Skip to main content

Full text of "كتب في علم الميكروبيولوجي"

See other formats




~V 






1237_C09.fm Page 285 Wednesday, November 12, 2003 1:08 PM 






Q Modeling the History 
Effect on Microbial 
Growth and Survival: 
Deterministic and 
Stochastic Approaches 



Jozsef Baranyi and Carmen Pin 



CONTENTS 

9.1 Introduction 

9.2 Modeling the History Effect at Population Level 
(Deterministic Modeling) 

9.2.1 Traditional Parameters 

9.2.2 Replacing the Lag Parameter with a Physiological 
State Parameter 

9.2.3 A Rescaling of the Physiological Parameter Provides 
a New Interpretation 

9.3 Modeling the History Effect at Single-Cell Level 
(Stochastic Modeling) 

9.3.1 Nature of Stochastic Modeling 

9.3.2 Physiological State and Lag Parameters for Single Cells 

9.3.3 A Simulation Program 

9.3.4 Lag/Survival Distribution of Single Cells and Lag/Shoulder 
of Population Curves 

9.3.5 History Effect on the Shoulder Periods of Survival Curves 

9.4 Concluding Remarks 
Acknowledgment 
References 




2004 by Robin C. McKellar and Xuewen Lu 










1237_C09.fm Page 286 Wednesday, November 12, 2003 1:08 PM 




~V 









9.1 INTRODUCTION 

The objective of food safety microbiology is to eliminate or at least significantly 
reduce the concentration of pathogenic microbes in food and to prevent their resus- 
citation and multiplication. Predictive microbiology focuses on the quantitative 
methods used to achieve this goal. 

In a mathematical sense, a predictive model is a mapping between the environ- 
ment of the microbial cells and their response. While primary models describe how 
the microbial cell concentration changes with time in a constant environment (see 
Chapter 2), a secondary model (such as D values vs. temperature; see Chapter 3) 
expresses a typical parameter of the response as a function of the environmental 
factors. Both the environment and the response are characterized by one or more, 
possibly dynamic, variables (i.e., they may change with time during observation). 
Temperature changing with time is an example of a dynamic environment (see 
Chapter 7); growth/survival curves are examples of dynamic responses, either in 
constant or dynamic environments. 

In accordance with the standard approach to modeling, microbial responses are 
studied from a distinguished time point — the zero time of observation. Naturally, 
this zero time is when the environment of the cells undergoes a sudden change. 
Researchers should know, or at least implicitly assume the conditions at this zero 
time to provide initial values for modeling. Such initial values are, for example, the 
initial composition of natural flora (in the case of food environment) or the inoculum 
level and initial physiological state of the studied cells. Respectively, we can speak 
about actual (current) and previous (historical or preinoculum) environments — the 
initial values are the result of the previous environment. 

A major obstacle in the development of predictive microbiology is that it is 
difficult to acquire data at low levels of cell concentration. Currently, commonly 
available counting methods can only measure cell concentrations of higher than 
approximately 10 cells per gram of sample. Therefore, it is frequently problematic 
to validate predictive microbiology models since most of them focus on situations 
when the cell concentration is relatively low and the measurements are inaccurate. 

Growth modeling concentrates on the early stages of microbial growth in the 
actual environment. As mentioned, the previous environment affects growth in the 
actual environment, which effect gradually diminishes after inoculation. This is called 
the adjustment process; the time during which it happens is called the lag period. 

Death modeling deals with decay in microbial concentration having two main 
causes: thermal inactivation, when death is relatively fast; and nonthermal death, 
which is generally a much slower process. Although the physiological mechanisms 
of the two can be completely different, it is possible to construct mathematical 
models for them in an analogous way. What is more, to some extent, growth and 
death modeling can also be discussed analogously. Accordingly, we show some basic 
techniques to model the effect of history on growth and death in a parallel way. 

We also present a comparative review of the applied deterministic and stochastic 
models. Deterministic models can always be treated as sort of "averaged out" 
versions of stochastic models. However, "smoothing down" the biological variability 
means a certain loss of information — the question is just how much. Biology itself 

2004 by Robin C. McKellar and Xuewen Lu 












~V 






1237_C09.fm Page 287 Wednesday, November 12, 2003 1:08 PM 






always has stochastic elements, but sometimes they are small enough (or we hope 
they are small enough) to be negligible. In this case, the applied model (and its one 
or more variables and processes) is reduced to be deterministic, which is simpler 
and easier to handle. However, such a reduction can be misleading when the random 
elements are significant. For example, when modeling the lag period of a population 
of relatively few cells (the "dormant" phase of the "log cone. vs. time curve" before 
exponential growth), the random effects originating from the variability among the 
cells cannot be ignored. Similarly, when modeling the number of survivors in an 
adverse environment, the stochastic variability of the resistance of individual cells 
can also become significant. 

Many authors have studied the effect of history on the lag time before growth 
(Augustin et al., 2000a; Augustin and Carlier, 2000; Beumer et al., 1996; Breand et 
al., 1999; Buchanan and Klavitter, 1991; Dufrenne et al., 1997; Gay et al., 1996; 
Hudson, 1993; Mackey and Kerridge, 1988; Membre et al., 1999; Stephens et al., 
1997; Walker et al., 1990; Wang and Shelef, 1992) and on the shoulder period before 
exponential death (Cotterill and Glauert, 1969; Manas et al., 2001; Ng et al., 1962; 
Sherman and Albus, 1923; Strange and Shon, 1964). In this chapter, we attempt to 
provide a simple modeling approach to take into account the history effect via the 
concept of initial work/damage done before the exponential growth/death phase. 
First, we apply a deterministic model at population level (i.e., when the population 
size is represented by a single continuous variable), after which we discuss similar 
questions at single-cell level, using stochastic processes, when the variability among 
individual cells is also taken into account. 

9.2 MODELING THE HISTORY EFFECT AT 
POPULATION LEVEL (DETERMINISTIC 
MODELING) 

9.2.1 Traditional Rvrameters 

Throughout this chapter, x(t) denotes the cell concentration at the time t, and y(t) 
denotes its natural logarithm. The slope of the "y(f) vs. time" curve is the (instan- 
taneous) specific rate, denoted by \l(t). Its maximum value, |i max , is the maximum 
specific rate. It is positive for growth and negative for death (Figure 9.1a and Figure 
9.1b). The maximum specific rate is measured at the inflexion point of the curve, 
if it has a sigmoid shape (Figure 9.1a). If the model is biphasic (Figure 9.1b), then 
this maximum does not necessarily exist. In this case, (J max is meant in an asymptotic 
sense: \l(t) — > (J max as t — > °°. In fact, we define the biphasic class of models by 
the criterion that \i(t) converges monotonically from a small, |i.(0) initial rate, to a 
finite value. 

In the case of growth, the most frequently modeled parameter is the maximum 
specific growth rate, or one of its rescaled versions: the r max maximum rate in terms 
of log 10 cell concentration, or the doubling time, T d . Note the relation between them: 

r„« = lWln 10; T d = In 2/^ max (9.1) 




2004 by Robin C. McKellar and Xuewen Lu 










1237_C09.fm Page 288 Wednesday, November 12, 2003 1:08 PM 







a) 



/ ,m 




t 






FIG U RE 9.1 (a) Main parameters of a sigmoid growth curve. The maximum specific growth 
rate |i. max can be measured at the inflexion point of the curve, (b) Main parameters of a biphasic 
survival curve. The instantaneous specific death rate converges to a limit value |i r 



'max. 



In the case of survival (or inactivation), instead of the analogous "halving time," 
it is more common to use the time to a decimal reduction, called D value: 



D = -In 10/fJ 



max 



(9.2) 



The reason for this slight inconsistency is historical: the first predictive models were 
created to describe thermal inactivation processes and, at high temperatures, the 
decimal reduction times were more practical to use than halving times. 

Some authors use the maximum specific rate and some the rate of the log 10 x(f) 
function, which can be obtained by an ln(10) ~ 2.3 conversion factor from [{(t). Most 
frequently, cell concentration data are given in terms of log 10 of the cell number per 
volume. However, the maximum specific rate has a more universal theoretical mean- 
ing as shown by the formulae below: 



dy d(\n x) dx I x 



dt 



dt 



dt 



m 



(9.3) 



2004 by Robin C. McKellar and Xuewen Lu 













~V 






1237_C09.fm Page 289 Wednesday, November 12, 2003 1:08 PM 






that is, 



— « ll At (9.4) 



max 
X 



The left-hand side of the latter equation denotes a relative increase or decrease of 
the population in At time. Therefore, the following interpretation can be given for 
the maximum specific rate: the probability that, at the time t, a single cell divides 
(growth) or dies (death), during the small [t, t + At] interval, is about \x(t)At; i.e., 
proportional to the length of At, with the factor \x(t). This is a fundamental link 
between the specific growth rate measured at population level and the probability 
of division at single-cell level. 

Another important, frequently modeled parameter is the lag time (for growth), 
or shoulder period (for survival), which is denoted by X throughout this chapter. For 
sigmoid curves (Figure 9.1a), it is traditionally defined as the time when the tangent 
of the y(t) curve with the maximum slope crosses the initial level, y = In x Q (Pirt, 
1975). (It is irrelevant whether the natural logarithm or log 10 is used to define the 
lag or shoulder.) For biphasic curves (Figure 9.1b), to find a good "geometrical" 
definition is not so straightforward as pointed out below. 

The definition of lag/shoulder does not have the strong basis that would connect 
the parameter to a single-cell equivalent, as in the case of the specific rate. Baranyi 
(1998) discussed the mathematical relation between the lag of the population (as 
defined above) and the lag times of the participating individual cells (see later in 
this chapter). He showed that the population lag depends on the subsequent specific 
growth rate and the initial cell number, even if the individual lag times do not. 
Another anomaly can be shown by means of biphasic models. It is natural to expect 
that, if a series of tangents drawn to the y(t) curves converges to a limit value (see 
Figure 9.1b), then the intersections with the initial v level also converge, in which 
case the lag could be defined by the limit value. Baranyi and Pin (2001) showed a 
biphasic death model where this was not true; the intersection, denoting the end of 
the shoulder period, would not converge, as t — > °°, although the specific death rate 
did converge. 

9.2.2 Replacing the Lag Rrameter with a Rjysiological 

Sate Parameter 

For growth, the "waiting time" of the individual cells is due to an adjustment process 
before the exponential growth. For survival curves, the shoulder period is due to an 
initial resistance before exponential death (such as in the case of the multitarget 
theory, Hermann and Horst, 1970). As shown by Baranyi (2002), it is not easy to 
relate the distribution of the "waiting times" of individual cells to the apparent 
lag/shoulder parameter of the population, measured in the traditional way. 

Baranyi and Roberts (1994) suggested that the lag should be considered as a 
derived parameter, a consequence of both the actual environment (via the specific 
growth rate), and the history of the cells (via a newly introduced parameter, the 
initial physiological state of the cells). Their method can be outlined as follows. 

2004 by Robin C. McKellar and Xuewen Lu 












~V 



1237_C09.fm Page 290 Wednesday, November 12, 2003 1:08 PM 






c 

3 

o 
o 




time 




FIGU RE 9.2 Interpretation of the oc physiological state. h = -ln(a ) = A,|i max is the "work 
to be done," which is the same for different temperatures. h is an initial value, as is \nx . 

Just as y is an initial value, we can introduce another value in order to quantify 
the "suitability" of the population to the current environment (i.e., the history effect). 
Let this quantity be denoted by oc (which comes from the term adjustment function, 
in the original paper). Let it be a dimensionless number between and 1. For growth, 
consider it as the fraction of the initial cells, which, without a lag, would be able to 
produce the same final growth as the original x initial cells with a lag. (Although 
the real situation is not like this, it is easier to think about a this way. In reality, 
a is the mean value of the respective parameters of the individual cells. The point 
is, as we will see later, that it does not matter if we think of it as "oc , the fraction 
of the cells that are able to grow," or "each cell is oc -suitable") 

By simple geometry (see Figure 9.2), it can be shown that 




X = 



-ln(cc n ) 



M 



(9.5) 



max 



This formula expresses our expectation that the lag time depends on both the actual 
environment, represented by |i max , and the history of the cells, characterized by oc . 
The history effect appears as an initial value, and this will be called the initial 
physiological state. Note that it does not represent simply the "health" of the cells, 
rather their suitability to the actual environment. Its extreme values are a = 0, in 
which case the lag is infinitely long, and oc = 1, when all the cells enter the 
exponential phase immediately and the lag time is 0. The effect of the initial 
physiological state on the lag is shown in Figure 9.3. 

In terms of biological interpretability, the main advantages of using oc , instead 
of lag, are: 

1. It is an initial value, affecting the lag, and this is in accord with our 
mechanistic thinking. 





2004 by Robin C. McKellar and Xuewen Lu 











~V 



1237_C09.fm Page 291 Wednesday, November 12, 2003 1:08 PM 







time 




FIG U RE 9.3 Depending on the cc physiological state, the lag can change while the maximum 
specific growth rate remains the same. 

2. It has a very strong connection with similar parameters of the individual 
cells: the physiological state of the population is equal to the average of 
the physiological states of the initial cells, irrespective of their number 
(Baranyi, 1998). 

3. It can be interpreted, via the individual cells, even if no traditional lag 
can be measured on the bacterial curve (for example, when it does not 
reach its inflexion point). 

9.2.3 A Rescaling of the Rmio logical Parameter Riovides 
a New Interpretatio n 

An interpretation of the history effect is also possible by means of the inverse of 
the physiological state. Introduce h Q = ln(l/a ) = -In a . It can be conceived as an 
initial hurdle, or the "work to be done" during the lag (Robinson et al., 1998; see 
Figure 9.2). As we see, it is the product of the lag and the maximum specific growth 
rate. Observe that, using the doubling time, T d = In 2/|i. max relation, the following 
formula can be obtained: 




h Q =ln2 



T, 



(9.6) 



The quantity X/T d is often called "relative lag" (McMeekin et al., 2002) and it is 
essentially the same as h . Some authors have reported that it is unaffected by small 
changes of temperature (Pin et al., 2002; Robinson et al., 1998) and pH (McKellar 
et al., 2002a). This can be explained as follows: the more suitable the cells are to 
the new environment, the less work they have to carry out to adapt. Therefore, the 
lag is longer at lower temperatures because the "work to be done," after the inocu- 
lation, is carried out more slowly, although the amount of work is the same (Figure 
9.4). Authors studying the effect of history and actual growth conditions on the work 
to be done during the lag include Delignette-Muller (1998), Robinson et al. (1998), 
Pin et al. (2002), and Augustin et al. (2000a). 





2004 by Robin C. McKellar and Xuewen Lu 










1237_C09.fm Page 292 Wednesday, November 12, 2003 1:08 PM 







lnx 



ln(a 




time 



FIGURE 9.4 As the temperature decreases, the work to be done (the difference between 
In x and ln(oc x )) is the same, while the maximum specific growth rate (i max becomes 
smaller and the lag becomes longer. The work to be done during the lag is also carried out 
at slower rates. 




The parameter h = -In oc , the "work to be done" before the exponential phase, 
plays the role of a bridge between the history and the actual environment. With an 
analogy, imagine that cells are crossing this bridge. If its angle is horizontal (h Q = 
0; no difficulty to overcome), then there is no lag. If it is vertical (h = <*>), then the 
work to be done is infinitely big, the lag is infinitely long. On the other hand, if the 
temperature is higher then every movement is quicker and so is the crossing (lag), 
although the work to be done is the same (Figure 9.4). It can be expected that this 
work is smaller when adjusting to a more favorable environment than the opposite 
way (Delignette-Muller, 1998; Robinson et al., 1998). This theory explains why the 
environmental effect on the lag has always been much less accurately predicted 
(Dufrenne et al., 1997; McKellar et al., 2002b; McMeekin et al., 2002; Robinson et 
al., 1998; Whiting and Bagi, 2002) than the specific growth rate. The lag also depends 
on the history, the details of which often have not even been recorded, let alone 
taken into account in the model. 




9.3 MODELING THE HISTORY EFFECT AT SINGLE-CELL 
LEVEL (STOCHASTIC MODELING) 

9.3.1 Nature of Stochastic Modeling 

Finer details of the effect of history on growth/death can be seen through a closer 
look at the distribution of the division/survival times of individual cells. As men- 
tioned in Section 9.1, if only deterministic models are used to describe responses 
of microbial populations then the variability among single cells is ignored. This is 
a problem because predicting microbial variability is increasingly important for 
Quantitative Microbial Risk Assessment. This variability of the responses increases 
under stress conditions, which makes predictions even less accurate. 

A characteristic feature of stochastic modeling is that it sometimes provides 
unexpected results. An example of this is the relation between the doubling time of 
a growing population and the average generation time of the individual cells in the 

2004 by Robin C. McKellar and Xuewen Lu 













~V 






1237_C09.fm Page 293 Wednesday, November 12, 2003 1:08 PM 






population. Depending on the distribution of the individual generation times, these 
two quantities can be markedly different. If the generation times follow the classical 
exponential distribution, then their average is l/|i max , while the doubling time is 
known to be In 2/|i. max , where |i. max is the (constant) maximum specific growth rate 
of the population. Baranyi and Pin (2001) described the exponential growth as a 
Poisson birth process and showed that with this approach the doubling time depends 
on the number of cells, from l/|i- max , when the number of cells is 1, converging to 
In 2/|l max , when the number of cells increases. 

9.3.2 Physiological Sate and Lag Parameters for 
Sngle Cells 

Let N(t) be the number of cells in a defined space, at the time t (remember, that x{f) 
denoted the concentration of the cells). Let the lag for the ith cell of the initial 
population be denoted by x, (/ = 1,2,..., N). We assume that they are identically 
distributed independent random variables, and their mean value is x. Note that x, is 
less than the time to the first division of the /th cell, which also includes the first 
generation time. 

Define the physiological states a, (i = 1,2,... JV) for the individual cells by a,= 
exp (-|i raax X,), analogously to the deterministic theory. A basic theorem of Baranyi 
(1998) is that the average of the individual physiological states is the same as the 
physiological state of the population, irrespective of the distribution of the individual 
lag times. Moreover, if ^(AO denotes the (traditionally defined) lag of the population 
consisting of TV cells, then 

N N 

-Hi,. 




,1" ,1 



e 



X(N) = --\n-^ = - — ln-^ (9.7) 

\L N \i N 

This formula is demonstrated in Figure 9.5. The more cells there are in the inoculum, 
the shorter is the expected population lag, converging to a limit value, A, min . Notice 
that the correct model for single-cell-generated curve is bi-phasic; the curvature for 
higher inoculum is caused by the distribution of the individual lag times. 

Various authors have reported that at a very low inoculum (approximate to 1 
cell per total volume of sample), the lag time is longer than expected (Augustin et 
al., 2000b; Gay et al., 1996; Stephens et al., 1997). However, as the above formula 
shows, this finding can also be explained by statistical means without extra biological 
assumptions. The lag time decreases with the initial cell number because the expo- 
nentially growing subgenerations of the cells with the shortest lag times will quickly 
dominate the whole population. 

In particular, Baranyi and Pin (2001) showed the mathematical relation between 
the respective parameters (mean and variance) of the individual lag times and those 
of the population consisting of TV cells. From these formulae, the distribution of 
individual lag times is per se different from the distribution of the population lag 
times. This was confirmed by the observation of Stephens et al. (1997) on heat- 

2004 by Robin C. McKellar and Xuewen Lu 










1237_C09.fm Page 294 Wednesday, November 12, 2003 1:08 PM 





~V 






« 3 



o 
o 




y =0 



MD=Tl 



time (h) 



FIG U RE 9.5 Simulating four single-cell-generated growth curves (y = 0) growing separately, 
and two separate growth curves generated by four cells each (y = In 4). The traditional lag 
times X of the subpopulations are estimated by curve fitting. Observe that the X(N) mean 
population lag converges to a limit value X Um while its variance (the spread of the growth 
curves as a function of their inoculum) converges to zero (see Figure 9.6). Symbols: simulated 
points; continuous thin lines: fitted curves; continuous thick lines: expected curves. 

injured Salmonella cells. The calculations provided another reason why the lag time 
should not be considered as a primary growth parameter but as the result of the initial 
physiological state (oc ), the inoculum size and the maximum specific growth rate. 
Smelt et al. (2002) estimated the population lag as the minimum of the individual 
lag times. Our formula shows that the population lag is greater than the minimum 
of the individual lag times, but closer to it than to the arithmetical mean. 




9.3.3 A Simulation Rogram 

To demonstrate the effect of the inoculum on the distribution of the population lag, 
we developed a simulation program. One hundred growth curves from different 
inoculum sizes were generated. The distribution functions used by the simulation 
were based on the theory of Baranyi and Pin (2001). 

We assumed that the individual lag times followed the gamma distribution, with 
the parameters p and v, so the mean individual lag is T = /?/V. The biological 
interpretation of this scenario is that the cells have to carry out p consecutive tasks, 
and the times required by the individual tasks are independent, exponentially dis- 
tributed random variables, with the parameter v (i.e., the average time required for 
each task was 1/v). 

The parameter p represented the amount of work to be done during the lag time 
(h = p). Partly for simplicity, partly for mechanistic reasons, v = |l max was assumed. 
The generation times, g j9 of individual cells, after the lag phase, followed the 
exponential distribution with the parameter |l raax . 

With the usual terminology for stochastic processes, the initial state of the system 
is (M0), meaning that there are N cells in the lag phase and in the exponential 
phase at the starting time (t ). The time t x of the first division in the population is 





2004 by Robin C. McKellar and Xuewen Lu 










1237_C09.fm Page 295 Wednesday, November 12, 2003 1:08 PM 





~V 





that of the cell with the smallest value for the sum of its lag time, x is and its first 
generation time, g t . 

t x =min(x 1 + g u ... X N + g N ) 

After the first division, N- 1 cells remain in lag phase while 2 cells are in exponential 
growth phase. Thus the state of the system is (N - 1, 2). The time t 2 of the second 
division will be the minimum of the first division times among the N - I cells 
remaining in the lag phase and the division times of the two daughter cells: 

t 2 = min(x, + g l9 ... X N _ { + g N _ l9 t x + g N+v f, + g N+2 ) 
This second division can happen to a cell in exponential or in lag phase. 




1. If a cell in lag phase divides, then the state of the system will be (N - 2, 
4) and the time of the third division will be t 3 = min(x y + g l3 ..., T N _ 2 + 

£/v-2> h + 8n+u h + 8n+2-> h + 8n+3> h + 8n+a)- 

2. If a cell in exponential phase divides, then the system jumps into the 

state (N - 1,3). Apart from numbering the daughter cells, the time to the 
third division will be t 3 = min^ + g l9 ..., x N _ { + g N _ u t x + g N+2 , t 2 + g N+3 , 

h + ^^+4)- 

The iteration can be continued in a similar manner and this is what our simulation 
program did, until the population had a given number of cells. One hundred growth 
curves were generated for each selected inoculum size. The parameters h and |l max 
were chosen close to practical values: h = 4 and |l max = 0.5 h -1 . Thus the gamma 
parameters to simulate the individual lag times were p = 4 and v = 0.5 Ir 1 (mean 
individual lag = 8 h) and the parameter of the exponential distribution to simulate 
the individual generation times was 0.5 Ir 1 (mean generation time = 2 h). The 
traditionally interpreted lag time of each population curve was estimated by fitting 
a biphasic (no upper asymptote) version of the model of Baranyi and Roberts (1994). 
Figure 9.5 shows typical examples of the generated population curves (originating 
from one and four cells). As can be seen, in the case of a single initial cell, when 
the inoculum is In x = 0, the population lag (X, measured by the tangent to the 
exponential phase) is close to the theoretical individual lag (x, = 8 h). 
If the inoculum size increases then, from the formula 




r 



UN) 



N->c 



urn 



Ln 



M 



1 + 



v 



MX 
P 



\ 



(9.8) 



/ 



the population lag converges to A aim = 8 -In 2 = 5.54. This convergence is shown in 
Figure 9.5. Figure 9.6 shows the distribution of the population lag times as a function 
of the inoculum size. 





2004 by Robin C. McKellar and Xuewen Lu 










1237_C09.fm Page 296 Wednesday, November 12, 2003 1:08 PM 








Frequency 



Lag time 



Inoculum 



FIGU RE 9.6 As the inoculum size increases, the population lag time decreases, converging 
to a limit value. Its variance converges to zero. 

9.3.4 Lag/Survival Distribution of Sngle Cells and 
Lag /Shoulder of R) pulatio n Curves 

A general result from Baranyi (2002) connects the expected population growth curve 
with the distribution of single-cell kinetics: 




x(t) = N 



ft \ 

je" ir - s) f(s)ds+jf(s)ds 




Vo 



(9.9) 



/ 



where /(0 denotes the (common) probability density function (pdf) of the individual 
lag times. It is analogous to the formula of Kormendy et al. (1998) for death curves: 



x(t) = N 



oo 

J>*( 



s)ds 



(9.10) 



where g(s) is the probability density function of the individual cells' death rate (i.e., 
the reciprocal of their survival time). The big difference is that while the probability 
distribution of the individual survival times determines the whole death curve (actu- 
ally by a Laplace transformation, as the formula shows) that of the individual lag 
times does not. In both models, the first compartments are "dormant" states, but the 
second ones are totally different: for growth, it is the state of exponential growth, 
characterized by the specific growth rate; for survival curves, it is a "death phase" 
without any parameter. 

The formulae above represent a one-to-one mapping between the f(f) distribution 
and the population growth/death curve. The problem is that, in practice, this mapping 
works only in one direction. The expected population curve can be easily derived 
from the distribution of the individual lag times, but vice versa it would need 



2004 by Robin C. McKellar and Xuewen Lu 












1237_C09.fm Page 297 Wednesday, November 12, 2003 1:08 PM 







unrealistically accurate data. This is an example where the performance of stochastic 
techniques is much better than that of their deterministic counterparts. 

9.3.5 History Bfect on the Sho ulder Rriods 
of Survival Curves 

For survival curves, with the notations as above, let x t be the time for the zth cell to 
die, and assume that it is distributed exponentially, with the mean value x (pure 
Poissonian death process). In this case, F{t) = 1 - e~ t/x , so the survival curve is y(f) 
= \n N - t/l. Therefore, in this case, the survival curve is linear and there is no 
shoulder parameter: X = 0. 

Analogously with our approach to modeling the distribution of individual lag 
times for growth, consider now the situation when T i (i = 1,..., TV) follow the gamma 
distribution, with the parameters p > 1 and v > 0, where the expected value of the 
survival time for a cell is x = pN . This is interpreted by Baranyi and Pin (2001) in 
the following way: the cell needs p damaging hits (see the multihit theory in Casolari, 
1988), and the times ; (j = 1,...,/?) between the hits are independent, exponentially 
distributed variables, with a common mean value = 7/V. Then the survival time 
of the z'th cell is 




X e ; (9-1D 



x. = 

therefore x, (i = 1,..., A/) are gamma distributed. In this case, as shown by the 
authors, the derivative of y(t) converges to a constant, namely v, the limit shoulder 
still does not exist if p > 1. In practice, the shoulder would be measured by v and 
the smallest detectable value of y(t), which means that the shoulder measurement 
would depend on the precision of the measurement method! This inconsistency 
makes the shoulder parameter unsuitable for modeling. Instead, the p parameter 
quantifying the "damage to be done" should be modeled, which is analogous to the 
h parameter for growth curves. 

Another well-known interpretation of the shoulder is the so-called multitarget 
model (Hermann and Horst, 1970). According to that, a cell has p targets that are 
being hit synchronously (not consecutively as in the previous case). The times needed 
to destroy the targets, 6^ (j = 1,..., /?), are independent, exponentially distributed 
variables with the common mean value of 1/v and the cell is live until all targets 
are inactivated. Therefore, the survival time of the zth cell is 



X.=max6. (9.12) 

' \<j<p J 

and F(f) = (1 - e~ vt ) p , where p > 1. The population survival curve is now 

y(t) = In AT + ln(l - (1 - e- v 9") (9.13) 



2004 by Robin C. McKellar and Xuewen Lu 














~V 






1237_C09.fm Page 298 Wednesday, November 12, 2003 1:08 PM 






which converges to the 

y a (t) = In N - v • (t - In pN) (9. 14) 

linear asymptote. Therefore, the limit shoulder parameter can be calculated as 

X = \np/v (9.15) 

This shows another analogy to the modeling of the history effect on growth: the "In 
/?" quantity could be used to characterize the initial "damage to be done," and the 
shoulder period depends on the magnitude of the initial damage and on the rate as 
the damage is being done. 

9.4 CONCLUDING REMARKS 

In this chapter, we showed that the history effect can be quantified either by the 
"initial physiological state" or the "work to be done" parameters; they are simply 
rescaled versions of each other. The traditional lag parameter has become a derived 
one, useful for static conditions only. By introducing a variable for the rate at which 
the work is carried out before the exponential phase, the model can be used for 
dynamic conditions, too. This concept was successfully applied in a series of papers 
(Baranyi et al., 1995, 1996) describing the entire bacterial growth profile (not only 
the exponential phase) under fluctuating conditions. 

The priority of a is purely a modeling concept; it does not mean that when 
fitting a bacterial growth curve, oc and not the lag should be estimated. The formula 
X = -lna /|i. max to calculate the lag expresses the view that both the history and the 
current environment influence the lag. This has practical consequences when devel- 
oping more complex, dynamic models or, for example, when developing predictive 
software packages. These commonly require the user to provide values for the 
environmental factors, from which the maximum specific rate (or one of its rescaled 
versions) and a default lag are predicted. If the user gives the inoculum, then the 
whole growth curve can be constructed, too. What our theory suggests is that it is 
not only the inoculum that cannot be predicted from the current environment, but 
also the initial physiological state, oc . This is shown in Figure 9.2: there are two 
independent initial points on the y axis: 




1. An inoculum value (y ) 

2. A point through which the tangent drawn to the inflexion will pass (which 
is h = -lna lower than the inoculum) 

Sometimes the value of the initial physiological state should be simply a = 1, 
expressing the fact that the preinoculation conditions are the same as the current 
environment. 

As has been mentioned, many features of the shoulder period before the expo- 
nential death phase can be discussed analogously to those of the lag period before 

2004 by Robin C. McKellar and Xuewen Lu 










1237_C09.fm Page 299 Wednesday, November 12, 2003 1:08 PM 




~V 









the exponential growth phase. Baranyi and Pin (2001) showed that the multitarget 
theory of inactivation (Hermann and Horst, 1970), modeling the shoulder before the 
exponential death rate, provides a parallel between the initial "damage to be done" 
for death and the h Q "work to be done" for growth. The "bridge" between the history 
and the current environment is characterized by h either for growth (Figure 9.2) or 
death (the mirror image of Figure 9.2). 

The question can be raised of why the distribution of individual lag times is 
important once different distributions can result in practically the same population 
curve. The answer lies in the interest of Quantitative Microbial Risk Assessment in 
the distribution of the quantity "time to a certain level" (such as time to infective 
dose or time to legally allowed concentration). The distribution of that time depends 
heavily on the distribution of individual lag times. 

Modeling the effect of history is more efficient with stochastic techniques, but 
also more data- and maths-demanding. Deterministic models are able to produce 
"averaged out" solutions only, and studying the variability around those predictions 
is vital in Quantitative Microbial Risk Assessment. 

ACKNOWLEDGMENT 

The authors are indebted to Susie George for preparing the manuscript. This work 
was funded by the EC project QLRTD-2000-01145 and the IFR project CSG 
434.1213A. 



REFERENCES 

Augustin, J.C., L. Rosso, and V. Carlier. 2000a. A model describing the effect of temperature 

history on lag time for Listeria monocytogenes. Int. J. Food Microbiol. 57, 169-181. 
Augustin, J.C., A. Brouillaud Delattre, L. Rosso, and V. Carlier. 2000b. Significance of 

inoculum size in the lag time of Listeria monocytogenes. Appl. Environ. Microbiol. 

66, 1706-1710. 
Augustin, J.C. and V. Carlier. 2000. Mathematical modelling of the growth rate and lag time 

for Listeria monocytogenes. Int. J. Food Microbiol. 56, 29-51. 
Baranyi, J. 1998. Comparison of stochastic and deterministic concepts of bacterial lag. 

J. Theor. Biol. 192, 403-408. 
Baranyi, J. 2002. Stochastic modelling of bacterial lag phase. Int. J. Food Microbiol. 73, 

203-206. 
Baranyi, J. and C. Pin. 2001. A parallel study on bacterial growth and inactivation. /. Theor. 

Biol. 210, 327-336. 
Baranyi, J. and T.A. Roberts. 1994. A dynamic approach to predicting bacterial growth in 

food. Int. J. Food Microbiol. 23, 277-294. 
Baranyi, J., A. Jones, C. Walker, A. Kaloti, and B.M. Mackey. 1996. A combined model for 

growth and thermal inactivation of Brochothrix thermosphacta. Appl. Env. Microbiol. 

62, 1029-1035. 
Baranyi, J., T.R Robinson, A. Kaloti, and B.M. Mackey. 1995. Predicting growth of Brocho- 
thrix thermosphacta at changing temperature. Int. J. Food Microbiol. 27, 61-75. 
Beumer, R.R., M.C. te Giffel, E. de Boer, and EM. Rombouts. 1996. Growth of Listeria 

monocytogenes on sliced cooked meat products. Food Microbiol. 13, 333-340. 

2004 by Robin C. McKellar and Xuewen Lu 











1237_C09.fm Page 300 Wednesday, November 12, 2003 1:08 PM 








Breand, S., G. Fardel, J.P. Flandrois, L. Rosso, and R. Tomassone. 1999. A model describing 

the relationship between regrowth lag time and mild temperature increase for Listeria 

monocytogenes. Int. J. Food Microbiol. 46, 251-261. 
Buchanan, R.L. and L.A. Klavitter. 1991. Effect of temperature history on the growth of 

Listeria monocytogenes at refrigeration temperatures. Int. J. Food Microbiol. 12, 

235-246. 
Casolari, A. (1988). Microbial death. In Physiological Models in Microbiology. Vol. II, Bazin, 

M.J. and Prosser, J.I., Eds. CRC Press, Boca Raton, FL, Ch. 7. 
Cotterill, O.J. and J. Glauert. 1969. Thermal resistance of Salmonellae in egg yolk products 

containing sugar or salt. Poult. Sci. 48, 1156-1166. 
Delignette-Muller, M.L. 1998. Relation between the generation time and the lag time of 

bacterial growth kinetics. Int. J. Food Microbiol. 43, 97-104. 
Dufrenne, J., E. Delfgou, W. Ritmeester, and S. Notermans. 1997. The effect of previous 

growth conditions on the lag phase time of some foodborne pathogenic micro- 
organisms. Int. J. Food Microbiol. 34, 89-94. 
Gay, M., O. Cerf, and K.R. Davey. 1996. Significance of pre-incubation temperature and 

inoculum concentration on subsequent growth of Listeria monocytogenes at 14°C. 

J.Appl. Bacteriol. 81, 433-438. 
Hermann, D. and J. Horst. 1970. Molecular Radiation Biology. Springer, Berlin. 
Hudson, J. A. 1993. Effect of pre-incubation temperature on the lag time of Aeromonas 

hydrophila. Lett. Appl. Microbiol. 16, 274-276. 
Kormendy, I., L. Kormendy, and A. Ferenczy. 1998. Thermal inactivation kinetics of mixed 

microbial populations. A hypothesis paper. J. Food Eng. 38, 439-453. 
Mackey, B.M. and A.L. Kerridge. 1988. The effect of incubation temperature and inoculum 

size on growth of Salmonellae in minced beef. Int. J. Food Microbiol. 6, 57-65. 
Manas, P., R. Pagan, I. Leguerinel, S. Condon, P. Mafart, and F. Sala. 2001. Effect of sodium 

chloride concentration on the heat resistance and recovery of Salmonella typhimu- 

rium. Int. J. Food Microbiol. 63, 209-216. 
McKellar, R.C., X. Lu, and K.P Knight. 2002a. Growth pH does not affect the initial 

physiological state parameter (p ) of Listeria monocytogenes cells. Int. J. Food Micro- 
biol. 73, 137-144. 
McKellar, R.C., X. Lu, and K.P. Knight. 2002b. Proposal of a novel parameter to describe 

the influence of pH on the lag phase of Listeria monocytogenes. Int. J. Food Microbiol. 

73, 127-135. 
McMeekin, T.A., J. Olley, D.A Ratkowsky, and T. Ross. 2002. Predictive microbiology: 

towards the interface and beyond. Int. J. Food Microbiol. 73, 395-407. 
Membre, J.M., T. Ross, and T. McMeekin. 1999. Behaviour of Listeria monocytogenes under 

combined chilling processes. Lett. Appl. Microbiol. 28, 216-220. 
Ng, H., J.L. Ingraham, and A.G. Marr. 1962. Damage and depression in Escherichia coli 

resulting from growth at low temperatures. /. Bacteriol. 84, 331-339. 
Pin, C., G.D. Garcia de Fernando, J.A. Ordonez, and J. Baranyi. 2002. Analysing the lag- 
growth rate relationship of Yersinia ente roc otitic a. Int. J. Food Microbiol. 73, 

197-201. 
Pirt, S.J. 1975. Principles of Microbe and Cell Cultivation. Blackwell, London. 
Robinson, T.P., M.J. Ocio, A. Kaloti, and B.M. Mackey. 1998. The effect of the growth 

environment on the lag phase of Listeria monocytogenes. Int. J. Food Microbiol. 44, 

83-92. 
Sherman, J.M. and W.R. Albus. 1923. Physiological youth in bacteria. J. Bacteriol. 8, 27-139. 




2004 by Robin C. McKellar and Xuewen Lu 












1237_C09.fm Page 301 Wednesday, November 12, 2003 1:08 PM 







Smelt, J.P.P.M., G.D. Otten, and A.P. Bos. 2002. Modelling the effect of sublethal injury on 
the distribution of the lag times of individual cells of Lactobacillus plantarum. Int. 
J. Food Microbiol. 73, 207-212. 

Stephens, P.J., J. A. Joynson, K.W. Davies, R. Holbrook, H.M. Lappinscott, and T.J. Humphrey. 
1997. The use of an automated growth analyser to measure recovery times of single 
heat-injured Salmonella cells. /. Appl. Microbiol. 83, 445-455. 

Strange, R.E. and M. Shon. 1964. Effects of thermal stress on viability and ribonucleic acid 
of Aerobacter aerogenes in aqueous suspension. J. Gen. Microbiol. 34, 99-114. 

Walker, S.J., P. Archer, and J.G. Banks. 1990. Growth of Listeria monocytogenes at refriger- 
ation temperatures. J. Appl. Bacteriol. 687, 157-162. 

Wang, C. and L.A. Shelef. 1992. Behaviour of Listeria monocytogenes and the spoilage 
microflora in fresh cod fish treated with lysozyme and EDTA. Food Microbiol. 9, 
207-213. 

Whiting, R.C. and L.K. Bagi. 2002. Modelling the lag phase of Listeria monocytogenes. Int. 
J. Food Microbiol. 73, 291-295. 





2004 by Robin C. McKellar and Xuewen Lu