et al. (2001) also commented that while kinetic models predict the mean 
growth rate, these estimates may be meaningless under stressed conditions owing 
to natural variability in biological responses. Similarly, Lund (1993) employed the 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 109 Wednesday, November 12, 2003 12:40 PM 




~V 









Gompertz model (see Chapter 2) to model the time-dependent probability of growth 
of L. monocytogenes Scott A as a function of environmental factors. Even at near- 
growth-limiting pH (4.3), however, the asymptote of the log(P growth ) was still close 
to 1. 

The above studies suggest that as environmental conditions become more inhib- 
itory to growth, not only does the probability that growth will be observed during 
the course of the experiment decrease, but the probability that growth is possible also 
decreases. This may be because the generation or lag time of all cells within the 
inoculum becomes in nitely long. Under these conditions, one begins to identify the 
absolute limits to microbial growth under combined stresses, i.e., the G/NG interface. 

3.4.3 Growth/No Growth Interface Models 

Microbial growth is restricted to nite ranges for any environmental factor, with 
growth rate sometimes declining abruptly within a very small increment of change 
of environment. Individual factor limits have been determined and collated (e.g., 
ICMSF, 1996a). That the growth range of microorganisms for one factor is reduced 
when a second environmental factor is less than optimal is also well recognized, and 
underlies the Hurdle concept (Leistner et al., 1985) also known as (multiple) barrier 
technology, or combined processes (Gorris, 2000). While the physiological basis of 
this synergy remains incompletely understood, the ability to de ne the limits to 
growth under combined environmental factors has enormous practical application in 
maintaining the microbial safely and quality of foods. Whether pathogens grow at 
all and the position of the G/NG boundary are of more interest than their growth 
rate because any growth implies a potential to cause harm to consumers. Similarly, 
so-called shelf-stable foods are sold, stored, and consumed over long periods of time. 
Therefore, the ability of spoilage organisms to grow at all implies that they have the 
potential to multiply to suf cient numbers to cause spoilage (Jenkins et al., 2000). 

In the early to mid-1990s, a vein of experimentation using logistic regression 
techniques was begun with the aim of developing models that could de ne absolute 
limits to microbial growth in multifactorial space, irrespective of time of incubation 
or number of cells in the inoculum. One impetus for this research was the problem 
of listeriosis (Parente et al., 1998; Tienungoon et al., 2000). Strategies proposed to 
control the threat of listeriosis included "zero tolerance" (i.e., not detectable in a 
25-g sample) of the presence of L. monocytogenes in foods that could support its 
growth, or to limit levels of contamination at the point of consumption to less than 
100 cfu/g. Thus, foods that did not support the growth of L. monocytogenes were 
considered to pose signi cantly less risk and to require much less regulatory "atten- 
tion" and testing. It was, therefore, of great commercial interest to be able to predict, 
without the need for protracted and expensive challenge testing, the potential for 
growth of specific bacteria within a particular food or, equivalently, product formu- 
lation options that would preclude growth. 

Models de ning combinations of environmental conditions that just prevent 
growth have become known as "G/NG interface," "growth boundary," or more simply 
"growth limits" models. The importance of growth boundary models for the design 
of safe foods and setting of food safety regulations, for the design of shelf-stable 

2004 by Robin C. McKellar and Xuewen Lu 












~V 






1237_C03.fm Page 110 Wednesday, November 12, 2003 12:40 PM 






foods, and as a means of empowering the Hurdle concept by allowing it to be applied 
quantitatively has been discussed by various authors (Masana and Baranyi, 2000b; 
McMeekm et al., 2000; Ratkowsky and Ross, 1995; Schaffner and Labuza, 1997). 
Various approaches have been suggested to de ne the G/NG boundary. For 
convenience, these are discussed below under three broad groupings: 

1. Empirical, deterministic, approaches 

2. Logistic regression techniques 

3. Arti cial neural networks 

Table 3.7 provides an overview of G/NG models published since 1990. 

3.4.3.1 Deterministic Approaches 

The rst explicit de nition of a microbial G/NG interface appears to be Pitt (1992), 
who derived regression equations from published data to describe the tempera- 
ture/water activity interface for a atoxin production and Aspergillus spp. growth. 
The equation used to describe the interaction between temperature and water activity 
limits for growth was: 



r (minmax) = 19 ^ ± /(g56.71 - 2289 X (1. 172 - tf )) 



g v v ^ w 

where J C 1 ™™ 8 *) are the upper and lower temperature limits for growth at the specified 
water activity. 

A similar equation was presented for a atoxin production. The predicted inter- 
faces from both models are shown in Figure 3.7 . 

To describe the pH/a w(NaC1) interface of the food spoilage organism Brocothrix 
thermosphacta, Masana and Baranyi (2000b) derived the midpoints of growth and 
no-growth observations by interpolation and fitted a polynomial function to those 
data. They noted that under some conditions, the interface was completely dominated 
by one factor or the other, so that their nal model consisted of a pH vs. a w parabolic 
curve and a NaCl-constant line. They also considered the effects of inoculum level 
on the interface, which was determined at 25°C for up to 24 days of incubation. 
Examples of the interface are shown in Figure 3.8. 

Membre et al. (2001) estimated levels of sorbate that prevented growth of 
Penicillium brevicomp actum in bakery products containing various levels of benzoate 
by extrapolation of kinetic data. Equations were derived to de ne growth-preventing 
combinations of sorbate and benzoate and were used to limit the range of predictions 
from the kinetic model they developed for P. brevicomp actum growth rate. 

Other workers have noted that the form and parameters of CPMs imply absolute 
limits to microbial growth, and suggested approaches to de ning the G/NG interface 
based on estimates of cardinal parameters. In this vein Ratkowsky and Ross (1995), 
recognizing the relationship between absolute limits for each environmental factor 
and their relationship to the parameters of square-root-type models and CPMs, 
experimented with the use of a kinetic model as the basis of a growth boundary 



2004 by Robin C. McKellar and Xuewen Lu 











TABLE 3.7 

Summary of Published Growth Boundary Models 



Experimental Design 














Environmental 


■ VII 


.. & «-o 




Reference 


Organism 


Strain 


Medium 


Factors 


Lower 


Upper 


Lev 


Presser et al. 


Escherichia 


M23 (non- 


Nutrient 




Temperature 


10 


37 


6 


(1998) 


coli 


pathogenic) 


Broth 




a w (NaCl) 
pH 

Lactic acid 
(mM) 


0.955 

2.8 




0.995 

6.9 
500 


4 

>10 

6 


Parente et al. 


Listeria 


Scott A, V7, 


Tryptone 


Soy 


Nisin (IU/ml) 


1 


2100 




(1998) 


mono- 
cytogenes 


and LI 1 


Broth + 0.6% 
Yeast Extract 


Leucocidin F10 
(AU/ml) 


1 


2100 














pH 


4.7 


6.5 














NaCl (% w/v) 


0.7 


4.5 














EDTA (mmol) 


0.1 


0.9 














Inoculum density 


1.6 x 10 3 


7.9 x 10 7 








Validation 






Nisin (IU/ml) 


8 


200 








Set 1 






Leucocidin F10 

(AU/ml) 
pH 

NaCl (% w/v) 
EDTA (mmol) 
Inoculum density 


8 

4.7 
0.7 
0.08 
0.6 x 10 3 


200 

6.5 

4.5 

4.72 

2.5 x 10 7 








Validation Set 






Nism (IU/ml) 


50 


250 








2 






Leucocidin F10 

(AU/ml) 
pH 

NaCl (% w/v) 
EDTA (mmol) 
Inoculum density 


1 

5.2 
1.8 
0.2 
1 x 10 5 


250 

6 

0.6 





1 to 4 



Total Data 
Points 

627 



Measured by? Time Limit 



OD Increase 
(con rmedas 
needed by 
culture) 



50 days 



Other 

Linear logistic 
regression, SAS 
PROCNONLIN 



7 days 
(@30°C) 



Logistic regression 
with polynomial 
using LOGIT 
1.14 module of 
Systat 



7 days 
(@30°C) 



10 



7 days 
(@30°C) 






I 
O 

o 

3' 

OG 
O 



< 
O 

a. 

i=s 

o 

IT. 

c 

< 

I 

a 



to 

o 
o 



to 

4- 

o 

-J 




2004 by Robin C. McKellar and Xuewen Lu 










TABLE 3.7 (Continued) 

Summary of Published Growth Boundary Models 



Reference Organism 



Bolton and 
Frank 
(1999) 



Listeria 
mono- 
cytogenes 




Strain 

Mixture 
(equal 

numbers) of 
Scott A, Brie 
1,71 

Switzerland, 
2379 LA 



Medium 

Soft fresh 
cheese 
(similar to 
"Mexican 
style" 
cheese) 



Experimental Design 



Environmental 
Factors 

Moisture 

(% w/w) 
salt (% w/w) 
pH 



Ranges 



Lower Upper Levels Replicates 



42 

2 
5 



60 

8 
6.5 



Total Data 
Points 

288 



Measured by? Time Limit 



Viable count 



21 and 42 
days 



4 
6 



Other 

Binary or "ordinal" 
logistic 

regression using 
SAS PROC 
LOGISTIC with 
link functions. 
For the latter, 
three responses : P 
of growth, stasis, 
or death 
(according to 
change in viable 
count; ±0.5 log 
CFU) were 
modeled 






I 
O 

o 

3' 

CTQ 
O 



to 

< 

o 

a. 

i=s 

o 

IT. 

c 

< 

I 

c 



to 

o 
o 



4- 

o 

-J 



Salter et al. 


Escherichia 


MR21 


Nutrient 


Temperature 


7.7 


37 


60 


(2000) 


coli 


(STEC) 


Broth 


a w (NaCl) 


0.943 


0.987 


28 



1-8, most 4 604 



OD increase 50 days 



Nonlinear logistic 
regression, SAS 
PROCLOGISTIC 
and 
PROCNONLIN 



Jenkins et al. 


Zygosaccha.ro 


4637, history 


Acidi edyeast 


Salt (NaCl, 


2.6 


4.2 


3 


(2000) 


myces bailii 


unknown 


nitrogen 


%w/v) 














broth 


Sugar (fructose, 
%w/v) 


7 


32 


3 










Total acetic acid 


1.8 


2.8 


3 










(%v/v) 
pH 


3.5 


4 


3 



243 



OD increase 29 days 



SAS LIFEREG 




2004 by Robin C. McKellar and Xuewen Lu 










Tienungoon 
et al. 
(2000) 



Listeria 
mono- 
cytogenes 



(Scott A, L5 
separate 
models) 



TSB-YE 



Temperature 3.1 

a w (NaCl) 0.928 

pH 3.7 

Lactic acid 

(mM) 



36.2 
0.995 

7.8 
500 



30 
60 
10 
14 



1 to 4 



Lopez-Malo 


Saccharo- 


Not stated 






^ 


0.93 


0.97 


3 


et al. 


myces 


Model based 






pH 


3 


6 


4 


(2000) 


cerevisiae 


on data of 
Cerruti et al. 
(1990) 






K-sorbate (ppm) 





1000 


6 


Masana and 


Brocothrix 


MR 165 


Tryptone 


Soya 


NaCl 


0.5 


10 


11 


Bar any l 


thermo- 




Broth 




pH 


4.4 


5.7 


7 


(2000b) 


sphacta 








Inoculum (cfu/ 
350 |ll) 


10 


1 million 


3 



2839 


OD increase 


90 days 


Nonlinear logistic 
regression, SAS 
PROCLOGISTIC 
and 
PROCNONLIN 


72 (x 2 


Viable count, 


50 h or 


Logistic regression 


observation 


including 


350 h 


with rst-order 


times) 


decrease in 




polynomial using 




viable count 




SPSS 




Viable count 




Polynomial fitted 
to midpoints of 
data-pairs of 
adjacent growth 
and no growth 
observed 
combinations 








Fine grid 
experiments 




NaCl 
pH 












Inoculum (cfu/ 
350 fll) 


Lanciotti et 
al. (2001) 


Bacillus 
cereus 

Staphylococ- 
cus aureus 


FG1 

S33 




BHI Broth 
BHI Broth 


Temperature 
&w (glycerol) 




Salmonella 
enteritidis 


B5 




BHI Broth 


pH 



45 combinations at 5 levels of 
NaCl and 5 levels of pH close 
to the G/NG boundary 

10 1000 2 



10 



450 



Viable count 



10 



0.89 



Ethanol (% v/v) 



45 
0.99 

8 

3 



5 


30 variables 


2 x 30 for 


OD increase 




combina- 


each strain 


(600nm) 


8 


tions over 
two 






5 


independent 






trials for 








each 






5 


organism 







2-7 days 



Generalized linear 
logistic 
regression, 
Statistica 
(Statsoft) 
software 






to 

I 
O 

o 

3' 

CTQ 
O 



< 

o 

a. 

i=s 

o 

IT. 

c 

< 

I 

a 



to 

o 
o 



to 

4- 

o 




2004 by Robin C. McKellar and Xuewen Lu 













I 
O 

o 

3' 

era 
o 




TABLE 3.7 (Continued) 

Summary of Published Growth Boundary Models 



Experimental Design 



Reference Organism 



Strain 



Stewart etal. Staphylococ- 5 strain 
(2001) cus aureus cocktail 



Medium 



BHI Broth 



Environmental 
Factors 



Ranges 



Lower Upper Levels Replicates 



a w (glycerol) 0.95 

Initial pH 4.5 

K-sorbate (ppm) 



or 



Total Data 

Points Measured by? Time Limit 



0.84 
7 



4 
4 











Ca-propionate 





1000 


3 










(ppm) 
















Temperature 


37 














(°C) 








McKellar 


Escherichia 


5 strain 


TS Broth 


Temperature 


10 


30 


5 


and Lu 


coli 015 :H7 


cocktail 




(°C) 








(2001) 








Acetic acid 
(modeled as 
undissociated 
form) 





4% 


8 










NaCl 


0.50% 


16.50% 


8 










Sucrose 





8% 


3 










pH 


3.5 


6.0 


6 



640 



OD increase 168 days 



1820 



Visible 
increase in 
turbidity 



3 days 



Other 

Toxin assayed 
Modeled "time to 

growth" using 

LIFEREG 



Linear logistic 
regression used 
(polynomial 
form) 



< 

o 
a. 
i=s 
o 

IT. 

c 

< 

I 

a 



to 

o 
o 
OJ 



4- 

o 

-a 





2004 by Robin C. McKellar and Xuewen Lu 














Membre et 


Penicillium 


Wild type 


MY50 agar 


pH 


2.5 


7.5 


±6 


1 


76 


Mycelial 


75 days 


Not directly 


al. (2001) 


compactum 


from bakery 
products 




Sorbic acid 

Propionic acid 

Sodium benzoate 

pH 

Sorbic acid 
(mg H) 

Na-benzoate 
(mg H) 


0.0 
0.0 


1000.0 
300.0 


1 
1 
1 
1 
6 

4 


5 


122 


growth 




modeled, growth 
limits due to 
sodium benzoate 
and sorbic acid at 
pH 5 were 
derived by 
extrapolation of 
growth rate data 








Commercial cakes 








6 


4 








Uljas et al. 


Escherichia 


3 strain 


Apple cider 


pH 


3.1 


4.3 


7 




1600 x 3 


Turbidity 


12 hours 


SAS 


(2001) 


coli 015 :H7 


cocktail 


(juice) 


Temperature 
Sodium benzoate 


5 


35 


4 






(growth 




PROCLOGISTIC 


In this case 


the response modeled was P > 5 log inactivation 





0.1% 


3 






within 48 hat 
35°C) after 




(dependency 
modeled as simple 


after various treatment times 


















/ 




c 






















dilution of 




rst order 










Potassium 





0.1% 


3 






treated 




equations of 










sorbate 












sample 




predictor 










Freeze-thaw 


Not applied 


Applied 


3 


756 








variables, no cross 










Ciders type 






3 


2 or 3 








products) 




Staphylococ- 


5 strain 


BHI Broth 


a w (NaCl, or 


0.84 


0.95 


4 


8 




OD increase 


168 days 


Toxin assayed 




cus aureus 


cocktail 




sucrose- 
fructose) 

pH 

K-sorbate (ppm) 


4.5 



7 
1000 


2 

4 

2 










Modeled "time to 
growth" using 
LIFEREG 






[Combined with data set of Stewart et al. (2001), 768 data] 








1792 










Listeria 


ATCC 33090 


BHI Broth 


Temperature 





43 


16 


(pH constant) 




Turbidimetry 


14 days 


Novel term based 




innocua 




(+0.2% w/w 
glucose, 

+0 3% w/w 


pH 


4.5 


9.4 


15 


(temperature 
constant) 




Viable count 
by culture 


1 month 


on relative 
inhibition of 
growth rate- 








yeast extract 


Propionic acid 


16 


64 


24 


(pH varied 








affecting factors 








(mM) 








from 5 to 








data generated for 


















7.5) 








combined kinetic 










Lactic acid (mM) 


20 


138 


27 


(pH varied 
from 4.8 to 
7.1) 








model that 
predicts no growth 
(NLINFIT in 
























MATLAB 5.2) 



I 
O 

o 

3' 

CTQ 
O 



< 
O 

a. 

i=s 

o 

IT. 

c 

< 

I 

o 



to 

o 
o 



to 

4- 

o 




2004 by Robin C. McKellar and Xuewen Lu 











TABLE 3.7 (Continued) 

Summary of Published Growth Boundary Models 



Reference Organism 



Battey and 
Schaffner 
(2001) 



Spoilage 
bacteria: 
Acineto- 
bacter 
calco- 
aceticus and 
Gluconobac- 
ter oxydans 



a 


JV 






v5 


V 








Battey et al. 


Spoilage 






(2002) 


yeasts: 






Saccharo- 






myces 






cerevisiae, 






Zygosaccha 






romyces 






bailii, 






Candida 






lipolytica 




Hajmeer and 


Data of Salter 




Basheer 


etal.(2000), 




(2002, 


see above 




2003a, 








2003b) 





Strain 

2 strain 
cocktail 



Medium 

Cold lied, 
ready to 
drink, 
beverages 



3 strain 
cocktail 



Cold lied, 
ready to 
drink 
beverages 



Experimental Design 



Environmental 
Factors 

pH 



Titratable acidity 3 

(%) 
Sodium benzoate 3 

(ppm) 

Sugar content 3 

(°C Brix) 

Potassium 3 

sorbate (pp) 

pH 3 

Titratable acidity 3 

(%) 
Sodium benzoate 3 

(ppm) 

Sugar content 3 

(°C Brix) 

Potassium 3 

sorbate (pp) 



Ranges 



Lower Upper Levels Replicates 



Total Data 
Points 



Measured by? Time Limit 



2.8 


3.8 


0.2 


0.6 


100 


350 


8 


16 


100 


350 


2.8 


3.8 


0.2 


0.6 


100 


350 


8 


16 


100 


350 



84 



Viable plate 
count 



8 weeks at 

25°C 



84 



Viable plate 

count 



8 weeks at 

25°C 



Other 

Model is based on 
growth and 
inactivation rates. 
Included 14 
duplicated 
validation trials 
(8 correctly 
predicted from 
model) 



Included 14 
duplicated 
validation trials 
(all correctly 
predicted from 
model) 



Probabilistic 
Neural Network 






I 
O 

o 

3' 

(a 

OQ 
O 



< 

o 

a. 

i=s 

o 

IT. 

c 

< 

I 

a 



to 

o 
o 



to 

4- 

o 

-a 





2004 by Robin C. McKellar and Xuewen Lu 










~V 



1237_C03.fm Page 117 Wednesday, November 12, 2003 12:40 PM 






model using linear logistic regression. This approach is discussed further below in 
Section 3.4.3.2. 

The approaches of Augustin and Carlier (2000b) and Le Marc et al. (2002) were 
presented in Section 3.2.3. Essentially, these approaches are empirical. They are 
based on assumed interactions between factors and are not fitted to G/NG data. An 
example of the response predicted by these approaches is shown in Figure 3.11. 




0.95- 



> 0.91 

'■*-• 
o 

(0 
i- 
(0 

% 0.8fr| 



0.8- 



0.75 




i i i i i i i i i 
10 15 20 25 30 35 40 45 50 55 



Temperature (°C) 

FIGURE 3.7 Predicted temperature-water activity interface for mold {Aspergillus spp.) 
growth (dashed line) and a atoxin production (solid line) from the model of Pitt (1992). 




5-80 



5-40 



? 500 



4-60 



4-20 








005 



0-10 



0-15 



0-20 



0-25 



0-30 



'w 



FIGURE 3.8 Data and modeled growth/no-growth boundary for Brochothrix thermosphacta 

i 

in response to pH and water activity at 25°C. Water activity data were rescaled to b w 



V 



1-a 



The data are for an inoculum of 1.5 x 10 6 cells/well (□), or for an inoculum of 1.5 x 10 1 and 
1.5 x 10 3 cells/well (A). (Reproduced from Masana, MO. and Baranyi, J. Food Microbiol., 
17, 485^193, 2000b. With permission.) 





2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 118 Wednesday, November 12, 2003 12:40 PM 







0.99 



0.98 

^^ 

> 

o 

<c 0.97 

I 

0.96 



0.95 



0.94 



: 






oo o 
oo 8 

COD 

8o 

8 



• o • 
oo 
oo 

O 00 

oo o 



: 
: 



: 
: 



«•♦♦ ♦ tr*t*t* 

i 

«»♦♦ ♦ ♦ ♦ 



♦ «►♦♦♦ 



m • • • •♦♦♦ ♦ ♦ ♦ f 



o° 



s 



s °% $ °s s «& h %° % x* t t J \ :* \ % 8 ♦ 4 



8 

o 



8 

o 



8 

o 



8 

o 



8 

o 



8 

o 



o 

8 

o 







10 



15 20 25 

Temperature (°C) 



30 



35 



40 



45 




FIGURE 3.9 Data obtained from separate experiments for the growth/no- growth (G/NG) 
boundary of Escherichia coli. Data are from Salter et al. (2000) (circles) and from unpub- 
lished results of the authors (diamonds). Near the G/NG boundary, the data obtained from 
discrete experiments do not form a smooth (monotonic) boundary, suggesting that small 
differences in experimental procedures can signi cantly affect the position of the boundary. 
Open symbols denote no-growth conditions, and solid symbols indicate that growth was 
observed. 




The above approaches can be considered to be deterministic; i.e., they predict 
only one position (-P grow th = 0.5) for the boundary, although the position of boundaries 
can be adjusted by "weighting" data in the case of Masana and Baranyi (2000b) or 
by selecting an appropriate value for in the case of the Le Marc et al. (2002) 
approach (see Section 3.2.3). While the data of Masana and Baranyi (2000b) included 
tenfold replication, the midpoints of the most "extreme" conditions that did allow 
growth, and the least "extreme" conditions that did not allow growth were estimated 
by interpolation and considered to represent 50% probability of growth. Other 
workers have suggested that some problems require higher levels of con dence that 
growth will not occur, so that methods that enable de nition of the interface at 
selected levels of statistical con dence may have greater utility. 

Another approach that implicitly characterizes the G/NG interface is that of 
combined growth and death models in which the rate of growth and rate of death 
under specified conditions are estimated simultaneously. The G/NG interface can be 
inferred from those combinations of conditions where growth rate and death rate 
are equal (see, e.g., Jones and Walker, 1993; Jones et al., 1994). A similar approach 
is evident in Battey et al. (2001) who modeled both the rates of growth and rates 
of death of spoilage molds in ready to drink beverages. The G/NG interface was 
given, implicitly, as that set of conditions where the rate of growth was equal to the 
rate of death. 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 119 Wednesday, November 12, 2003 12:40 PM 




~V 









Ratkowsky et al. (1991) noted that as environmental conditions become more 
inhibitory to microbial growth the variability in growth rates increases widely, which 
implies that the probability that growth occurs at all becomes uncertain, because the 
left-hand tail of the growth rate distribution falls below zero. This is supported in 
the results of Whiting and colleagues (Whiting and Call, 1993; Whiting and Oriente, 
1997; Whiting and Strobaugh, 1998), where P max (the proportion of spores that 
successfully germinated and initiated growth) was shown to decline at increasingly 
stringent conditions. Conversely, Masana and Baranyi (2000b) observed, as have 
other workers (McKellar and Lu, 2001; Presser et al, 1998; Salter et al., 2000; 
Tienungoon et al., 2000), that the difference in conditions that allow growth and 
those that do not is usually abrupt, and often at or beyond the limits of resolution 
of instruments commonly used to measure such differences. Thus, Masana and 
Baranyi (2000b) questioned the need for approaches that model the transition 
between conditions leading to high probability of growth and those leading to low 
probabilities of growth. While this abrupt transition appears consistent within rep- 
licated experiments it is less certain, however, that the same consistency is true 
between experiments. Figure 3.9, showing experimental data, suggests that responses 
near the boundary may be inconsistent when data from several discrete experiments 
are combined. This may suggest subtle, but important, differences in response related 
to the physiology of the inoculum, or its concentration. Furthermore, it suggests that 
the ability to characterize probabilities of growth under specified sets of conditions 
may be an important element of growth boundary models and that the boundary 
may not be "absolute" but depend on the physiological state of the cell and, by 
inference, on the size of the inoculum. This will be discussed further in Section 3.4.4. 

3.4.3.2 Logistic Regression 

Ratkowsky and Ross (1995) and others (Bolton and Frank, 1999; Jenkins et al., 
2000; Lanciotti et al, 2001; Lopez-Malo et al., 2000; McKellar and Lu, 2001; 
Parente et al., 1998; Stewart et al., 2001, 2002; Uljas et al., 2001) reintroduced the 
use of logistic regression to model categorical data (i.e., growth or no growth) in 
predictive microbiology, enabling probabilistic determination of the G/NG bound- 
ary. Use of the logit function enabled the probability of growth under specific sets 
of conditions to be estimated, so that the G/NG boundary could be specified at 
selected levels of con dence. 

Ratkowsky and Ross (1995) aimed to model absolute limits to growth in mul- 
tifactorial space, but only had available data based on a 72-h observation period. 
While most other workers have preferred polynomial functions to describe the effect 
of independent variables on the logit function, in the former approach a square -root- 
type kinetic model was In-transformed and used as the basis of the function relating 
the logit of probability of growth to independent variables, e.g., temperature, water 
activity, pH. This approach was adopted in an attempt to retain some level of 
biological interpretability of the models, a desire echoed by others (Augustin and 
Carher, 2000a,b; Le Marc et al., 2002). 

The form of the G/NG interface model of Presser et al. (1998) was derived from 
the kinetic model of Presser et al. (1997) for the growth rate of E. coli (see Equation 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 120 Wednesday, November 12, 2003 12:40 PM 








3.10). Novel data were generated specifically to assess the limits of E. coli growth 
under combinations of temperature, pH, a w and lactic acid. The corresponding G/NG 
model had the form: 

LogitP = 28.0 + 8.90 ln(a„ - 0.943) 

+ 2.01n(r-4.00) + 4.59 ln(l - 10 39 -p h ) (3.44) 

+ 6.961n[l -LACI{\0.1 x (1 + 10p h " 3 - 86 ))] 
+ 3.061n[l -LAC/(S23 x (1 + 10 386 -p h ))] 

where all terms are as de ned in Section 3.2.1. 

Some parameters in that model had to be determined independently, i.e., were 
not determined in the regression, and were derived from the fitted values of square- 
root-type kinetic models. Essentially the same approach was adopted by Lanciotti 
et al. (2001) to develop G/NG models for B. cereus, S. aureus, and Salmonella 
enteritidis. Ratkowsky (2002) commented on the increased exibility in being able 
to determine all of the parameters in the model during the regression, and subsequent 
studies developed the approach, eventually leading to a novel nonlinear logistic 
regression technique (Salter et al., 2000; Tienungoon et al., 2000). Ratkowsky (2002) 
pointed out that nonlinear logistic regression was a new statistical technique and 
discussed bene ts and problems with that approach specifically in relation to growth 
limits modeling. A problem with models of the form of Equation 3.44 is that for 
conditions more extreme than the parameters corresponding to T min , pH min , a w min , 
etc., and which are tested experimentally though not expected to permit growth, the 
terms containing those parameters would become negative. As all of those terms are 
associated with a logarithmic transformation, the expression cannot be calculated 
during the regression and such values are ignored in the model tting process, or 
have to be eliminated from the data set before the tting process begins. This, in 
turn, affects the values of the parameters of the fitted model. Ratkowsky (2002) 
comments that an objective method for selection and deletion of such data is nec- 
essary', but does not yet exist. 

Bolton and Frank (1999) extended the binary logistic regression approach by 
recoding growth and no growth data to allow a third category: survival, or stasis. 
They termed this approach ordinal logistic regression. Parente et al. (1998) 
"reversed" the response variable, and applied logistic regression techniques to the 
probability of survival/no survival of L. monocytogenes in response to bacteriocins, 
pH, EDTA, and NaCl. Stewart et al. (2001) modeled the probability of growth of 
S. aureus within 6 months of incubation at 37°C, and at reduced water activity 
achieved by various humectants. They also compared the growth boundary with the 
boundary for enterotoxin production, and observed a close correlation between the 
two criteria. 

Growth limits models have also been developed for spoilage organisms including 
Saccharomyces cerevisiae (Lopez -Malo et al., 2000) and Zygosaccharomyces bailii 
(Jenkins et al., 2000) and cocktails of Saccharomyces cerevisiae, Zygosaccharomyces 
bailii, and Candida lipolytica (Battey et al., 2002). Interestingly, the study of Jenkins 
et al. (2000), while encompassing broader ranges of factor combinations, con rmed 



2004 by Robin C. McKellar and Xuewen Lu 














~V 



1237_C03.fm Page 121 Wednesday, November 12, 2003 12:40 PM 






the simpler and earlier model of Tuynenburg Muys (1971). That model, which 
specifies combinations of molar salt plus sugar and percent undissociated acetic acid 
for stability of acidic sauces, still forms the basis of the industry standard for those 
products. This observation suggests that limits to growth under combined conditions 
can be highly reproducible. 




3.4.3.3 Relationship to the Minimum Convex 
Polyhedron Approach 

The concept of the MCP was introduced by Baranyi et al. (1996) (see Figure 3.10) 
to describe the multif actor "space" that just encloses the interpolation region of a 
predictive kinetic model. If the interpolation region exactly matched the growth 
region of the organism then the MCP would also describe the growth limits of the 
organism. In practice, however, it would be impossible to undertake suf cient 
measurements to completely de ne the MCP; i.e., the MCP has "sharp" edges 
because of the method of its calculation, whereas from available studies (see Figures 
3.8 and 3.9) the G/NG interface forms a continuously curved surface. However, it 
might also be possible to use no-growth data to create a no-growth MCP and to 
combine the growth MCP and no -growth MCP to de ne a region within which the 
G/NG boundary must lie. This approach has been assessed and compared to a model 
of the form of Equation 3.43 by Le Marc and colleagues (Le Marc et al., 2003). 
These workers concluded that the logistic regression modeling approach produced 
a smoother response surface, more consistent with observations, but that the MCP 
approach had the advantage of being directly linked to observations and therefore 
was not a prediction from a model. 





FIGURE 3.1 Interpolation region (MCP) for a model that includes four-factor combinations 
(7, pH, NaCl, NaN0 2 ). The interpolation region shown is that for NaCl = 0.5%, but is based 
on the complete data set. Solid circles indicate conditions under which observations have 
been made, while the lines represent the edges of the MCP. (From Masana, M.O. and Baranyi, 
J. Food Microbiol. , 17, 367-374, 2000a. With permission.) 





2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 122 Wednesday, November 12, 2003 12:40 PM 








3.4.3.4 Artificial Neural Networks 

Recently, Hajmeer and Basheer (2002, 2003a,b) demonstrated the use of a Proba- 
bilistic Neural Network (PNN) approach to de nition of the G/NG interface. PNNs 
are a form of ANN (see Section 3.2.6). In a series of papers, based on modeling the 
data of Salter et al. (2000) for the effects of temperature and water activity (due to 
NaCl) on the growth limits of E. coli, Hajmeer and Basheer concluded that their 
PNN models provided a better description of the data of Salter et al. (2000) than 
did the nonlinear logistic regression method referred to above. Their conclusion is 
considered in more detail in Section 3.4.3.5 below. 

It should be noted that neither the logistic regression models described above, 
nor the PNN, produce an equation that describes the interface. Rather, the output of 
those models is the probability that a given set of conditions will allow growth. To 
de ne the interface, it is necessary to rearrange the model for some selected value 
of P to generate an equation that describes the G/NG boundary. 

3.4.3.5 Evaluation of Goodness of Fit and Comparison 
of Models 

Methods for evaluation of performance of logistic regression models include the 
receiver operating curve (ROC; also referred to as the concordance rate), the Hos- 
mer-Lemeshow goodness-of- t statistic, and the maximum rescaled R 2 statistic. 
These are considered in greater detail in Tienungoon et al. (2000). 

Brie y, the ROC is obtained from the proportion of events that were correctly 
predicted compared to the proportion of nonevents that were correctly predicted. 
The closer the value is to 1 , the better the level of discrimination. In epidemiological 
studies, ROC values > 0.8 are considered excellent. ROC values for G/NG models 
are typically much higher. 

The Hosmer-Lemeshow index involves grouping objects into a contingency 
table and calculating a Pearson chi-square statistic. Small values of the index indicate 
a good t of the model. 

The maximum rescaled R 2 value is proposed for use with binomial error as an 
analogy to the R 2 value used with normally distributed error. The closer the value 
is to 1 , the greater is the success of the model in predicting the observed outcome 
from the independent variables. Zhao et al. (2001) cite the deviance test and graphical 
tools such as the index plot and half normal plot as methods for determining goodness 
of t of linear logistic regression models. 

Other methods based on calculation of indices from the "confusion matrix" 
(Hajmeer and Basheer, 2002, 2003b) or the equivalent "contingency matrix" 
(Hajmeer and Basheer, 2003a) were used to compare performance between models 
derived from different approaches and applied to the same data. 

Another method of evaluation is to compare the fitted model to independent data 
sets (Bolton and Frank, 1999; Masana and Baranyi, 2000b; Tienungoon et al., 2000) 
although, generally, such data are not readily available (see, e.g., McKellar and Lu, 
2001). The model of Tienungoon et al. (2000) for L. monocytogenes growth bound- 
aries showed very good agreement with the data of McClure et al. (1989) and George 

2004 by Robin C. McKellar and Xuewen Lu 














~V 






1237_C03.fm Page 123 Wednesday, November 12, 2003 12:40 PM 






et al. (1988) despite that different strains were involved. There is also a remarkable 
level of similarity between the observations of Tienungoon et al. (2000) and the 
observations of Le Marc et al. (2002) on growth limits of L. innocna. Several 
publications, however, report growth of L. monocytogenes at temperatures lower 
than the minimum growth limit predicted by the Tienungoon et al. (2000) model, 
possibly indicating strain variation or that the experimental design failed to recognize 
important elements that facilitate L. monocytogenes growth at temperatures < 3°C, 
i.e., that an inappropriate growth substrate was used. Similarly, McKellar and Lu 
(2001) reported that their model failed to predict growth of E. coli 0157:H7 under 
conditions where it had been previously reported, although it should be noted that 
their model was limited to observation of growth within 72 h. Bolton and Frank 
(1999) compared the predictions of their growth limits models for L. monocytogenes 
in cheese to the data of Genigeorgis et al. (1991) for L. monocytogenes growth in 
market cheese. The models predicted correctly in 65% of trials (42-day model) and 
81% of trials (21 -day model). 

Given the diversity of approaches, it is pertinent to ask: does one method for 
de ning the G/NG interface perform better than another? As with kinetic models, 
the ability to describe a specific experimental data set does not necessarily reflect 
the ability to predict accurately the probability of growth under novel sets of con- 
ditions. While measures of performance of logistic regression models are available, 
they can be readily affected by the data set used. Perfect agreement between observed 
and modeled data responses may not be possible if there are anomalies in the data. 
Figure 3.11 provides a clear example. Nonetheless, for many growth limits models 
high rates of concordance (typically >90%) have been reported. As noted earlier, in 
epidemiological logistic regression modeling, rates higher than 70% are considered 
to represent good ts to the data, implying that the limits to microbial growth are 
highly reproducible when well -controlled experiments are conducted. 

To date, only one direct comparison of G/NG modeling approaches has been 
presented (Hajmeer and Basheer, 2002, 2003a,b) but this was based on one data set 
only, i.e., that of Salter et al. (2000) for the growth limits of E. coli in tempera- 
ture/water activity space. Only by comparing the performance of different modeling 
approaches applied to multiple data sets does an appreciation of overall model 
performance emerge. Nonetheless, to illustrate differences between models and give 
some appreciation of their overall performance we compare several models using 
the data of Salter et al. (2000) for the growth limits of E. coli R3 1 in response to 
temperature and water activity. The model types compared are: 

1. The PNN of Hajmeer and Basheer (2003a), which those authors were 
able to summarize as a relatively simple equation 

2. A model of the type of Equation 3.44 fitted to a subset of the Salter et 
al. (2000) data set by Hajmeer and Basheer (2003a) (It should be noted 
that, contrary to what is stated in that publication, the model presented 
by Hajmeer and Basheer was not generated by nonlinear logistic regres- 
sion but by a two-step linear logistic regression as described in Presser et 
al., 1998) 




2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 124 Wednesday, November 12, 2003 12:40 PM 







0.99 



0.98 

>. 

■«-» 

> 

o 

ro 0.97-1 

o 

0.96-1 



0.95 



0.94 



*♦> ♦ ♦ ♦ ♦ ooo d 




%% $ t s°s 



10 



20 30 

Temperature (°C) 



40 



50 



60 




FIGURE 3.1 1 Comparison of predicted no growth boundaries for four modeling approaches 
applied to the data of Salter et al. (2000) (circles) for the growth limits of Escherichia coli 
R31 in response to temperature and water activity (NaCl) combinations. Approach of Hajmeer 
and Basheer PNN (heavy solid line); Linear Logistic /Equation 3.44 (heavy dashed line); Le 
Marc et al. (2002) (light solid line); Augustin and Carlier (2000a) (light dashed line). The 
data set was subsequently augmented with new data (diamonds), which reveals that none of 
the models extrapolate reliably. (Solid symbols: growth; open symbols: no growth.) 

3. A model of the type of Le Marc et al. (2002; Equation 3.25 to Equation 
3.27), where T max = 49.23°C (to be consistent with the logistic regression 
model parameter), a wmm = 0.948, and T mhl = 8.8°C, based on the minimum 
water activity and temperature, respectively, at which growth were 
observed 

4. A model of the type of Augustin and Carlier (2000b; Equation 3.24) 
assuming that T min = 8.8°C and a wmin = 0.948, consistent with the param- 
eter values used for the Le Marc et al. (2002) model 




The predicted interfaces are shown in Figure 3.1 1, together with the data used 
to generate the models. (Note that the subsets of 143 of the 179 data of Salter et al. 
(2000) used by Hajmeer and Basheer (2003a) to t the PNN and the Equation 3.44 
type model were not identi ed.) 

When compared to the full data set, the level of misprediction ranged from ~15 
to 20 of the 179 data points for each of the models, suggesting that the level of 
performance was not greatly different despite very different modeling approaches.* 
A complication in the comparison of G/NG model performance is that most of the 

* It should be noted that this analysis disagrees with the results of Hajmeer and Basheer (2002) who 
reported only two to four mispredictions for the total (i.e., training and validation) data set. 



2004 by Robin C. McKellar and Xuewen Lu 













~V 






1237_C03.fm Page 125 Wednesday, November 12, 2003 12:40 PM 






data are readily predicted, e.g., those that fall outside the known limits for growth 
for individual environmental factors when all other factors are optimal. Such data 
can "overwhelm" the data in which one is most interested, i.e., in the relatively 
narrow region where factors interact to reduce the biokinetic ranges and, yet more 
specifically, where the probability of growth rapidly changes from "growth is very 
likely" to "growth is very unlikely." These data de ne the interface and, consequently, 
data closest to the interface are more important when comparing model performance. 
This has implications for experimental design, as discussed in Section 3.4.4 below. 

To assess whether one model might be preferred on theoretical grounds, as 
adjudged by its ability to extrapolate reliably, the predictions of all models in the 
temperature range above 35°C can be compared to data subsequently generated, 
shown in Figure 3.11, and not used to generate the models. Clearly, none of the 
models extrapolate well. 

From the above comparison, it appears that despite very different modeling 
approaches and degrees of complexity of modeling, there is currently little to dif- 
ferentiate those approaches on the basis of their ability to describe the G/NG interface 
or on their ability to predict outside the interpolation region. 

3.4.4 Experimental Methods and Design Considerations 

As suggested above, currently there is little mechanistic understanding of how 
environmental factors interact to prevent bacterial growth and it must be recognized 
as a possibility that there is no single, common mechanism underlying the observed 
boundaries for different factor combinations. Consequently, it is not possible from 
rst principles to design the optimal experiment that captures the essential informa- 
tion that will characterize the response and lead to reliable models. Instead, at this 
time, experimental methods must be focused toward gaining enough data in the 
interface region to be able to describe empirically the limits to growth. 

First of all, two approaches may be distinguished that could affect the experi- 
mental methods chosen. In one, the interest is in whether growth/toxin production, 
etc. is possible within some specified time limit, which may be related to the shelf 
life of the product. In other approaches, the objective is to de ne absolute limits to 
growth, i.e., the most extreme combinations of factors that just allow growth. McKel- 
lar and Lu (2001) argue that there is always a time limit imposed on G/NG modeling 
studies. Strategies exist, however, that provide greater con dence that the "absolute" 
limits to growth are being measured. Some of these are discussed below. 

3.4.4.1 Measuring Both Growth and Inactivation 

Several groups have assessed both growth and inactivation in their experimental 
treatments (McKellar and Lu, 2001; Parente et al., 1998; Presser et al., 1999; 
Razavilar and Genigeorgis, 1998). In this way the position of the boundary is inferred 
from two "directions." If growth is not observed, an observer cannot be sure whether 
growth is not possible or has not occurred yet. If it is known that at some more 
extreme condition inactivation occurs, it can be inferred that the G/NG boundary 
lies between those two sets of conditions. 




2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 126 Wednesday, November 12, 2003 12:40 PM 








A potential problem with this strategy is that cultures can initially display some 
loss of viability, but with survivors eventually initiating growth; i.e., population 
decline cannot unambiguously be interpreted as "growth is not possible." Numerous 
studies (e.g., Mellefont et al., 2003) have demonstrated that rapid transfer of a culture 
from one set of conditions to another that is more stressful can induce injury and 
death, but that survivors will eventually adjust and be able to grow. This has been 
termed the Phoenix phenomenon (Shoemaker and Pierson, 1976). Such regrowth 
has been reported in the context of G/NG modeling (Bolton and Frank, 1 999; Masana 
and Baranyi, 2000b; Parente et al., 1998; Tienungoon et al., 2000). 

Clearly, an experimenter interested in determining the "absolute" G/NG bound- 
ary will need to maximize the resistance of the inoculum to stress on exposure to 
the new, more stressful, environment. The use of stationary phase cultures as inocula 
would seem to be a minimum requirement. It may be necessary to habituate cultures 
to the test conditions (e.g., growth at conditions just less harsh) prior to inoculation 
into the test conditions to maximize the chance that growth, if possible, will be 
observed. One way to maximize the likelihood of observing the most extreme growth 
limits would be to use cultures growing at the apparent limits as inocula into slightly 
more stringent conditions. This also has the advantage of minimizing growth lags 
on inoculation into a harsher environment. 

3.4.4.2 Inoculum Size 

Masana and Baranyi (2000b) indicated that inoculum size affected the position of 
the boundary. Robinson et al. (1998) reported similar effects of inoculum density 
on bacterial lag times. While it is clear that time to detection would depend on 
inoculum density, growth detection methods were not cell-density -dependent in 
those studies. Parente et al. (1998) also reported that a decrease in inoculum size 
decreased the probability of survival. If the shock of transfer is known to inactivate 
a xed proportion of the cells in the inoculum, to develop a robust model it will be 
necessary to use an inoculum that ensures that even after inactivation there is a high 
probability that at least one cell will survive. 

The above observations lend support to the hypothesis that it is the distribution 
of physiological states of readiness to survive and multiply in a new environment 
that determines the position of the G/NG boundary, i.e., all other things being equal, 
the more cells in the inoculum the more likely it is that there is one cell that has 
the capacity to survive and grow. This also reinforces the equivalence between G/NG 
boundary modeling and the modeling of conditions that lead to in nite lag times. 
The importance of the distribution of lag times on the observed lag times of whole 
populations has been discussed by Baranyi (1998). 

There may be more involved reasons for inoculum density -dependent responses 
also, such as chemical messaging between cells (see, e.g., Miller and Bassler, 2001 ; 
Winans and Bassler, 2000). 

In conclusion, if the aim is to determine absolute limits to growth, a higher 
number of cells is preferable. Masana and Baranyi (2000b) stated that growth 
boundaries "represented the chance of growth for each sample; therefore, to assure 
a low probability of growth in many samples, it will be more relevant to consider 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 127 Wednesday, November 12, 2003 12:40 PM 




~V 








boundaries for high inoculum levels." Equally, as noted above, steps to maximize 
the cell's chances of survival and growth in the environment are also recommended. 
There is potentially a caveat, however, that needs to be applied. Maximum 
population densities of batch cultures are reported to decline under increasingly 
harsh growth conditions. Thus, the use of high inocula may mask the true position 
of the G/NG boundary if the inoculum used is already denser than the MPD of the 
organism in a very stressful test environment. 

3.4.4.3 Are There Absolute Limits to Microbial Growth? 

In the above discussion it has been implicitly assumed that there are absolute limits 
to microbial growth under combined environmental stresses. It is pertinent to exam- 
ine this assumption. 

Numerous authors have noted that, within an experiment, the transition between 
conditions that allow growth, and those that do not, is abrupt and that usually all 
replicates at the last growth condition grow, while all the replicates at the rst-growth- 
preventing condition do not (Masana and Baranyi, 2000; Presser et al. , 1 998; Tienungoon 
et al., 2000). McKellar and Xu (2001), for example, reported that of 1820 conditions 
tested, all ve replicates of each condition either grew or did not grow. This abruptness, 
however, is not always evident in the modeled results (Tienungoon et al., 2000). 

Conversely, between experiments by the same researcher, using the same meth- 
ods and the same strain, results are not always reproducible. Figure 3.9 provides an 
example and Masana and Baranyi (2000b) make the same observation of their data 
for Brochothrix thermosphacta. At the same time, however, there is evidence of 
excellent reproducibility of boundaries between independent workers, using different 
strains, and different methods in different locations. The results of Tienungoon et 
al. (2000) were highly consistent between two strains tested, and more notably, with 
those of George et al. (1988) and Cole et al. (1990) presented a decade earlier, 
including different strains in one case. There is also a remarkable similarity between 
the pH/temperature G/NG interface of Listeria innocua reported by Le Marc et al. 
(2002) and the same interface for two species of L. monocytogenes presented in 
Tienungoon et al. (2000). 

Jenkins et al. (2000) noted that the boundary they derived for the growth limits 
of Zygosaccharomyces baillii in beverages was very consistent with a model devel- 
oped 30 years earlier for the stability of acidic sauces. 

Stewart et al. (2002) noted that with S. aureus, as conditions became increasingly 
unfavorable for growth, the contour lines (P grow th) they generated drew closer and 
closer together, suggesting that conditions were approaching absolute limits that do 
not allow growth. Conversely, there are examples where one group's observations 
do not agree well with another's for an analogous organisms/environmental pair 
(e.g., Bolton and Frank, 1999; McKellar and Xu, 2001). Delignette-Muller and Rosso 
(2000) reported strain variability in the minimum temperature for growth. 

While the above discussion points to heterogeneity in the physiological readiness 
of bacteria to grow in a new environment, Masana and Baranyi (2000b) also infer 
that differences in microenvironments, particularly within foods, could also be a 
source of heterogeneity in observed growth limits. 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 128 Wednesday, November 12, 2003 12:40 PM 




~V 








In conclusion, there is a body of experimental evidence that suggests that growth 
boundaries, if carefully determined, might be highly reproducible. Conversely, coun- 
terexamples exist. It remains to be determined whether the incongruous results arise 
from signi cant and measurable differences in methodology, e.g., the detection time 
used in the respective studies, or are due to uncontrollable sources (Table 3.7). 

3.4.4.4 Experimental Design 

As noted above, it is not possible from rst principles to design the optimal exper- 
iment that captures the information to characterize the G/NG boundary. Various 
authors have suggested physiological interpretations (Battey et al, 2001; Battey and 
Schaffner, 2001; Jenkins et al., 2000; Lopez-Malo et al., 2000; McMeekin et al., 
2000) but none have yet been experimentally tested. 

Thus, an empirical approach that aims to collect as much information in the 
region of most interest, i.e., the G/NG interface, is recommended by most workers. 
Several groups of researchers have indicated that they use a two-stage modeling 
process. The rst uses a coarse grid of conditions of variables to roughly establish 
the position of the boundary. The second phase monitors responses at conditions 
near the boundary and at close intervals of the environmental parameters. Variable 
combinations far from the interface, at which growth is either highly likely or highly 
unlikely, do not provide much information to the modeling process, which seeks to 
de ne the interface with a high level of precision. Equally, it is ideal to use a design 
that gives roughly equal numbers of conditions where growth is, and growth is not, 
observed (Jenkins et al., 2000; Legan et al., 2002; Masana and Baranyi, 2000b, Uljas 
et al., 2001). Pragmatically, Legan et al. (2002) recommend setting up "marginal" 
and "no-growth" treatments rst because these treatments will run for the longest 
time (possibly several months). Those conditions in which growth is expected to be 
relatively quick can be set up last because they only need monitoring until growth 
is detected. 

The nature of these studies necessarily involves long incubation times. Legan 
et al. (2002) noted that particular care must be taken to ensure that the initial 
conditions do not change over time solely as a result of an uncontrolled interaction 
with the laboratory environment. Prevention of dessication or uptake of water vapor 
requires particular attention. Changes resulting from microbial activity may, how- 
ever, be an important part of the mechanism leading to growth initiation and should 
not be stabilized at the expense of growth that would naturally occur in a food. 
Legan et al. (2002) comment that, for example, maintaining the initial pH over time 
is typically neither possible nor practical, even in buffered media, and that allowing 
a change in pH due to growth of the organism more closely mimics what would 
happen in a food product than maintaining the initial pH over time. 

3.4.4.5 Conclusion 

From the above discussion, unambiguous de nition of the G/NG boundary of an 
organism in multidimensional space presents several paradoxical challenges. While 
an experimenter will do well to remember these considerations in the interpretation 



2004 by Robin C. McKellar and Xuewen Lu 











1237_C03.fm Page 129 Wednesday, November 12, 2003 12:40 PM 








of his/her results, it seems probable that methods that have been used to date will 
have come close to identifying the "true" G/NG boundary, and that the position of 
the boundary will move only slightly if an experimenter acts to control all of the 
above variables and to maximize the potential for the observation of growth in the 
chosen experimental system. 

While the discussion has not focused specifically on appropriate methods for 
probability of growth within a defined time, many of the same principles and 
considerations will apply. 

Moreover, the field of growth limits modeling, while having an equally long 
history as kinetic modeling, now seems to be quite disjointed, with little rigorous 
comparison of approaches, let alone agreement on the most appropriate model 
structures or experimental methods. In particular, the earlier work in probability 
modeling seems to have been ignored by some more contemporary workers, without 
reasons being indicated. 

The results of G/NG studies are clearly of great interest to food producers and 
food safety managers. It is perhaps time, then, that the G/NG modeling community 
seeks to find common ground and to begin to develop a rigorous framework for the 
development, and interpretation, of growth limits studies. 

APPENDIX A3.1 — CHARACTERIZATION OF 
ENVIRONMENTAL PARAMETERS AFFECTING 
MICROBIAL KINETICS IN FOODS 

A3. 1.1 Temperature 

In most situations, temperature is the major environmental parameter in uencing 
kinetics of microorganisms in food and its effect is included in most predictive 
microbiology models. During processing, storage, and distribution the temperature 
of foods can vary substantially, frequently including periods of temperature abuse 
for chilled foods (see, e.g., Audits International, 1999; James and Evans, 1990; 
O'Brien, 1997; Sergelidis et al., 1997). Thus, it is an important property of secondary 
models that they can predict the effect of changing temperatures on microbial 
kinetics and application of these models relies on information about product tem- 
perature and its possible variation over time. Numerous types of thermometers, 
temperature probes, and data loggers are available (McMeekin et al, 1993, pp. 
257-269; seagrant.oregonstate.edu/extension/ sheng/loggers.html) to measure the 
temperature of foods or food processing equipment. Infrared non-contact thermom- 
eters are often appropriate for foods but their use is limited for process equipment 
with stainless surfaces. 

A3. 1.2 Storage Atmosphere 

Foods are typically stored aerobically, vacuum packed, or by using modi ed atmo- 
sphere packing (MAP). "Controlled atmosphere packaging" can be considered a 
special case of MAP. MAP foods are exposed to an atmosphere different from both 



2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 130 Wednesday, November 12, 2003 12:40 PM 








air and vacuum packed usually involving mixtures of the gasses carbon dioxide 
(C0 2 ), nitrogen (N 2 ), and oxygen (0 2 ). 

2 and C0 2 in uence growth of most microorganisms and secondary predictive 
models must take their effect into account. The solubility of 2 in water, and thereby 
into the water phase of foods, is low (~0.03 1/1) but it can be important for growth 
and metabolism of microorganisms in both aerobic and MAP-stored products 
(Dainty and Mackey, 1992). Numerous techniques and instruments are available to 
determine 2 in the gas phase or dissolved in food. Microelectrodes to determine 
gradients of dissolved 2 in foods are available (www.instechlabs.com/oxy- 
gen.html; www.microelectrodes.com/) but models to predict the effect of such 
gradients remain to be developed. To account for the effect of aerobic or vacuum 
packed storage of foods a categorical approach has been used within predictive 
microbiology. For aerobic conditions growth media with access to air have been 
agitated. For vacuum packed foods microorganisms typically have been grown 
under 100% N 2 . 

C0 2 inhibits growth of some microorganisms substantially and, to predict micro- 
bial growth in MAP foods, it is important to determine the equilibrium concentration 
in the gas phase or the concentration of C0 2 dissolved into the foods water phase. 
At equilibrium, the concentration of C0 2 dissolved into the water phase of foods 
is proportional to the partial pressure of C0 2 in the atmosphere surrounding the 
product. Henry's law (Equation A3.1) provides a good approximation for the solu- 
bility of C0 2 . 




CO A— =K H PC0 2 (A3.1) 



In Equation A3. 1, K H is Henry's constant (mg/l/atm) and pC0 2 is the partial pressure 
(atm) of C0 2 . Between and 160°C the temperature dependence of the Henry's 
constant can be predicted by Equation A3. 2: 



r-i _^_ -i 



K H (mg - 1 - atm ) = 

101325-2.4429 ( A3 - 2 ) 



exp(-6.8346 + 1 .2817 ■ 10 4 / K - 3.7668 • 10 6 / K 2 + 2.997 • 10 3 / K 3 ) 

where K is the absolute temperature (Carroll et al., 1991). Those authors expressed 
K u as MPa/mole fraction. In Equation A3. 2 the constants 101,325 Pa/atm and 2.4429 
was used to convert this unit into mg C0 2 /1 H 2 0/atm. 

For MAP foods in exible packaging the partial pressure of C0 2 is conveniently 
determined from the percentage of C0 2 inside the pack. A range of analytical 
methods is available to determine C0 2 concentration in gas mixtures or concentra- 
tions of dissolved C0 2 (Dixon and Kell, 1989; www.pbi-dansensor.com/Food.htm). 

As shown from Equation A3.1 and Equation A3. 2, the concentration of C0 2 
dissolved in the water phase of a MAP food with 50% C0 2 in the headspace gas at 
equilibrium is 1.67 g/1 at 0°C and 1.26 g/1 at 8°C. Because of the high solubility of 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 131 Wednesday, November 12, 2003 12:40 PM 







C0 2 in water the gas composition in the headspace of MAP foods changes after 
packaging. The equilibrium gas composition is in uenced by several factors, e.g., 
the percentage of C0 2 in the initial headspace gas (%C0 2 Imtial ), the initial gas/product 
volume ratio (G/P), temperature, pH, lipids in the food, respiration of the food, and, 
of course, permeability of the packing lm. Different mass-balance equations to 
predict the rate of adsorption and solubility of C0 2 have been suggested (Devlieghere 
et al, 1998; Dixon and Kell, 1989; Gill, 1988; Lowenadler and Ronner, 1994; 
Simpson et al., 2001 a,b; Zhao et al., 1995). In chilled foods the rate of absorption 
of C0 2 is rapid compared to growth of microorganisms. Therefore, to predict micro- 
bial growth in these MAP foods it is suf cient to take into account the equilibrium 
concentration of C0 2 . 

Devlieghere et al. (1998) suggested Equation A3. 3 to predict the concentration 
of C0 2 in the water phase as a function of %C0 2 Imtial and G/P. In Equation A3. 3, 
d C02 is the density of C0 2 (1.976 g/1). 



/ 



V 



G 
P 



\ 



■ dC0 2 + K H 



) 



(G 



\(~ir} "1 Equilibrium 

L 2-1 aqueous 



V 



P 



■ dC0 2 + K H 



) 



( 



4 



V 



100 



K H —- %CO* nitial ■ dC0 2 



P 



J 




2 



(A3.3) 




Equation A3. 3 does not take into account the effect of the storage temperature and 
Devlieghere et al. (1998) developed a polynomial model to predict the concentration 
of dissolved C0 2 as a function of %C0 2 Imtial , G/P, and temperature. If, for example, 
%C0 2 Imtial is 25, the polynomial model predicts that a G/P ratio of three results in 
higher concentration of dissolved C0 2 than does a G/P ratio of 4, which is not 
logical. In contrast we have found that the combined use of Equation A3. 2 and 
Equation A3. 3 provides realistic predictions for concentrations of dissolved C0 2 . It 
also seems relevant to include the effect of product pH on dissolved C0 2 , and thereby 
the equilibrium concentration of C0 2 in the gas phase of MAP foods. 



A3. 1.3 Salt, Water-Phase Salt, and Water Activity 

While temperature is the single most important storage condition in uencing growth 
of microorganisms in foods, NaCl is the most important product characteristic in 
many foods. The concentration of NaCl in foods can be determined as chloride by 
titration (Anon., 1995a). Instruments to determine NaCl indirectly from conductivity 
measurements are available but extensive calibration for particular types of products 
may be required. In fresh and intermediate moisture foods, NaCl is dissolved in the 
water phase of the products. 

To predict the effect of NaCl on growth of microorganisms in these products 
the concentration of water-phase salt (WPS) or relative humidity, i.e., the water 
activity (a w ) must be determined (Equation A3. 4 to Equation A3. 7). 

Water-phase salt can be calculated from Equation A3. 4: 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 132 Wednesday, November 12, 2003 12:40 PM 







% Water phase salt = 
%NaCl (w/v) x 100/(100 - % dry matter +%NaCl (w/v)) (A3.4) 

Water activity is a fundamental property of aqueous solutions and is de ned as: 



a = -5- (A3.5) 



w 



Po 




where p is the vapor pressure of the solution and p is the vapor pressure of the 
pure water under the same conditions of temperature, etc. 

For mixtures of NaCl and water there is a direct relation between the WPS 
content and a w (Resnik and Chirife, 1988; Equation A3. 6 and Equation A3. 7). For 
cured foods where NaCl is the only major humectant these relations are valid as 
documented, e.g., for cold-smoked salmon (Jorgensen et al., 2000) and processed 
"delicatessen" meats (Ross and Shadbolt, 2001). To determine water activity of 
foods, instruments relying on the dew point method are now widely used because 
of their speed (providing results within a few minutes), robustness, and reliability 
but other methods and instruments are available (Mathlouthi, 2001). 

a w = 1-0.0052411 -%WPS- 0.00012206 -%WPS 2 (A3.6) 

% WPS =8-140.01 -(a -0.95) -405. 12 -(a -0.95) 2 (A3.7) 




w / v w 



A3.1.4 pH 

For many microorganisms, small pH variations in the pH range ~6 to -7 have very 
little or no effect on population kinetics. In more acidic foods, however, pH per se 
can greatly in uence microbial kinetics but can also accentuate the effect of other 
added preservative compounds. The pH of solid foods is often determined by homog- 
enizing 10 g of a sample with 10 to 20 ml of distilled water and measuring the pH 
of the suspension using a standard combined electrode. 

A3. 1.5 Added Preservatives Including Organic Acids, 
Nitrate, and Spices 

High concentrations of organic acids occur naturally in some foods and various 
organic acids including acetic acid, ascorbic acid, benzoic acid, citric acid, lactic 
acid, and sorbic acid are frequently added to foods. Organic acids can inhibit growth 
of microorganisms markedly and secondary models to predict their inhibitory effect 
are frequently needed. As for NaCl the secondary models must take into account 
the concentration of organic acids in the water phase of products. In addition, 
secondary models may need to describe the combined effect of organic acids and 
other environmental parameters particularly the pH. 

2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 133 Wednesday, November 12, 2003 12:40 PM 








In solution, organic acids exist either as the dissociated (ionized) or undissociated 
species. The Henderson-Hasselbalch equation (Equation A3. 8) relates the proportion 
of undissociated and dissociated forms of organic acid to pH and pK a according to 
the following expression: 

[A-]/[HA] = 10 H -P K a (A3.8) 

where [HA] is the concentration of undissociated form of the acid, [A~] the concen- 
tration of dissociated (ionized) form of the acid, and pK a is the pH at which the 
concentrations of the two forms are equal. 

While both the dissociated and the undissociated forms of organic acids have 
inhibitory effects on bacterial growth the undissociated form is more inhibitory, 
usually by two to three orders of magnitude, than the dissociated form (Eklund, 1989). 

Cross-multiplying and rearranging Equation A3. 8 to solve for [HA] gives: 

[HA] = [LAC]/(1 + 10P H "P K a) (A3. 9) 

where [LAC] is the total lactic acid concentration and all other terms are as previously 
de ned. 

As the concentration of an undissociated acid increases the growth rate of 
microorganisms decreases, eventually ceasing completely at a level described as the 
MIC. This behavior, and its dependence on the interaction of pH and total organic 
acid concentration, is included explicitly in several secondary models (Augustin and 
Carrier, 2000a; Presser et al., 1997). 

Simple enzyme kits are available to determine several of the organic acids that 
are important in foods. Simultaneous determination of a range of organic acids is 
possible by HPLC analysis and is often an appropriate method to use (Dalgaard and 
Jorgensen, 2000; Pecina et al., 1984). 

Nitrite can be added to some types of meat products and its concentration in the 
water phase of products must be taken into account when secondary predictive 
models for these products are developed. Colorimetric methods are available to 
measure the concentration of nitrite in foods (Anon., 1995b; Karl, 1992). 

Spices and herbs can have substantial antimicrobial activity and appropriate 
terms may need to be included in secondary models (Koutsoumanis et al., 1999; 
Skandamis and Nychas, 2000). The concentration of active antimicrobial compo- 
nents in spices, herbs, and essential oils can vary substantially as a function, e.g., 
of geographical region and season (Nychas and Tassou, 2000; Sofos et al., 1998). 
Therefore, the development of accurate secondary predictive models most likely will 
have to rely on the concentration of their active antimicrobial components. Recently, 
Lambert et al. (2001) showed the antimicrobial effect of the oregano essential oil 
quantitatively corresponded to the effect of its two active components, i.e., thymol 
and carvacrol. To quantitatively determine active components in spices, herbs, and 
essential oils appropriate extracts can be analyzed by GC/MS techniques (Cosentino 
et al., 1999; Cowan, 1999). 




2004 by Robin C. McKellar and Xuewen Lu 













~V 






1237_C03.fm Page 134 Wednesday, November 12, 2003 12:40 PM 






A3. 1.6 Smoke Components 

It has long been known that high concentrations of smoke components have strong 
antimicrobial activity (Shewan, 1949). Today many meat and seafood products are 
smoked but typically less intensively than some decades ago. However, even mod- 
erate concentrations of smoke components can in uence growth rates, growth 
limits, and rates of death/inactivation of microorganisms in foods (Leroi et al., 
2000; Leroi and Joffraud, 2000; Ross et al., 2000b; Sunen, 1998; Thurette et al., 
1998). Thus, to obtain accurate prediction of microbial kinetics in smoked foods 
it is important to include terms for the effect of smoke components in secondary 
models. Phenols are important antimicrobials in wood smoke, or in liquid smokes, 
and a few secondary models include the total phenol concentration as an environ- 
mental parameter (Augustin and Carlier 2000a,b; Gimenez and Dalgaard, in press; 
Membre et al., 1997). 

Classical colorimetric methods can be used to determine the total concentration 
of phenols in smoked foods. These methods rely on formation of colored complexes, 
e.g., between phenols and Gibb's reagent (2,6-dichloroquinone-4-chloroimide) or 4- 
aminoantipyrine (Leroi et al., 1998; Tucker, 1942). The total phenol concentration 
is a crude measure of how intensely foods have been smoked. By using GC/MS 
techniques more detailed information about specific smoke components can be 
obtained (Guillen and Errecalde, 2002; McGill et al., 1985; Toth andPotthast, 1984). 
In the future, secondary models may be developed to include the effect of specific 
phenols, other specific smoke components, and possibly their interaction with NaCl. 
During the smoking of foods, phenols and other smoke components are mainly 
deposited in the outer 0.5 cm of the product (Chan et al., 1975). Modeling the effect 
of the spatial distribution in foods is another challenge. 

A3. 1.7 Other Environmental Parameters 

The environmental parameters discussed above include those that are of major 
importance in traditional methods of food preservation. Many modern methods of 
food preservation also rely on combinations of these environmental parameters. 
However, the effect of a few well-known and several emerging food processing 
technologies relies on the antimicrobial effect of other environmental parameters, 
e.g., bacteriocins, gamma irradiation, high electric field pulses, high pressure, and 
UV light. Secondary models for the effect of some of these environmental parameters 
have been developed but will not be discussed here in detail. Other environmental 
parameters related to food structure and to the effect of microbial metabolism on 
changes in environmental parameters are discussed in Chapter 5 whereas the effect 
of time- varying environmental parameters is discussed in Chapter 7. 




2004 by Robin C. McKellar and Xuewen Lu 










1237_C03.fm Page 135 Wednesday, November 12, 2003 12:40 PM 








REFERENCES 

Abdullah, B., Gardini, F., Paparella, A., and Guerzoni, ME. Growth modelling of the pre- 
dominant microbial groups in hamburgers in relation to modulation of atmosphere 

composition, storage temperature, and diameter of meat particle. Meat Sci., 38, 

511-526, 1994. 
Adair, C, Kilsby, D.C., and Whittall, P.T. Comparison of the Schoolfield (non-linear Arrhe- 

nius) model and the square root model for predicting bacterial growth in foods. Food 

Microbiol, 6, 7-18, 1989. 
Adams, M.R., Little, C.L., and Easter, M.C. Modelling the effect of pH, acidulant and 

temperature on growth of Yersinia enterocolitica. J. Appl. Bacteriol, 71, 65-71, 1991. 
Anon. Salt (chlorine as sodium chloride) in seafood. Volumetric method. In Of cial Methods 

of Analysis, AOAC, Arlington, VA, 1995a, p. 937.09. 
Anon. Nitrites in cured meat. Colorimetric method. In Of cial Methods of Analysis, AOAC, 

Arlington, VA, 1995b, p. 937.31. 
Audits International, US Food Temperature Evaluation. Audits International, Northbrook, IL, 

1999. http://www.foodriskclearinghouse.umd.edu/audits_international.htm. Febru- 
ary, 2002. 
Augustin, J.-C. and Carlier, V Mathematical modelling of the growth rate and lag time for 

Listeria monocytogenes. Int. J. Food Microbiol, 56, 29—51, 2000a. 
Augustin, J.-C. and Carlier, V Modelling the growth of Listeria monocytogenes with a 

multiplicative type model including interactions between environmental factors. Int. 

J. Food Microbiol, 56, 53-70, 2000b. 
Augustin, J.-C, Rosso, L., and Carlier, V A model describing the effect of temperature history 

on lag time fox Listeria monocytogenes. Int. J. Food Microbiol, 57, 169-181, 2000. 
Bajard, S., Rosso, L., Fardel, G., and Flandrois, J.R The particular behaviour of Listeria 

monocytogenes under sub-optimal conditions. Int. J. Food Microbiol. , 29, 201-211, 

1996. 
Baker, D.A. Probability models to assess the safety of foods with respect to Clostridium 

botulinum. J. Ind. Microbiol, 12, 156—161, 1993. 
Baker, D.A. and Genigeorgis, C. Predicting the safe storage of fresh sh under modi ed 

atmospheres with respect to Clostridium botulinum toxi genesis by modelling length 

of the lag phase of growth. J. FoodProt, 53, 131-140, 1990. 
Baker, D.A. and Genigeorgis, C. Predictive modeling. In Clostridium botulinum Ecology and 

Control in Foods, Eds. A.W.H. Hauschild and K.L. Dodds. Marcel Dekker, New York, 

1993, 343^106. 
Baker, D., Genigeorgis, C, Glover, J., and Razavilar, V Growth and toxigenesis of 

C botulinum type E in shes packaged under modi ed atmospheres. Int. J. Food 

Microbiol, 10,269-290, 1990. 
Baranyi, J. Comparison of stochastic and deterministic concepts of bacterial lag. J. Theor 

Biol, 192,403^108, 1998. 
Baranyi, J. and Roberts, T.A. A dynamic approach to predicting bacterial growth in food. Int. 

J. Food Microbiol, 23, 277-294, 1994. 
Baranyi, J. and Roberts, T.A. Mathematics of predictive food microbiology. Int. J. Food 

Microbiol, 26,199-218, 1995. 
Baranyi, J., Ross, T., McMeekin, T.A., and Roberts, T.A. Effects of parametrization on the 

performance of empirical models used in "predictive microbiology." Food Microbiol. , 

13. 83 -91, 1996. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 136 Wednesday, November 12, 2003 12:40 PM 








Battey, A.S. and Schaffner, D.W. Modelling bacterial spoilage in cold- lied ready to drink 
beverages by Acinetobacter calcoaceticus and Gluconobacter oxydans. J. Appl. 
Microbiol, 91, 237-247, 2001. 

Battey, A.S., Duffy, S., and Schaffner, D.W. Modelling mould spoilage on cold- lied ready- 
to-drink beverages by Aspergillus niger and Penicillium spinulosum. Food Microbiol. , 
18,521-529,2001. 

Battey, A.S., Duffy, S., and Schaffner, D.W. Modeling yeast spoilage in cold- lied ready-to- 
drink beverages with Saccharomyces cerevisiae, Zygosaccharomyces bailii, and Can- 
dida lipolytic a. Appl. Environ. Microbiol. , 68, 1901—1906, 2002. 

Benedict, R.C., Pertridge, T., Wells, D., and Buchanan, R.L. Bacillus cereus: aerobic growth 
kinetics. J. FoodProt, 56, 211-214, 1993. 

Bernaerts, K., Versyck, K.J., and van Impe, J.F. On the design of optimal experiments for 
parameter estimation of a Ratkowsky-type growth kinetic at suboptimal temperatures. 
Int. J. Food Microbiol. , 54, 27-38, 2000. 

Bhaduri, S., Buchanan, R.L., and Phillips, J.G. Expanded response surface model for predict- 
ing the effects of temperature, pH, sodium chloride contents and sodium nitrite 
concentration on the growth rate of Yersinia enterocolitica. J. Appl. Bacteriol. , 79, 
163-170, 1995. 

Bolton, L.F. and Frank, J.F. De ning the growth/no growth interface for Listeria monocyto- 
genes in Mexican-style cheese based on salt, pH and moisture content. J. FoodProt., 
62, 601-609, 1999. 

Bouttefroy, A., Mansour, M., Linder, M., and Milliere, J.B. Inhibitory combinations of nisin, 
sodium chloride, and pH on Listeria monocytogenes ATCC 15313 in broth by an 
experimental design approach. Int. J. Food Microbiol, 54, 109-115, 2000. 

Broughall, J.M. and Brown, C. Hazard analysis applied to microbial growth in foods: devel- 
opment and application of three-dimensional models to predict bacterial growth. Food 
Microbiol, 1, 13-22, 1984. 

Broughall, J.M., Anslow, P., and Kilsby, D.C. Hazard analysis applied to microbial growth 
in foods: development of mathematical models describing the effect of water activity. 
J. Appl Bacteriol, 55, 101-110, 1983. 

Buchanan, R.L. Developing and distributing user- friendly application software. J. Ind. Micro- 
biol, 12,251-255, 1993a. 

Buchanan, R.L. Predictive food microbiology. Trends Food Sci. Technol, 41, 6-11, 1993b. 

Buchanan, R.L. and Bagi, K. Expansion of response surface models for the growth of 
Escherichia coli 0157:H7 to include sodium nitrite as a variable. Int. J. Food Micro- 
biol, 23, 317-332, 1994. 

Buchanan, R.L. and Phillips, J.G. Updated models for the effects of temperature, initial pH, 
NaCl, and NaN0 2 on aerobic and anaerobic growth of Listeria monocytogenes. Quant. 
Microbiol, 2, 103-128, 2000. 

Buchanan, R.L., Smith, J.L., McColgan, C, Marmer, B.S., Golden, M., and Dell, B. Response 
surface models for the effects of temperature, pH, sodium chloride, and sodium nitrite 
on the aerobic and anaerobic growth of Staphylococcus aureus 196E. J. Food Saf, 
13, 159-175, 1993. 

Buchanan, R.L., Golden, M.H., and Phillips, J.G. Expanded models for the non- thermal 
inactivation of Listeria monocytogenes. J. Appl Microbiol, 82, 561-511, 1997. 

Budavari, S. The Merck Index: An Encyclopaedia of Chemicals, Drugs, and Biological s, 11th 
ed., Merck and Co., Inc., Rahway, NJ, 1989. 

Carroll, J.J., Slupsky, J.D., and Mather, A. E. The solubility of carbon dioxide in water at low 
pressure. J. Phys. Chem. Ref. Data, 20, 1201-1209, 1991. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 137 Wednesday, November 12, 2003 12:40 PM 








Chan, W.S., Toledo, R.T., and Deng, J. Effect of smokehouse temperature, humidity and air 
ow on smoke penetration into sh muscle. J. Food Sci., 40, 240-243, 1975. 

Chandler, R.E. and McMeekin, T.A. Combined effect of temperature and salt concentra- 
tion/water activity on growth rate of Halobacterium spp. J. Appl. Bacteriol, 67, 
71-76, 1989. 

Chea, F.P., Chen, Y, Montville, T.J., and Schaffner, D.W. Modelling the germination kinetics 
of Clostridium botulinum 56A spores as affected by temperature, pH, and sodium 
chloride. J. FoodProt, 63, 1071-1079, 2000. 

Cheroutre-Vialette, M. and Lebert, A. Modelling the growth of Listeria monocytogenes in 
dynamic conditions. Int. J. Food Microbiol, 55, 201-207, 2000. 

Chorin, E., Thuault, D., Cleret, J.- J., and Bourgeois, C.-M. Modelling Bacillus cereus growth. 
Int. J. Food Microbiol. , 38, 229-234, 1997. 

Cole, M.B., Jones, M. V, and Holy oak, C. The effect of pH, salt concentration and temperature 
on the survival and growth of Listeria monocytogenes. J. Appl. Bacteriol, 69, 63-72, 
1990. 

Connil, N., Plissoneau, L., Onno, B., Pilet, M.-F., Prevost, H., and Dousset, X. Growth of 
Carnobacterium diver gens V41 and production of biogenic amines and divergicin in 
sterile cold- smoked salmon extract at varying temperatures, NaCl levels, and glucose 
concentrations. J. FoodProt, 65, 333-338, 2002. 

Cosentino, S., Tuberoso, C.I.G., Pisano, B., Satta, M., Mascia, Y, Arzedi, E., and Palmas, R 
In-vitro antimicrobial activity and chemical composition of Sardinian thymus essential 
oils. Lett. Appl Microbiol, 29, 130-135, 1999. 

Cowan, M.M. Plant products as antimicrobial agents. Clin. Microbiol. Rev., 12, 564-582, 
1999. 

Cuppers, H.G.A., Oomes, S., and Brul, S. A model for the combined effects of temperature 
and salt concentration on growth rate of food spoilage molds. Appl Environ. Micro- 
biol, 63, 3764-3769, 1997. 

Dainty, R.H. and Mackey, B.M. The relationship between the phenotypic properties of bacteria 
from chill-stored meat and spoilage processes. J. Appl Bacteriol, 73, 103S-114S, 
1992. 

Dalgaard, P. Evaluation and Prediction of Microbial Fish Spoilage, Ph.D. thesis, Technolog- 
ical Laboratory, Danish Ministry of Fisheries, Lynby, Denmark, 1993. 

Dalgaard, P. Modelling of microbial activity and prediction of shelf life for packed fresh sh. 
Int. J. Food Microbiol. , 26, 305-317, 1995. 

Dalgaard, P. Modelling and prediction the shelf-life of seafood. In Safety and Quality Issues 
in Fish Processing, Bremner, H.A. (Ed.), Woodhead Publishing, Cambridge, England, 
2002, pp. 191-219. 

Dalgaard, P. and Jorgensen, L. V. Cooked and brined shrimps packed in a modi ed atmosphere 
have a shelf-life of >7 months at 0°C, but spoil at 4-6 days at 25°C. Int. J. Food Sci. 
Technol, 35, 431^142, 2000. 

Dalgaard, P., Ross, T., Kamperman, L., Neumeyer, K., and McMeekin, T.A. Estimation of 
bacterial growth rates from turbidimetric and viable count data. Int. J. Food Micro- 
biol, 23, 391^104, 1994. 

Dalgaard, P., Mejlholm, O., and Huss, H.H. Application of an iterative approach for devel- 
opment of a microbial model predicting the shelf-life of packed sh. Int. J. Food 
Microbiol, 38, 169-179, 1997. 

Dalgaard, P., Cowan, B. J., Heilmann, J., and Silberg, S. The seafood spoilage and safety 
predictor (SSSP). In Predictive Modeling in Foods — Conference Proceedings, Van 
Impe, J.F.M., Geeraerd, A.H., Leguerinel, I., and Mafart, P. (Eds.), Katholieke Uni- 
versiteit Leuven/BioTec, Belgium, 2003, pp. 256-258 . 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 138 Wednesday, November 12, 2003 12:40 PM 








Dantigny, P. Dimensionless analysis of the microbial growth rate dependence on sub-optimal 

temperatures. J. Ind. Microbiol. Biotechnol, 21, 215-218, 1998. 
Dantigny, P. and Molin, P. In uence of the modelling approach on the estimation of the 

minimum temperature for growth in Belehradek-type models. Food Microbiol, 17, 

597-604, 2000. 
Daughtry, B.J., Davey, K.R., and King, K.D. Temperature dependence of growth kinetics of 

food bacteria. Food Microbiol, 14, 21-30, 1997. 
Davey, K.R. A predictive model for combined temperature and water activity on microbial 

growth during the growth phase. J. Appl Bacteriol, 67, 483^188, 1989. 
Davey, K.R. Applicability of the Davey (linear Arrhenius) predictive model to the lag phase 

of microbial growth. J. Appl Bacteriol, 70, 253-257, 1991. 
Davey, K.R. Modeling the combined effect of temperature and pH on the rate coef cient for 

bacterial growth. Int. J. Food Microbiol , 23, 295-303, 1994. 
Davey, K.R. Belehradek models: application to chilled foods preservation. Refrig. Sci. Technol 

Proc, EUR 18816, 31-41, 1999. 
Davey, K.R. Models for predicting the combined effect of environmental process factors on 

the exponential and lag phases of bacterial growth: development and application and 

an unexpected correlation. In Proceedings of the 6th World Conference of Chemical 

Engineering, Melbourne, 2001. 
Davey, K.R. and Daughtry, B.J. Validation of a model for predicting the combined effect of 

3 environmental -factors on both exponential and lag phases of bacterial- growth — 

temperature, salt concentration and pH. Food Res. Int., 28, 233-237, 1995. 
Davies, K.W. Design of experiments for predictive microbial modeling. J. Ind. Microbiol, 

12, 295-300, 1993. 
Deak, T. and Beuchat, L.R. Use of indirect conductimetry to predict the growth of spoilage 

yeasts, with special consideration of Zagosaccharomyces bailii. Int. J. Food Micro- 
biol, 23, 405^17 ', 1994. 
Delignette-Muller, M.L. Relation between the generation time and the lag time of bacterial 

growth kinetics. Int. J. Food Microbiol. , 43, 97-104, 1998. 
Delignette-Muller, M.L. and Rosso, L. Biological variability and exposure assessment. Int. 

J. Food Microbiol, 58, 203-212, 2000. 
Delignette-Muller, M.L., Baty, F., Cornu, M., and Bergis, H. The effect of the temperature 

shifts on the lag phase of Listeria monocytogenes . In Predictive Modeling in Foods — 

Conference Proceedings, Van Impe, J.F.M., Geeraerd, A.H., Leguerinel, I., and 

Mafart, P. (Eds.), Katholieke Universiteit Leuven/BioTec, Belgium, 2003, pp. 

209-211. 
Dengremont, E. and Membre, J.M. Statistical approach for comparison of the growth rates 

of the strains of Staphylococcus aureus. Appl Environ. Microbiol, 61, 4389^1395, 

1997. 
Devlieghere, F., Debevere, J., and van Impe, J. Concentration of carbon dioxide in the water- 
phase as a parameter to model the effect of a modi ed atmosphere on microorganisms. 

Int. J. Food Microbiol. , 43, 105-113, 1998. 
Devlieghere, F., Geeraerd, A.H., Versyck, K.J., Bernaert, H., van Impe, J., and Debevere, J. 

Shelf life of modi ed atmosphere packed cooked meat products: addition of Na- 

lactate as a fourth shelf life determinative factor in a model and product validation. 

Int. J. Food Microbiol. , 58, 93-106, 2000a. 
Devlieghere, F., Lefevre, I., Magnin, A., and Debevere, J. Growth of Aeromonas hydrophila 

in modi ed-atmosphere-packed cooked meat products. Food Microbiol. , 17, 1 85-196, 

2000b. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 139 Wednesday, November 12, 2003 12:40 PM 








Devlieghere, F., Geeraerd, A. H., Versyck, K.J., Vandewaetere, B., van Impe, J., and Debevere, 
J. Growth of Listeria monocytogenes in modi ed atmosphere packed cooked meat 
products: a predictive model. Food Microbiol. , 18, 53-66, 2001. 

Dixon, N.M. and Kell, D.B. The control and measurement of "C0 2 " during fermentations. 
J.Microbiol. Methods, 10, 155-176, 1989. 

Dodds, K.L. An introduction to predictive microbiology and the development and use of 
probability models with Clostridium botulinum. J. Ind. Microbiol, 12, 139-143, 1993. 

Doe, RE. and Heruwati, E. A model for the prediction of the microbial spoilage of sun-dried 
tropical sh. J. FoodEng, 8, 47-72, 1988. 

Draper, N.R. and Smith, H. Applied Regression Analysis, 2nd ed., John Wiley & Sons, New 
York, 1981, 709 pp. 

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

Eklund, T. Organic acids and esters. In Mechanisms of Action of Food Preservation Proce- 
dures, Gould, G.W. (Ed.), Elsevier Science Publishers, London, 1989, pp. 160-200. 

Eyring, H. The activated complex in chemical reactions. J. Chem. Phys., 3, 107-115, 1935. 

Farber, J. M. Predictive modeling of food deterioration and safety. In Foodborne Microor- 
ganisms and Their Toxins: Developing Methodology, Pierson, M.D. and Stern, N.J. 
(Eds.), Marcel Dekker, New York, 1986, pp. 57-90. 

Fernandez, PS., Baranyi, J., and Peck, M.W. A predictive model of growth from spores of 
non-proteolytic Clostridium botulinum in the presence of different C0 2 concentrations 
as in uenced by chill temperature, pH and NaCl. Food Microbiol. , 1 8, 453^161 , 2001 . 

Ganzle, M.G., Ehmann, M., and Hammes, W.P Modeling of growth of Lactobacillus san- 
franciscensis and Candida milleri in response to process parameters of sourdough 
fermentation. Appl. Environ. Microbiol, 64, 2616-2623, 1998. 

Garcia-Gimeno, R.M., Hervas-Martinez, C, and de Siloniz, M.I. Improving arti cial neural 
networks with a pruning methodology and genetic algorithms for their application in 
microbial growth predictions in food. Int. J. Food Microbiol, 72, 19-30, 2002. 

Garcia-Gimeno, R.M., Hervas-Martinez, C, Barco-Alcaia, E., Zurera-Cosano, G., and Sanz- 
Tapia, E. An arti cial neural network approach to Escherichia coli 0157:H7 growth 
estimation. J. FoodSci., 68, 639-645, 2003. 

Geeraerd, A.H., Herremans. C.H., Cenens, C, and Van Impe J.F. Application of arti cial 
neural networks as a non-linear modular modeling technique to describe bacterial 
growth in chilled food products. Int. J. Food Microbiol. , 44, 49-68, 1998a. 

Geeraerd, A.H., Herremans, C.H., Ludikhuyze, L.R., Hendrickx, M.E., and Van Impe, J.F. 
Modeling the kinetics of isobaric-isothermal inactivation of Bacillus subtilis alpha- 
amylase with arti cial neural networks. J. Food Eng, 36, 263-279, 1998b. 

Geeraerd, A.H., Valdramidis, VP, Devlieghere, F., Bernaert, H., Debevere, J., and van Impe, 
J. Development of a novel modelling methodology by incorporating a priori micro- 
biological knowledge in a black box approach. Int. J. Food Microbiol, in press. 

Genigeorgis, C.A. Factors affecting the probability of growth of pathogenic microorganisms 
in foods. J. Am. Vet. Med. Assoc, 179, 1410-1417, 1981. 

Genigeorgis, C, Carniciu, M., Dutulescu, D., and Farver, T.B. Growth and survival of Listeria 
monocytogenes in market cheeses stored at 4°C to 30°C. J. Food Protect., 54, 
662-668, 1991. 

George, S.M., Lund, B.M., and Brocklehurst, T.F. The effect of pH and temperature on 
initiation of growth of Listeria monocytogenes. Lett. Appl Microbiol, 6, 153—156, 
1988. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 140 Wednesday, November 12, 2003 12:40 PM 








Gibson, A.M., Bratchell, N., and Roberts, T.A. The effect of sodium chloride and temperature 
on the rate and extent of growth of Clostridium botulinum type A in pasteurized pork 
slurry. J. Appl. Bacteriol, 62, 47SM90, 1987. 

Gibson, A.M., Bratchell, N, and Roberts, T.A. Predicting microbial growth: growth responses 
of salmonellae in a laboratory medium as affected by pH, sodium chloride and storage 
temperature. Int. J. Food Microbiol., 6, 155-178, 1988. 

Gibson, A.M., Baranyi, J., Pitt, J.I., Eyles, M.J., and Roberts, T.A. Predicting fungal growth: 
the effect of water activity on Aspergillus avus and related species. Int. J. Food 
Microbiol., 23, 419^131, 1994. 

Gill, CO. The solubility of carbon dioxide in meat. Meat Sci., 22, 65-71, 1988. 

Gimenez, B.C. and Dalgaard, P. Modelling and predicting the simultaneous growth of Listeria 
monocytogenes and spoilage microorganisms in cold-smoked salmon. J. Appl. Micro- 
biol., in press. 

Gorris, L. Hurdle technology. In Encyclopaedia of Food Microbiology, Robinson, R., Batt, 
C.A., and Patel, P. (Eds.), Academic Press, London, 2000, pp. 1071-1076. 

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

Guillen, M.D. and Errecalde, C. Volatile components of raw and smoked black bream (Brama 
rail) and rainbow trout (Oncorhynchus mykiss) studied by means of solid phase 
microextraction and gas chromatography /mass spectrometry. J. Sci. FoodAgric, 82, 
945-952, 2002. 

Hajmeer, M.N. and Basheer, I. A probabilistic neural network approach for modeling and 
classi cation of bacterial growth/no- growth data. J. Microbiol. Methods, 51, 217-226, 
2002. 

Hajmeer, M.N. and Basheer, LA. A hybrid Bayesian- neural network approach for probabilistic 
modeling of bacterial growth/no- growth interface. Int. J. Food Microbiol., 82, 
233-243, 2003a. 

Hajmeer, M.N. and Basheer, I. Comparison of logistic regression and neural network-based 
classi ers for bacterial growth. Food Microbiol., 20, 43-55, 2003b. 

Hajmeer, M.N., Basheer, LA., and Najjar, Y.M. Computational neural networks for predictive 
microbiology. 2. Application to microbial growth. Int. J. Food Microbiol. , 34, 51-66, 
1997. 

Heitzer, A., Kohler, H.E., Reichert, P., and Hamer, G. Utility of phenomenological models 
for describing temperature dependence of bacterial growth. Appl. Environ. Microbiol., 
57,2656-2665, 1991. 

Hinshelwood, C.N. In uence of temperature on the growth of bacteria. In The Chemical 
Kinetics of the Bacterial Cell, Clarendon Press, Oxford, 1946, 254-257. 

Hosmer, D.W. and Lemeshow, S . Applied Logistic Regression, John Wiley & Sons, New York, 
1989. 

Hultin, E. The in uence of temperature on the rate of enzymic processes. Acta Chem. Scand., 
9, 1700-1710, 1955. 

ICMSF (International Commission on Microbiological Specifications for Foods). Microor- 
ganisms in Foods. 5. Microbiological Specifications of Food Pathogens, Blackie 
Academic & Professional, London, 1996a. 

ICMSF. Modelling microbial responses in foods. In Microorganisms in Foods. 5. Microbio- 
logical Specifications of Food Pathogens, Roberts, T.A., Baird-Parker, A.C., and 
Tompkin, R.B. (Eds.), Blackie Academic & Professional, London, 1996b, pp. 
493-500. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 141 Wednesday, November 12, 2003 12:40 PM 








Ikawa, J.Y. and Genigeorgis, C. Probability of growth and toxin production by nonproteolytic 
Clostridium botulinum in rock sh llets stored under modi ed atmospheres. Int. J. 
Food Microbiol, 4, 167-181, 1987. 

James, S. and Evans, J. Performance of domestic refrigerators and retail display cabinets for 
chilled products. Refrig. Sci. Technol. Proc, 1990-1994, 401^09, 1990. 

Jenkins, P., Poulos, P.G., Cole, M.B., Vandeven, M.H., and Legan, J.D. The boundary for 
growth of Zygosaccharomyces bailii in acidi ed products described by models for 
time to growth and probability of growth. J. Food Prot., 63, 222-230, 2000. 

Jeyamkondan, S., Jay as, D.S., and Holley, R.A. Microbial growth modelling with arti cial 
neural networks. Int. J. Food Microbiol, 64, 343-354, 2001. 

Johnson, F.H. and Lewin, I. The growth rate of E. coli in relation to temperature, quinine and 
coenzyme. J. Cell Comp. Physiol, 28, 47-75, 1946. 

Jones, J.E. and Walker, S. Advances in modeling microbial growth. J. Ind. Microbiol, 12, 
200-205, 1993. 

Jones, J.E., Walker, S. J., Sutherland, J.P, Peck, M.W., and Little, C.L. Mathematical modelling 
of the growth, survival and death of Yersinia enterocolitica. Int. J. Food Microbiol. , 
23, 433^137, 1994. 

Jorgensen, L.V., Dalgaard, P., and Huss, H.H. Multiple compound quality index for cold- 
smoked salmon (Salmo salar) developed by multivariate regression of biogenic 
amines and pH. J. Agric. Food Chem., 48, 2448-2453, 2000. 

Juneja, V.K., Marmer, B.S., Phillips, J.G., and Palumbo, S.A. Interactive effects of tempera- 
ture, initial pH, sodium chloride, and sodium pyrophosphate on the growth kinetics 
of Clostridium perfringens. J. Food Prot., 59, 963—968, 1996. 

Kalina, V. Dynamics of microbial growth and metabolic activity and their control by aeration. 
Antonie Van Leeuwenhoek, 63, 353—373, 1993. 

Karl, H. Bestimmung des nitritgehaltes in rucher schen und anden schprodukten. Deutsche 
Lebensmittel-Rundschau, 2, 41^5, 1992. 

Koutsoumanis, K. Predictive modeling of the shelf life of sh under nonisothermal conditions. 
Appl. Environ. Microbiol, 67, 1821-1829, 2001. 

Koutsoumanis, K. and Nychas, G.-J.E. Application of a systematic experimental procedure 
to develop a microbial model for rapid sh shelf life prediction. Int. J. Food Micro- 
biol, 60, 171-184,2000. 

Koutsoumanis, K., Tassou, C.C., Taoukis, PS., and Nychas, G.-J.E. Modelling the effective- 
ness of a natural antimicrobial on Salmonella enteritidis as a function of concentra- 
tion, temperature and pH, using conductance measurements. J. Appl. Microbiol, 84, 
981-987, 1998. 

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

Koutsoumanis, K.P, Taoukis, PS., Drosinos, E.H., and Nychas, G.-J.E. Application of an 
Arrhenius model for the combined effect of temperature and C0 2 packaging on the 
spoilage micro ora of sh. Appl Environ. Microbiol, 66, 3528-3534, 2000. 

Lambert, R.J.W., Skandamis, P.N., Coote, P. J., and Nychas, G.-J.E. A study of the minimum 
inhibitory concentration and mode of action of oregano essential oil, thymol and 
carvacrol. J. Appl Microbiol, 91, 453^162, 2001. 

Lanciotti, R., Sinigaglia, M., Guardini, F., Vannini, L, and Guerzoni, M.E. Growth/no growth 
interface of Bacillus cereus, Staphylococcus aureus and Salmonella enteritidis in 
model systems based on water activity, pH, temperature and ethanol concentration. 
Food Microbiol, 18. 659-668. 2001. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 142 Wednesday, November 12, 2003 12:40 PM 








Legan, D., Vandeven, M., Stewart, C, and Cole, M.B. Modelling the growth, survival and 
death of bacterial pathogens in food. In Foodborne Pathogens: Hazards, Risk Analysis, 
and Control, Blackburn, C. de W. and McClure, P.J. (Eds.), Woodhead Publishing, 
Cambridge, England/CRC Press, Boca Raton, FL, 2002, pp. 53-95. 

Leistner, L. Hurdle technology applied to meat products of the shelf stable and intermediate 
moisture food types. In Properties of Water in Foods in Relation to Quality and 
Stability, Simatos, D. and Multon, J.L., Eds., Martinus Nijhoff Publishers, Dordrecht, 
1985. 

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

Le Marc, Y., Pin, C, and Baranyi, J. Methods to determine the growth domain in a multidi- 
mensional environmental space. In Predictive Modeling in Foods — Conference Pro- 
ceedings ', Van Impe, J.F.M., Geeraerd, A.H., Leguerinel, I., and Mafart, P. (Eds.), 
Katholieke Universiteit Leuven/BioTec, Belgium, 2003, pp. 34-36. 

Leroi, F. and Joffraud, J.J. Salt and smoke simultaneously affect chemical and sensory quality 
of cold-smoked salmon during 5°C storage predicted using factorial design. J. Food 
Prot, 63, 1222-1227, 2000. 

Leroi, E, Joffraud, J.-J., Chevalier, F., and Cardinal, M. Study of the microbial ecology of 
cold-smoked salmon during storage at 8°C. Int. J. Food Microbiol., 39, 111-121, 
1998. 

Leroi, E, Joffraud, J. J., and Chevalier, F. Effect of salt and smoke on the microbiological 
quality of cold-smoked salmon during storage at 5°C as estimated by the factorial 
design method. J. Food Prot., 63, 502-508, 2000. 

Leroy, F. and de Vuyst, L. A combined model to predict the functionality of the bacteriocin- 
producing Lactobacillus sakei strain CTC 494. Appl. Environ. Microbiol, 69, 
1093-1099, 2003. 

Levenspiel, O. Chemical Reaction Engineering, 2nd ed., John Wiley & Sons, New York, 1972. 

Lindroth, S. and Genigeorgis, C. Probability of growth and toxin production by nonproteolytic 
Clostridium botulinum in rock sh stored under modi ed atmospheres. Int. J. Food 
Microbiol, 3, 167-181, 1986. 

Lopez-Malo, A., Guerrero, S., and Alzamora, S.M. Probabilistic modelling of Sac charomyces 
cerevisiae inhibition under the effects of water activity, pH and potassium sorbate 
concentration. J. Food Prot, 63, 91-95, 2000. 

Lou, W. and Nakai, S. Application of arti cial neural networks for predicting the thermal 
inactivation of bacteria: a combined effect of temperature, pH and water activity. 
Food Res. Int., 34, 573-579, 2001. 

Lowenadler, J. and Ronner, U. Determination of dissolved carbon dioxide by colorimetric 
titration in modi ed atmosphere systems. Lett. Appl Microbiol, 18, 285-288, 1994. 

Lund, B.M. Quanti cation of factors affecting the probability of development of pathogenic 
bacteria, in particular Clostridium botulinum, in foods. J. Ind. Microbiol, 12, 144-155, 
1993. 

Lund, B.M., Graham, S.M., and Franklin, J.G. The effect of acid pH on the probability of 
growth of proteolytic strains of Clostridium botulinum. Int. J. Food Microbiol, 4, 
215-226, 1987. 

Masana, M.O. and Baranyi, J. Adding new factors to predictive models: the effect on the risk 
of extrapolation. Food Microbiol. , 17, 367-374, 2000a. 

Masana, M.O. and Baranyi, J. Growth/no growth interface of Brochothrix thermosphacta as 
a function of pH and water activity. Food Microbiol. , 17, 485^193, 2000b. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 143 Wednesday, November 12, 2003 12:40 PM 








Mathlouthi, M. Water content, water activity, water structure and the stability of foodstuffs. 

Food Control 12,409-417,2001. 
McClure, P.J., Roberts, T.A., and Otto Oguru, P. Comparison of the effects of sodium chloride, 

pH and temperature on the growth of Listeria monocytogenes on gradient plates and 

in liquid medium. Lett. Appl. Microbiol, 9, 95-99, 1989. 
McClure, P. J., Baranyi, J., Boogard, E., Kelly, T.M., and Roberts, T.A. A predictive model 

for the combined effect of pH, sodium chloride and storage temperature on the growth 

of Brochotrix thermosphacta. Int. J. Food Microbiol., 19, 161—178, 1993. 
McClure, P.J., Blackburn, C. de W., Cole, M.B., Curtis, PS., Jones, J.E., Legan, ID., Ogden, 

ID., and Peck, M.W. Modelling the growth, survival and death of microorganisms 

in foods: the UK Food Micromodel approach. Int. J. Food Microbiol. , 23, 265-275, 

1994a. 
McClure, P.J., Cole, M.B., and Davies, K.W. An example of the stages in the development 

of a predictive mathematical model for microbial growth: the effect of NaCl, pH and 

temperature on the growth of Aeromonas hydrophila. Int. J. Food Microbiol, 23, 

359-375, 1994b. 
McDonald, K. and Sun, D.W. Predictive food microbiology for the meat industry: a review. 

Int. J. Food Microbiol, 52, 1-27, 1999. 
McGill, A.S., Murray, J., and Hardy, R. Some observations on the determination of phenols 

in smoked sh by solvent extraction: dif culties and improvements in the method- 
ology. Z Lebens. Unters. Forsch., 181, 363-369, 1985. 
McKellar, R.C. and Lu, X.W. A probability of growth model for Escherichia coli 0157:H7 

as a function of temperature, pH, acetic acid, and salt. J. FoodProt, 64, 1922-1928, 

2001. 
McMeekin, T.A., Chandler, R.E., Doe, P.E., Garland, CD., Olley, J., Putro, S., and Ratkowsky, 

D.A. Model for the combined effect of temperature and salt concentration/ water 

activity on growth rate of Staphylococcus xylosus. J. Appl Bacteriol, 62, 543-550, 

1987. 
McMeekin. T.A., Olley, J., Ross, T., and Ratkowsky, D.A. Predictive Microbiology. Theory 

and Application, Research Studies Press, Taunton, U.K., 1993. 
McMeekin, T.A., Presser, K., Ratkowsky, D., Ross, T., Salter, M., and Tienungoon, S. Quan- 
tifying the hurdle concept by modelling the bacterial growth/no growth interface. Int. 

J. Food Microbiol, 55, 93-98, 2000. 
Mellefont, L.A. and Ross, T. The effect of abrupt shifts in temperature on the lag phase 

duration of Escherichia coli and Klebsiella oxytoca. Int. J. Food Microbiol, 83, 

295-305, 2003. 
Mellefont, L.A., McMeekin, T.A., and Ross, T. The effect of abrupt osmotic shifts on the lag 

phase duration of foodborne bacteria. Int. J. Food Microbiol, 83, 281-293, 2003. 
Mellefont, L.A., McMeekin, T.A., and Ross, T. The effect of abrupt osmotic shifts on the lag 

phase duration of physiologically distinct populations of Salmonella typhimurium. 

Int. J. Food Microbiol. , in press. 
Membre, J.-M. and Burlot, P.M. Effect of temperature, pH, and NaCl on growth and pecti- 

nolytic activity of Pseudomonas marginalis. Appl. Environ. Microbiol, 60, 

2017-2022, 1994. 
Membre, J.M. and Kubaczka M. Degradation of pectic compounds during pasteurised vege- 
table juice spoilage by Chryseomonas luteola: a predictive microbiology approach. 

Int. J. Food Microbiol. , 42, 159-166, 1998. 
Membre, J.M., Thurette, J., and Catteau, M. Modelling the growth, survival and death of 

Listeria monocytogenes. J. Appl. Bacteriol, 82, 345—350, 1997. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 144 Wednesday, November 12, 2003 12:40 PM 








Membre, J-M., Ross, T., and McMeekin, T.A. Behaviour of Listeria monocytogenes under 
combined chilling processes. Lett. Appl. Bacteriol, 28, 216-220, 1999. 

Membre, J.M., Kubaczka, M., and Chene, C. Growth rate and growth-no-growth interface 
of Penicillium brevicompactum as functions of pH and preservative acids. Food 
Microbiol, 18, 531-538, 2001. 

Membre, J.-M., Leporq, B., Vilette, M., Mettler, E., Perrier, L., and Zwietering, M. Experi- 
mental protocols and strain variability of cardinal values (pH and a w ) of bacteria 
using Bioscreen C: microbial and statistical aspects. In Microbial Adaptation to 
Changing Environments, Proceedings and Abstracts of the 18th Symposium of the 
International Committee on Food Microbiology and Hygiene (ICFMH), Axelson, L., 
Tronrud, E.S., and Merok, K.J. (Eds.), MATFORSK, Norwegian Food Research 
Institute, As, Norway, 2002, pp. 143-146. 

Messens, W., Verluyten, J., Leroy, F., and De Vurst, L. Modelling growth and bacteriocin 
production by Lactobacillus curvatus LTH 1174 in response to temperature and pH 
values used for European sausage fermentation processes. Int. J. Food Microbiol. , 
81,41-52,2002. 

Miles, D.W., Ross, T., Olley, J., and McMeekin, T.A. Development and evaluation of a 
predictive model for the effect of temperature and water activity on the growth rate 
of Vibrio par ahaemolyticus. Int. J. Food Microbiol. , 38, 133—142, 1997. 

Miller, MB. and Bassler, B.L. Quorum sensing in bacteria. Ann. Rev. Microbiol, 55, 165-199, 
2001. 

Molina, M. and Giannuzzi, L. Combined effect of temperature and propionic acid concen- 
tration on the growth of Aspergillus parasiticus. Food Res. Int., 32, 677-682, 1999. 

Najjar, Y.M., Basheer, LA., and Hajmeer, M.N. Computational neural networks for predictive 
microbiology. 1. Methodology. Int. J. Food Microbiol. , 34, 27^19, 1997. 

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

Ng, T.M., Viard, E., Caipo, M.L., Duffy, S., and Schaffner, D.W. Expansion and validation 
of a predictive model for the growth of Bacillus stearothermophilus in military rations. 
J. FoodSci., 67, 1872-1878, 2002. 

Nichols, D.S., Olley, J., Garda, H., Brenner, R.R., and McMeekin, T.A. Effect of temperature 
and salinity stress on growth and lipid composition of Shewanella gelidimarina. Appl 
Environ. Microbiol, 66, 2422-2429, 2002a. 

Nichols, D.S., Presser, K.A., Olley, J., Ross, T., and McMeekin, T.A. Variation of branched- 
chain fatty acids marks the normal physiological range for growth in Listeria mono- 
cytogenes. Appl Environ. Microbiol, 68, 2809-2813, 2002b. 

Nychas, G.-J.E. and Tassou, C.C. Traditional preservatives: oils and spices. In Encyclopaedia 
of Food Microbiology, Robinson, R.K., Batt, C.A., and Patel, PD. (Eds.), Academic 
Press, San Diego, 2000, pp. 1717-1722. 

O'Brien, G.D. Domestic refrigerator air temperatures and the public's awareness of refriger- 
ator use. Int. J. Environ. Health Res., 1, 41-148, 1997. 

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

Oscar, T.P. Development and validation of a tertiary simulation model for predicting the 
potential growth of Salmonella typhimurium on cooked chicken. Int. J. Food Micro- 
biol, 16, 111 '-190,2002. 

Palumbo, S.A., Williams, A.C., Buchanan, R.L., and Phillips, J.G. Model for the anaerobic 
growth of Aeromonas hydrophila K144. J. Food Prot., 55, 260-265, 1992. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 145 Wednesday, November 12, 2003 12:40 PM 








Palumbo, S.A., Williams, A. C, Buchanan, R.L., Call, J.C., and Phillips, J.G. Expanded model 
for the aerobic growth of Aeromonas hydrophila. J. Food Saf, 16, 1-13, 1996. 

Panagou, E.Z., Skandamis, P.N., and Nychas, G.-J.E. Modelling the combined effect of 
temperature, pH and a w on growth of Monascus ruber, a heat-resistant fungus isolated 
from green table olives. J. Appl. Microbiol, 94, 146-156, 2003. 

Parente, E., Giglio, M.A., Ricciardi, A., and Clementi, F. The combined effect of nisin, 
leucocin F10, pH, NaCl and EDTA on the survival of Listeria monocytogenes in 
broth. Int. J. Food Microbiol, 40, 65-74, 1998. 

Passos, F.V., Fleming, H.P, Felder, R.M., and Ollis, D.F. Modeling growth of Sac charomyces 
rosei in cucumber fermentation. Food Microbiol, 14, 533-542, 1997. 

Passos, F.V., Fleming, H.P, Ollis, D.F., Hassan, H.M., and Felder, R.M. Modeling the specific 
growth rate of Lactobacillus plantarum in cucumber extract. Appl. Microbiol. Bio- 
technol, 40, 143-150, 1993. 

Pecina, R., Bonn, G., and Burtscher, O. High-performance liquid chromatography elution 
behaviour of alcohols, aldehydes, ketones, organic acids and carbohydrates on a 
strong cation-exchange stationary phase. J. Chromatogr., 287, 245-258, 1984. 

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

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

Pitt, R.E. A descriptive model of mold growth and a atoxin formation as affected by envi- 
ronmental conditions. J. FoodProt., 56, 139-146, 1992. 

Pitt, R.E. A model of a atoxin formation in stored products. Trans. ASAE, 38, 1445-1453, 
1995. 

Pouillot, R., Albert, I., Cornu, M., and Denis, J.-B. Estimation of uncertainty and variability 
in bacterial growth using Bayesian inference. Application to Listeria monocytogenes. 
Int. J. Food Microbiol. , 81, 87-104, 2003. 

Presser, K., Ratkowsky, DA., and Ross, T. Modelling the growth rate of Escherichia coli as 
a function of pH and lactic acid concentration. Appl Environ. Microbiol, 63, 
2355-2360, 1997. 

Presser, K.A., Ross, T., and Ratkowsky, DA. Modelling the growth limits (growth/no growth 
interface) of E. coli as a function of pH, lactic acid and temperature. Appl Environ. 
Microbiol, 64, 1773-1779, 1998. 

Presser, K.A., Salter, M.A., Ratkowsky, D.A., and Ross, T. (1999). Development of growth 
limits (growth/no growth interface) modelling and its application to predictive food 
microbiology. Recent Res. Dev. Microbiol, 3, 535-549, 1999. 

Rasch, M. The in uence of temperature, salt and pH on the inhibitory effect of reuterin on 
Escherichia coli. Int. J. Food Microbiol, 72, 225—231, 2002. 

Rasmussen, S.K.J., Ross, T., Olley, J., and McMeekin, T. A process risk model for the shelf- 
life of Atlantic salmon llets. Int. J. Food Microbiol, 73, 47-60, 2002. 

Ratkowsky, DA. Nonlinear Regression Modeling: A Uni ed Practical Approach, Marcel 
Dekker, New York, 1983, 276 pp. 

Ratkowsky, DA. Principles of nonlinear- regression modeling. J. Ind. Microbiol, 12, 195-199, 
1993. 

Ratkowsky, DA. Some examples of, and problems with, the use of nonlinear logistic regres- 
sion in predictive food microbiology. Int. J. Food Microbiol, 73, 119-125, 2002. 

Ratkowsky, DA. and Ross, T. Modelling the bacterial growth/no growth interface. Lett. Appl 
Microbiol, 20, 29-33, 1995. 

Ratkowsky, DA., Olley, J., McMeekin, T.A., and Ball, A. Relation between temperature and 
growth rate of bacterial cultures. J. Bacteriol, 149, 1-5, 1982. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 146 Wednesday, November 12, 2003 12:40 PM 








Ratkowsky, D.A., Lowry, R.K., McMeekin, T.A., Stokes, A.N., and Chandler, R.E. Model 

for bacterial culture growth rate throughout the entire biokinetic temperature range. 

J. Bacterid., 154, 1222-1226, 1983. 
Ratkowsky, D.A., Ross, T., McMeekin, T.A., and Olley, J. Comparison of Arrhenius-type and 

Belehradek-type models for prediction of bacterial growth in foods. J. Appl. Bacte- 

riol, 71,452-459, 1991. 
Razavilar, V. and Genigeorgis, C. Prediction of Listeria spp. growth as affected by various 

levels of chemicals, pH, temperatures and storage time in a model broth. Int. J. Food 

Microbiol, 40, 149-157, 1998. 
Resnik, S.L. and Chirife, J. Proposed theoretical water activity values at various temperatures 

for selected solutions to be used as reference sources in the range of microbial growth. 

J. FoodProt, 51, 419-423, 1988. 
Riemann, H. The effect of the number of spores on growth and toxin formation by C. 

botulinum type E in inhibitory environments. In Botulism, 1966. Proceedings of the 

5th International Symposium on Food Microbiology, Moscow, July 1966, Ingram, 

M. and Roberts, T.A., Eds., Chapman and Hall, London, 1967, 150-169. 
Roberts, T.A. and Jarvis, B. Predictive modelling of food safety with particular reference to 

Clostridium botulinum in model cured meat systems. In Food Microbiology: Advances 

and Prospects, Roberts, T.A. and Skinner, F.A. (Eds.), Academic Press, New York, 

1983, pp. 85-95. 
Robinson, T.P, Ocio, M.J., Kaloti, A., and Mackey, B.M. The effect of growth environment 

on the lag phase of Listeria monocytogenes. Int. J. Food Microbiol . , 44, 83-92, 1998. 
Rodriguez, A.M.C., Alcala, E.B., Gimeno, R.M.G., and Cosano, G.Z. Growth modelling of 

Listeria monocytogenes in packed fresh green asparagus. Food Microbiol, 17, 

421^127,2000. 
Ross, T. A Philosophy for the Development of Kinetic Models in Predictive Microbiology, 

Ph.D. thesis, University of Tasmania, Hobart, Australia, 1993a. 
Ross, T. Belehradek-type models. J. Ind. Microbiol, 12, 180-189, 1993b. 
Ross, T. Predictive Food Microbiology Models in the Meat Industry, Meat and Livestock 

Australia, Sydney, 1999a, 196 pp. 
Ross, T. Assessment of a theoretical model for the effects of temperature on bacterial growth 

rate. Refrig Sci. Technol Proc, EUR 18816, 64-71, 1999b. 
Ross, T. and McMeekin, T.A. Predictive microbiology: a review. Int. J. Food Microbiol, 23, 

241-264, 1994. 
Ross, T. and Shadbolt, C.T. Predicting Escherichia coli Inactivation in Uncooked Comminuted 

Fermented Meat Products, Meat and Livestock Australia, Sydney, 2001, 66 pp. 
Ross, T., Fazil, A., and Paoli, G. Listeria monocytogenes in Australian processed meat 

products: risks and their management. Meat and Livestock Australia, Sydney, in press. 
Ross, T., Baranyi, J., and McMeekin, T.A. Predictive microbiology and food safely. In 

Encyclopaedia of Food Microbiology, Robinson, R., Batt, C.A., and Patel, P. (Eds.), 

Academic Press, London. 2000a, pp. 1699-1710. 
Ross, T., Dalgaard, P., and Tienungoon, S. Predictive modelling of the growth and survival 

of Listeria in shery products. Int. J. Food Microbiol . , 62, 231-245, 2000b. 
Ross, T., Ratkowsky, D.A., Mellefont, L.A., and McMeekin, T.A. Modelling the effects of 

temperature, water activity, pH and lactic acid concentration on the growth rate of 

Escherichia coli. Int. J. Food Microbiol, 82, 33^3, 2003. 

— - J - 

Rosso, L. Modelisation et Microbiologic Previssionnelle: Elaboration d'un nouvel outil pour 

VAgro-alimentaire, Universite Claude Bernard, Lyon, 1995. 
Rosso, L. Models using cardinal values. Refrig. Sci. Technol Proc, EUR 18816, 48-55, 1999. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 147 Wednesday, November 12, 2003 12:40 PM 








Rosso, L., Lobry, R., and Flandrois, J.R An unexpected correlation between cardinal temperatures 
of microbial growth highlighted by a new model. J. Theor. Biol, 162, AA1-463, 1993. 

Rosso, L., Lobry, J.R., Bajard, S., and Flandrois, J.R Convenient model to describe the 
combined effects of temperature and pH on microbial growth. Appl. Environ. Micro- 
biol, 61, 610-616, 1995. 

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

Salter, M., Ross, T., and McMeekin, T.A. Applicability of a model for non-pathogenic 
Escherichia coli for predicting the growth of pathogenic Escherichia coli. J. Appl. 
Microbiol, 85, 357-384, 1998. 

Salter, M.A., Ratkowsky, D.A., Ross, T., and McMeekin, T.A. Modelling the combined 
temperature and salt (NaCl) limits for growth of a pathogenic Escherichia coli strain 
using generalised nonlinear regression. Int. J. Food Microbiol, 61, 159-167, 2000. 

Sautour, M., Dantigny, P., Divies, C, and Bensoussan, M. A temperature- type model for 
describing the relationship between fungal growth and water activity. Int. J. Food 
Microbiol, 67, 63-69, 2001. 

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

Schepers, A.W., Thibault, J., and Lacroix, C. Comparison of simple neural networks and 
nonlinear regression models for descriptive modelling of Lactobacillus helveticus 
growth in pH-controlled batch cultures. Enzyme Microb. Technol, 26, 431^145, 2000. 

Schoolfield, R.M., Sharpe, P.J.H., and Magnuson, C.E. Non-linear regreaon of biological 
temperature-dependent rate models based on absolute reaction-rate theory. J. Theor. 
Biol, 88,719-731, 1981. 

Seman, D.L., Borger, A.C., Meyer, ID., Hall, PA., and Milkowski, A.L. Modeling the growth 
of Listeria monocytogenes in cured ready- to-eat processed meat products by manip- 
ulation of sodium chloride, sodium diacetate, potassium lactate, and product moisture 
content. J. FoodProt, 65, 651-658, 2002. 

Sergelidis, D., Abrahim, A., Sarimvei, A., Panoulis, C, Karaioannoglou, P., and Genigeorgis, 
C. Temperature distribution and prevalence of Listeria spp. in domestic, retail, and 
industrial refrigerators in Greece. Int. J. Food Microbiol, 34, 171-177, 1997. 

Sharpe, P.J.H. and DeMichele, D.W. Reaction kinetics and poikilotherm development. 
J. Theor. Biol, 64, 649-670, 1977. 

Shewan, J.M. The biological stability of smoked and salted sh. Chem. Ind., July, 501-505, 
1949. 

Shoemaker, S.P and Pierson, M.D. "Phoenix phenomenon" in the growth of Clostridium 
perfringens. Appl Environ. Microbiol. , 32, 803—806, 1976. 

Simpson, R., Almonacid, S., and Acevedo, C. Development of a mathematical model for 
MAP systems applied to nonrespiring foods. J. Food Sci., 66, 561-567, 2001a. 

Simpson, R., Almonacid, S., and Acevedo, C. Mass transfer in Paci c hake (Merluccius 
australis) packed in refrigerated modi ed atmosphere. J. Food Process. Eng, 24, 
405^121, 2001b. 

Sinigaglia, M., Corbo, M.R., Altieri, C, and Massa, S. Response surface model for effects 
of temperature, water activity and pH on germination of Bacillus cereus spores. 
J. FoodProt, 22, 121-133, 2002. 

Skandamis, P.N. and Nychas, G.- J.E. Development and evaluation of a model predicting the 
survival of Escherichia coli 0157:H7 NCTC 12900 in homemade eggplant at various 
temperatures, pHs, and oregano essential oil concentrations. Appl Environ. Micro- 
biol, 66, 1646-1653.2000. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C03.fm Page 148 Wednesday, November 12, 2003 12:40 PM 








Skandamis, P.N., Davies, K.W., McClure, P.J., Koutsoumanis, K., and Tassou, C. A vialistic 
approach for non-thermal inactivation of pathogens in traditional Greek salads. Food 
Microbiol, 19,405-421,2002. 

Skinner, G.E., Lerkin, J.W., and Rhodehamel, E.J. Mathematical modeling of microbial 
growth: a review. J. Food Saf, 14, 175-217, 1994. 

Smith, M.G. The generation time, lag time and minimum temperature of growth of coliform 
organisms on meat, and the implications for codes of practice in abattoirs. J. Hyg. 
(Camb.), 94, 289-300, 1985. 

Sofos, J.N., Beuchat, L.R., Davidson, P.M., and Johnson, E.A. Naturally Occurring Antimi- 
crobials in Food, Task Force Report 132, Council of Agricultural Science and Tech- 
nology, Ames, LA, 1998. 

Stewart, CM., Cole, M.B., Legan, J.D., Slade, L., Vandeven, M.H., and Schaffner, D.W. 
Modeling the growth boundary of Staphylococcus aureus for risk assessment pur- 
poses. J. FoodProt., 64, 51-57, 2001. 

Stewart, CM., Cole, M.B., Legan, J.D., Slade, L., Vandeven, M.H., and Schaffner, D.W. Sta- 
phylococcus aureus growth boundaries: moving towards mechanistic predictive models 
based on solute specific effects. Appl. Environ. Microbiol, 68,1864-1871, 2002. 

Sunen, E. Minimum inhibitory concentration of smoke wood extracts against spoilage and 
pathogenic micro-organisms associated with foods. Lett. Appl. Microbiol, 27, 45^18, 
1998. 

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

Sutherland, J.P, Aherne, A., and Beaumont, A. L. 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, 1996. 

Sutherland, J.P, Bayliss, A. J., and Braxton, D.S. Predictive modelling of growth of Escher- 
ichia coli 0157:H7: the effects of temperature, pH and sodium chloride. Int. J. Food 
Microbiol, 25, 29^9, 1995. 

Sutherland, J.P, Bayliss, A.J., Braxton, D.S., and Beaumont, A.L. 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, 1997. 

Thurette, J., Membre, J.M., Han Ching, L., Tailliez, R., and Catteau, M. Behavior of Listeria 
spp. in smoked sh products affected by liquid smoke, NaCl concentration, and 
temperature. J. FoodProt., 61, 1475-1479, 1998. 

Tienungoon, S. Some Aspects of the Ecology of Listeria monocytogenes in Salmonoid Aquac- 
ulture, Ph.D. thesis. University of Tasmania, Hobart, Australia, 1988. 

Tienungoon, S., Ratkowsky, D.A., McMeekin, T.A., and Ross, T. Growth limits of Listeria 
monocytogenes as a function of temperature, pH, NaCl and lactic acid. Appl Environ. 
Microbiol, 66, 4979^1987, 2000. 

Toth, L. and Potthast, K. Chemical aspects of the smoking of meat and meat products. Adv. 
Food Res., 29, 87-158, 1984. 

Tu, J.V. Advantages and disadvantages of using arti cial neural networks versus logistic 
regression for predicting medical outcomes. J. Clin. Epidemiol, 11, 1225-1231, 1996. 

Tucker, I.W. Estimation of phenols in meat and fat. Assoc. Off. Agric. Chem., 25, 779-782, 
1942. 

Tuynenberg Muys, G. Microbial safety in emulsions. Process Biochem., 6, 25-28, 1971. 

Uljas, H.E., Schaffner, D.W., Duffy, S., Zhao, L., and Ingham, S.C Modeling of combined 
steps for reducing Escherichia coli 0157:H7 populations in apple cider. Appl Envi- 
ron. Microbiol, 67, 133-141, 2001. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 149 Wednesday, November 12, 2003 12:40 PM 








Valik, L., Baranyi, J., and Gorner, F. Predicting fungal growth: the effect of water activity on 
Penicillium rocqueforti. Int. J. Food Microbiol, 47, 141—146, 1999. 

Wei, Q.K., Fang, T.J., and Chan, W.C. Development and validation of growth model for 
Yersinia enterocolitica in cooked chicken meats packed under various atmosphere 
packaging and stored at different temperatures. J. Food Prot., 64, 987-993, 2001. 

Whiting, R.C. Microbial modeling in foods. Crit. Rev. Food Sci. Nutr., 35, 461-494, 1995. 

Whiting, R.C. and Bagi, L.K. Modeling the lag phase of Listeria monocytogenes. Int. J. Food 
Microbiol., 73, 291-295, 2002. 

Whiting, R.C. and Buchanan, R.L. Letter to the editor: a classi cation of models in predictive 
microbiology. A reply to K.R. Davey. Food Microbiol., 10, 175-177, 1993. 

Whiting, R.C. and Call, J.E. Time of growth model for proteolytic Clostridium botulinum. 
Food Microbiol, 10,295-301, 1993. 

Whiting, R.C. and Cygnarowic-Provost, M. A quantitative model for bacterial growth and 
decline. Food Microbiol. , 9, 269-277, 1992. 

Whiting, R.C. and Golden, M.H. Modeling temperature, pH, NaCl, nitrite and lactate on the 
survival of Escherichia coli 0157:H7 in broth. J. Food Safety, 23, 61-74, 2003. 

Whiting, R.C. and Oriente, J.C. Time-to-turbidity model for non-proteolytic type B Clostrid- 
ium botulinum. Int. J. Food Microbiol, 36, 49-60, 1997. 

Whiting, R.C. and Strobaugh, T.P Expansion of the time-to-turbidity model for proteolytic 
Clostridium botulinum to include spore numbers. Food Microbiol, 15,449^153, 1998. 

WHO/FAO (World Health Organisation and Food and Agriculture Organisation). Risk Assess- 
ment of Listeria monocytogenes in ready- to-eat foods. Food and Agriculture Organ- 
isation, Rome, in press. 

Wijtzes, T., De Wit, J.C, Huis in't Veld, J.H., van't Riet, K., and Zwietering, M.H. Modelling 
bacterial growth of Lactobacillus curvatus as a function of acidity and temperature. 
Appl. Environ. Microbiol, 61, 2533-2539, 1995. 

Wijtzes, T., van't Riet, K., Huis in't Veld, J.H., and Zwietering, M.H. A decision support 
system for the prediction of microbial food safely and food quality. Int. J. Food 
Microbiol, 42, 79-90, 1998. 

Wijtzes, T., Rombouts, F.M., Kant-Muermans, M.L., van't Riet, K., and Zwietering, M. Devel- 
opment and validation of a combined temperature, water activity, pH model for bac- 
terial growth rate of Lactobacillus curvatus. Int. J. Food Microbiol, 63, 57-64, 2001. 

Winans, S.C. and Bassler, B.L. Mob psychology. J. Bacteriol, 184, 873-888, 2000. 

Yan, W. and Hunt, L.A. An equation for modelling the temperature response of plants using 
only the cardinal temperatures. Ann. Bot., 84, 607-614, 1999. 

Yin, W. and Wallace, D.H. A nonlinear model for crop development as a function of temper- 
ature. Agric. Forest Mete or ol, 11, 1-16, 1995. 

Zaika, L.L., Moulden, E., Weimer, L., Phillips, J.G., and Buchanan, R.L. Model for the 
combined effects of temperature, initial pH, sodium chloride and sodium nitrite 
concentrations on anaerobic growth of Shigella exneri. Int. J. Food Microbiol. , 23, 
345-358, 1994. 

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

Zhao, Y, Wells, J.H., and McMillin, K.W. Dynamic changes of headspace gases in C0 2 and 
N 2 packed fresh beef. J. Food Sci., 60, 571-575, 591, 1995. 

Zhao, L., Chen, Y, and Schaffner, D.W. Comparison of logistic regression and linear regres- 
sion on modelling percentage data. Appl Environ. Microbiol, 67, 2129-2135, 2001. 

Zwietering, M. Microbial and engineering characteristic numbers for process evaluation. 
Refrig Sci. Technol Proc, EUR 18816, 272-279, 1999. 

2004 by Robin C. McKellar and Xuewen Lu 













1237_C03.fm Page 150 Wednesday, November 12, 2003 12:40 PM 







Zwietering, M.H., Cuppers, H.G.A.M., De Wit, J.C., and van't Riet, K. Evaluation of data 
transformations and validation of models for the effect of temperature on bacterial 
growth. Appl. Environ. Microbiol, 60, 195-203, 1994. 

Zwietering, M.H., de Koos, J.T., Hasenack, B.E., De Wit, J.C., and van't Riet, K. Modeling 
of bacterial growth as a function of temperature. Appl. Environ. Microbiol, 57, 
1094-1101, 1991. 

Zwietering, M.H., De Wit, J.C., and Notermans, S. Application of predictive microbiology 
to estimate the number of Bacillus cereus in pasteurised milk at the time of consump- 
tion. Int. J. Food Microbiol, 30, 55-70, 1996. 

Zwietering, M.H., Jongenburger, I., Rombouts, F.M., and van't Reit, K. Modeling of the 
bacterial growth curve. Appl Environ. Microbiol, 56, 1875-1881, 1990. 

Zwietering, M.H., Wijtzes, T., De Wit, J.C., and van't Riet, K. A decision support system for 
prediction of the microbial spoilage in foods. J. Food Prot., 55, 973-979, 1992. 





2004 by Robin C. McKellar and Xuewen Lu