247
CALORIMETRIC AND MICROBIOLOGICAL EVALUATION OF BACTERIA AFTER EXPOSURE TO FOOD PRESERVATION TREATMENTS DISSERTATION Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The Ohio State University By Jaesung Lee, M. S. * * * * * The Ohio State University 2004 Dissertation Committee: Approved By Dr. Gönül Kaletunç, Adviser Dr. Polly D. Courtney Dr. Michael E. Mangino Adviser Dr. Olli H. Tuovinen Food Science and Nutrition Graduate Program

Calorimetric and Microbiological Evaluation of Bacteria

Embed Size (px)

Citation preview

Page 1: Calorimetric and Microbiological Evaluation of Bacteria

CALORIMETRIC AND MICROBIOLOGICAL EVALUATION OF BACTERIA

AFTER EXPOSURE TO FOOD PRESERVATION TREATMENTS

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree

Doctor of Philosophy in the Graduate School of

The Ohio State University

By

Jaesung Lee, M. S.

* * * * *

The Ohio State University 2004

Dissertation Committee: Approved By

Dr. Gönül Kaletunç, Adviser

Dr. Polly D. Courtney

Dr. Michael E. Mangino Adviser

Dr. Olli H. Tuovinen Food Science and Nutrition Graduate Program

Page 2: Calorimetric and Microbiological Evaluation of Bacteria

ABSTRACT

Thermal and non-thermal food preservation treatments affect cellular components

of foodborne microorganisms that cause physiological changes in cells and eventually

death of bacteria. Differential scanning calorimetry (DSC) thermograms of whole

bacterial cells display thermally-induced transitions revealing the response of bacteria to

heat under linearly increasing temperature condition. Therefore, DSC of the whole

microbial cells can allow the detection in vivo of changes in their cellular components

including ribosomes, nucleic acids, proteins and cell envelopes. The main purpose of this

study was to evaluate the effects of physical and chemical treatments on microorganisms

based on the changes in thermal stability (Tm) of the cellular components and the total

apparent enthalpy (∆H) from the calorimetric data. To compare with DSC data, the

viability data from microbiological methods (plate counting) was also evaluated.

The viability and the change in the thermal stability of individual transitions of

Escherichia coli and Lactobacillus plantarum were evaluated after pre-heating in the

DSC to various temperatures. The fractional viability based on calorimetric data [(∆H-

∆Hf)/(∆H0-∆Hf)] and plate count data (N/N0) showed a linear relationship. Viability loss

and the irreversible changes in DSC thermograms of whole cells pre-treated in DSC to

various temperatures were highly correlated between 55 and 70oC. Comparison of DSC

ii

Page 3: Calorimetric and Microbiological Evaluation of Bacteria

scans for isolated ribosomes showed that the thermal stability of ribosomes from E. coli

is greater than the thermal stability of L. plantarum ribosomes, consistent with the greater

thermal tolerance of E. coli observed from viability loss and DSC scans of whole cells.

The apparent enthalpy data obtained from DSC of E. coli cells were applied to a

theoretical formalism to predict the number of surviving microorganisms as a function of

linearly increasing temperature. The decimal reduction time (D) and thermal resistance

constant (z) values for E. coli determined from the calorimetric data were compared to

the corresponding values from plate count data obtained after heat treatment in the DSC

and after isothermal treatment to validate the new approach. The calculated D values

using both apparent enthalpy and viability data for cells heat treated in the DSC were

similar to the D values obtained from isothermal treatment. Temperatures for 1 through

10-log microbial population reductions, calculated from plate count and enthalpy data

were in agreement within 0.5-2.4oC at a 4oC min-1 heating rate.

The effect of chemical agents (acids, ethanol or NaCl) on the cellular components

and the survivability in subsequent heat treatment of E. coli was evaluated using DSC

and viable counting methods. The thermal stability for ribosomal subunits denaturation

decreased as concentration of the chemical agents increased. The apparent enthalpy also

decreased, mainly due to reduction of ribosomal subunit peak as the concentration

increased. The size and thermal stability of DNA transition were reduced by inorganic

(HCl) and organic (CH3COOH) acid treatments. The survival of cells received chemical

treatments was lower than that of non-treated cells after mild heat treatment (at 60, 62.5

and 65oC) indicating that the conformational changes in cellular components by the

chemical treatments may cause sensitization to heat.

iii

Page 4: Calorimetric and Microbiological Evaluation of Bacteria

High hydrostatic pressure (HHP)-induced changes in cell structures of E. coli were

determined using DSC and electron microscopy (EM) to relate the structural changes to

viability of the cells. The reversibility of transition and the change in the thermal stability

of ribosome of E. coli are affected by 200 MPa and above pressures in HHP treatment.

The enthalpy and the thermal stability of the DNA melting transition were reduced by

HHP treatments above 300 MPa. The pressure-induced changes in ribosome and DNA

were also detected in thin sections under transmission electron microscopy. In EM study,

integrity of cell envelope was maintained in pressure- or heat-inactivated cells; however,

the leakage of cell wall or outer membrane substance and empty space between cell

envelope and inside structure were exclusively observed in pressure- inactivated cells.

The effects of HHP and nisin treatment alone and in combination on cellular

components and viability of Salmonella enterica subsp. enterica serova Enteritidis

(Salmonella Enteritidis) FDA and OSU 799 strains were evaluated by DSC and plate

counting in order to evaluate the relative resistance and to optimize the treatment

conditions. An 8-log cfu/ml reduction was observed after a pressure treatment at 500

MPa for FDA strain and 450 MPa for the OSU 799 strain. When nisin was added, a

similar reduction was obtained at 400 MPa for FDA strain and 350 MPa for the OSU 799

strain. The decrease in apparent enthalpy appeared to be mainly due to reduction in the

ribosome denaturation peak for the pressure alone and nisin-combination treatments.

DNA might be irreversibly damaged by the combination treatments. There is a linear

relationship in a logarithmic plot of fractional apparent enthalpy values [(∆H-∆Hf)/(∆H0-

∆Hf)] versus the fractional survivors from plate count data (N/N0) for treated cells.

iv

Page 5: Calorimetric and Microbiological Evaluation of Bacteria

Dedicated to my parents and my daughters

v

Page 6: Calorimetric and Microbiological Evaluation of Bacteria

ACKNOWLEDGMENTS

I thank Dr. Gönül Kaletunç, my adviser, for her guidance, encouragement, patience

and suggestions throughout this work. I also to express my gratitude to committee

members, Dr. Olli H. Tuovinen, Dr. Michael E. Mangino and Dr. Polly D. Courtney, for

their valuable criticism, suggestions and comments.

I wish to express appreciation to my colleagues, Kelley Yosik, Hyunjung Chung,

and U.C. Rakhith for their help. I thank the USDA laboratory for the instrumental

support to complete my research.

I am especially grateful to my parents and sisters for their concern, encouragement

and moral support throughout my education. Finally, this dissertation is dedicated to Jiae

Park, my wife, and Yunjung Lee and Yunmi Lee, my daughters, for their love, support,

and patience.

vi

Page 7: Calorimetric and Microbiological Evaluation of Bacteria

VITA

January 5, 1964 ……………………………….

Born in Inchon, Korea

1994 …………………………………………... B.S., Microbiology, University of

Minnesota, Minneapolis, MN

1999 …………………………………………... M.S., Food Science, University of

Delaware, Newark, DE

1999-present ………………………………….. Graduate Research Associate, The Ohio

State University, Columbus, OH

PUBLICATIONS

Lee, J. and Kaletunç, G. 2002. Evaluation by differential scanning calorimetry of the heat inactivation of Escherichia coli and Lactobacillus plantarum. Appl. Environ. Microbiol. 68:5379-5386. Lee, J. and Kaletunç, G. 2002. Calorimetric determination of inactivation parameters of microorganisms. J. Appl. Microbiol. 93:178-189. Alpas, H., Lee, J., Bozoglu, F. and Kaletunç, G. 2003. Differential scanning calorimetry of pressure-resistant and pressure-sensitive strains of Staphylococcus aureus and Escherichia coli O157:H7. Int. J. Food Microbiol. 87:229-237. Kaletunç, G., Lee, J., Alpas, H. and Bozoglu, F. 2004. Evaluation of structural changes induced by high hydrostatic pressure in Leuconostoc mesenteroides. Appl. Environ. Microbiol. 70:1116-1122.

vii

Page 8: Calorimetric and Microbiological Evaluation of Bacteria

PUBLISHED ABSTRACTS

Lee, J. and Kaletunç, G. Inactivation of Salmonella Enteritidis treated by a combination of high hydrostatic pressure (HHP) and Nisin: A calorimetric investigation. American Society of Microbiology General Meeting, Washington, DC, 2003. Lee, J. and Kaletunç, G. Calorimetric evaluation of the thermal stability of ribosomes isolated from Escherichia coli and Lactobacillus plantarum. Institute of Food Technology (IFT) Annual Meeting, Anaheim, CA, 2002. Lee, J., Alpas, H., Bozoglu, F. and Kaletunç, G. Studies on the effect of high hydrostatic pressure (HHP) on cell morphology of Leuconostoc mesenteroides with scanning electron microscopy (SEM). IFT Annual Meeting, Anaheim, CA, 2002. Lee, J. and Kaletunç, G. Calorimetric determination of microbial survival curve of Escherichia coli. IFT Annual Meeting, New Orleans, LA. 2001. Lee, J. and Kaletunç, G. Calorimetric evaluation of chemically stressed Escherichia coli. Ohio Valley IFT, Columbus, OH. 2001. Lee, J. and Kaletunç, G. Evaluation of the influence of environmental stresses on inactivation of microorganisms by differential scanning calorimetry. IFT Annual Meeting, Dallas, TX, 2000.

FIELDS OF STUDY

Major Field: Food Science and Nutrition

viii

Page 9: Calorimetric and Microbiological Evaluation of Bacteria

TABLE OF CONTENTS Page ABSTRACT………………………………………………………………………… ii

DEDICATION……………………………………………………………………… v

ACKNOWLEDGMENTS…………………………………………………………. vi

VITA………………………………………………………………………………... vii

LIST OF TABLES………………………………………………………………….. xii

LIST OF FIGURES………………………………………………………………… xiii

INTRODUCTION…………………………………………………………………. 1

Chapters:

1. Literature Review………………………………………………………………... 3

Differential Scanning Calorimetry (DSC)………………………………………. 3 Principles of DSC………………………………………………………………. 3 Bacterial Thermal Analysis by Differential Scanning Calorimetry……………. 6

Thermal Processing Effect on Microorganisms…………………………………. 11 Effect on cell components……………………………………………………… 11 Mechanism of cell death……………………………………………………….. 16

Chemical Effect on Microorganisms……………………………………………. 18 Hurdle technology……………………………………………………………… 18 Effect of ethanol………………………………………………………………... 18 Effect of NaCl………………………………………………………………….. 19 Effect of acids…………………………………………………………………. 20

High hydrostatic pressure (HHP)……………………………………………….. 20 HHP technology for food preservation………………………………………… 21 Effects of HHP on microorganisms……………………………………………. 22 HHP in combination with other processing technologies……………………… 24 HHP in combination with bacteriocins………………………………………… 25

ix

Page 10: Calorimetric and Microbiological Evaluation of Bacteria

2. Evaluation by differential scanning calorimetry of the heat inactivation of

Escherichia coli and Lactobacillus plantarum………………………………….

28

Abstracts………………………………………………………………………... 28 Introduction…………………………………………………………………….. 30 Materials and Methods…………………………………………………………. 33 Results………………………………………………………………………….. 39 Discussions…………………………………………………………………….. 55 References……………………………………………………………………… 63 3. Calorimetric determination of inactivation parameters of microorganisms ……. 66 Abstracts………………………………………………………………………... 66 Introduction…………………………………………………………………….. 68 Materials and Methods…………………………………………………………. 71 Theory………………………………………………………………………….. 75 Results………………………………………………………………………….. 80 Discussions…………………………………………………………………….. 86 References……………………………………………………………………… 102 4. Evaluation by differential scanning calorimetry of the effects of ethanol, NaCl,

acetic acid and pH on Escherichia coli …………………………………………

107 Abstracts………………………………………………………………………... 107 Introduction…………………………………………………………………….. 109 Materials and Methods…………………………………………………………. 111 Results………………………………………………………………………….. 115 Discussions…………………………………………………………………….. 125 References……………………………………………………………………… 131 5. Evaluation of viability and structural changes induced by high hydrostatic

pressure in Escherichia coli ……….……………………………………………

137 Abstracts………………………………………………………………………... 137 Introduction…………………………………………………………………….. 139 Materials and Methods…………………………………………………………. 141 Results………………………………………………………………………….. 146 Discussions…………………………………………………………………….. 159 References……………………………………………………………………… 167 6. Inactivation of Salmonella Enteritidis FDA by combination of high hydrostatic

pressure and nisin.……………………………………………………………….

172

x

Page 11: Calorimetric and Microbiological Evaluation of Bacteria

Abstracts………………………………………………………………………... 172 Introduction…………………………………………………………………….. 174 Materials and Methods…………………………………………………………. 177 Results………………………………………………………………………….. 182 Discussions…………………………………………………………………….. 195 References……………………………………………………………………… 202 General Conclusions……………………………………………………….……….. 208 Bibliography…………………………………………………………………….….. 212 Appendix. Figures and Table of the evaluation of Salmonella Enteritidis inactivation after HHP treatment with different concentrations of nisin……………

225

xi

Page 12: Calorimetric and Microbiological Evaluation of Bacteria

LIST OF TABLES

Table Page

1.1 The G+C content of each strain DNA and the temperature recorded for peak associated with the melting of putative DNA from whole cell DSC………….

9

1.2 Major transition temperatures in the thermograms of whole cells of E. coli….. 10 1.3 Survivors of four pathogens by pressurizing in the absence and presence of

bacteriocins……………………………………………………………………

27 2.1 Transition temperature and apparent enthalpy values for E. coli and L.

plantarum ribosomes after DSC in different pH………………………………

44 3.1 Viability and apparent enthalpy values for E. coli cells after pre-treatment in

the DSC……………………………………………………………………….

83 3.2 D and z values reported for E. coli from isothermal and non-isothermal heat

treatments……………………………………………………………………...

84 4.1 Effects of chemicals on viability and calorimetry of E. coli………………….. 116 5.1 Viability, apparent enthalpy values and transition temperatures of each peak

for E. coli cells after treatments…………………………………….…………

147 6.1 Viability and apparent enthalpy values for cells of Salmonella Enteritidis

strains after HHP treatments…………………………………………………..

186 Appendix. 1 Viability and apparent enthalpy values for cells of Salmonella

Enteritidis FDA after HHP treatments in combination with nisin…………….

231

xii

Page 13: Calorimetric and Microbiological Evaluation of Bacteria

LIST OF FIGURES

Figure Page 1.1 Chamber of cylindrical type DSC …………………………………………….. 4 1.2 Diagram of the DSC chamber…………………………………………………. 4 1.3 Typical DSC curve of starch.………………………………………………….. 5 1.4 Scheme for the sequence of events leading to the death of microorganism

from heating…………………………………………………………………...

17 2.1 DSC thermogram of whole cells of E. coli ATCC 14948…………………….. 32 2.2 Experimental scheme of calorimetric and microbial analysis…………………. 38 2.3 Thermograms of whole cells of E. coli and L. plantarum obtained by DSC….. 40 2.4 Thermograms of isolated intact ribosomes of E. coli and L. plantarum

obtained by DSC………………………………………………………………

42 2.5 Thermograms of whole cells (A) and isolated intact ribosomes (B) of E. coli

(a) and L. plantarum (b) obtained by DSC after HEPES buffer (pH 7.5) wash.…………………………………………………………………………..

43 2.6 DSC thermogram of isolate ribosome of E. coli (a) and L. plantarum (b) at

different pH of phosphate buffer………………………………………………

45 2.7 Effect of heat pre-treatment on the thermogram of E. coli……………………. 48 2.8 Effect of heat pre-treatment on the thermogram of L. plantarum……………... 49 2.9 Viable counts and DSC thermograms of E. coli after heat pre-treatment at

60oC, 62.5oC, 64oC, 65oC and 70oC……………………………………….…..

52 2.10 Viable counts and DSC thermograms of L. plantarum after heat pre-treatment

at 55oC, 57.5oC, 60oC, 65oC and 70oC………………………………………...

53

xiii

Page 14: Calorimetric and Microbiological Evaluation of Bacteria

2.11 Correlation between fractional apparent enthalpy and fractional viability for

E. coli and L. plantarum………………………………………………………

54 3.1 Experimental scheme of calorimetric and microbial analysis…………………. 79 3.2 Apparent specific heat capacity versus temperature curves of control and

heat-treated E. coli…………………………………………………………….

81 3.3 A typical DSC thermogram for whole cells of E. coli K12 after empty

crucible baseline subtraction…………………………………………………..

89 3.4 DSC thermogram for whole cells of E. coli K12 displaying curve base line

used to determine the apparent enthalpy value………………………………..

90 3.5 Temperature dependence of fractional survivor population determined from

plate count data after heat pre-treatment of E. coli cells in the DSC………….

92 3.6 Temperature dependence of fractional survivor population determined from

calorimetric data after heat pre-treatment of E. coli cells in the DSC………...

93 3.7 Comparison of D values calculated from the calorimetric and viability data

obtained under non-isothermal heat treatment in the DSC and D values obtained from isothermal heat treatment………………………………………

96 3.8 Comparison of D values calculated from the calorimetric and viability data

obtained under non-isothermal heat treatment in the DSC and D values obtained from isothermal heat treatment………………………………………

99 4.1 Experimental scheme of calorimetric and microbial analysis…………………. 114 4.2 DSC thermogram of E. coli pellet after ethanol treatment…………………….. 117 4.3 DSC thermogram of E. coli pellet after NaCl treatment………………………. 119 4.4 DSC thermogram of E. coli pellet after inorganic acid (HCl) treatment……… 120 4.5 DSC thermogram of E. coli pellet after organic acid (acetic acid) treatment…. 122 4.6 Survival of untreated and chemically treated E. coli after heat treatment under

linearly increasing temperature………………………………………………..

124 5.1 Experimental scheme of calorimetric, EM and microbial analysis……………. 145 5.2 Pressure dependence of fractional viability determined by plate count……….. 146

xiv

Page 15: Calorimetric and Microbiological Evaluation of Bacteria

5.3 DSC thermograms of pellets of E. coli whole cell after HHP (35oC for 5 min) or heat (65oC for 6 min) treatments…………………………………………...

150

5.4 Pressure dependence of fractional apparent enthalpy determined by DSC…… 151 5.5 Correlation between fractional apparent enthalpy and fractional viability for

HHP treated E. coli……………………………………………………………

152 5.6 SEM micrograph of control (a), pressure-inactivated (b, at 700 MPa, 35oC, 5

min), and heat-inactivated (c, at 65oC for 6 min) E. coli cells.………………..

153 5.7 TEM micrographs of untreated (a), pressure-inactivated (b, at 700 MPa, 35oC,

5 min), and heat-inactivated (c, at 65oC for 6 min) E. coli cells.……………...

156 6.1 Experimental scheme of calorimetric and microbial analysis…………………. 181 6.2 Thermograms of whole cells of Salmonella Enteritidis OSU 799 and

Salmonella Enteritidis FDA obtained by DSC………………………………..

183 6.3 Pressure dependence of fractional viability of Salmonella Enteritidis strains

determined by plate count. The cells treated with or without nisin…………..

187 6.4 DSC thermograms of Salmonella Enteritidis FDA pellets after pressure alone

treatment (a) and pressure-nisin combination treatment (b)…………………..

190 6.5 DSC thermograms of Salmonella Enteritidis OSU 799 pellets after pressure

alone treatment (a) and pressure-nisin combination treatment (b………….….

191 6.6 Pressure dependence of fractional viability of Salmonella Enteritidis strains

determined by calorimetric data. The cells treated with or without nisin…….

192 6.7 Correlation between fractional apparent enthalpy and fractional viability for

Salmonella Enteritidis FDA after HHP treatment……………………………..

193 6.8 Correlation between fractional apparent enthalpy and fractional viability for

Salmonella Enteritidis OSU 799 after HHP treatment………………………..

194 Appendix.1 DSC thermograms of Salmonella Enteritidis FDA pellets after

combinations of pressure and nisin (200 IU/ml) treatments..…………………

226 Appendix.2 DSC thermograms of Salmonella Enteritidis FDA pellets after

combinations of pressure and nisin (400 IU/ml Nisaplin) treatments………...

227

xv

Page 16: Calorimetric and Microbiological Evaluation of Bacteria

Appendix.3 DSC thermograms of Salmonella Enteritidis FDA pellets after combinations of pressure and nisin (600 IU/ml Nisaplin) treatments………...

228

Appendix.4 Pressure dependence of fractional viability of Salmonella Enteritidis

FDA determined by plate count……………………………………………….

229 Appendix.5 Pressure dependence of fractional viability of Salmonella Enteritidis

FDA determined by calorimetric data…………………………………………

230

xvi

Page 17: Calorimetric and Microbiological Evaluation of Bacteria

INTRODUCTION

Thermal and non-thermal processing technologies are widely applied in the food

industry for the preservation of food materials. The main goals of food preservation

technologies are to inactivate the spoilage and pathogenic microorganisms to produce a

safe product with enhanced shelf life. An understanding of the mechanism of the

microbial inactivation by physical and chemical stress is vital to assess the foodborne

disease and spoilage risk associated with food processing. Inactivation of

microorganisms results from irreversible denaturation of cell walls, membranes,

ribosomes, nucleic acids, and proteins such as enzymes by application of chemical and

physical stresses. Therefore, the investigation of the patterns of above macromolecular

changes that induce cell death during given treatments will provide knowledge for

designing optimal food process conditions.

Differential scanning calorimetry (DSC) thermograms of whole bacterial cells

display thermally-induced transitions revealing the response of bacteria to heat under

linearly increasing temperature condition. A number of overlapping transitions with a

net endothermic effect are observed when microorganisms are heated (Miles et al., 1986;

Anderson et al., 1991; Mackey et al., 1991; Belliveau et al., 1992; Kaletunç, 2001). The

observed transition peaks correspond to the denaturation of cellular components. A peak

1

Page 18: Calorimetric and Microbiological Evaluation of Bacteria

temperature corresponding to each transition represents the thermal stability of a cellular

component of bacteria. Mackey et al. (1991) investigated the origins of apparent

individual transitions on the thermogram of Escherichia coli. Individual peaks observed

in thermograms of whole cells of E. coli were assigned to cell components by comparing

the transition temperatures of isolated cell components with corresponding transitions in

whole cells. In addition, DSC measurement provides information about amount of heat

energy (apparent enthalpy, ∆H) associated with the transition. Therefore, DSC can be

utilized for thermal characterization of microorganisms before and after exposure to a

treatment to evaluate the impact of such treatment. Comparison of various final states

achieved under different treatment conditions starting from same initial state will allow

one to predict the effectiveness of various treatments for inactivation of bacteria.

My research focused on the investigation of changes in thermal properties of

cellular components after physical (heat, pressure) or chemical (acids, salt, ethanol and

nisin) treatment alone and in combination. Effectiveness of each treatment was evaluated

by calorimetric and microbiological data. Using the results of this study, a fundamental

understanding of the cause of those inactivation treatments of bacteria can be developed.

2

Page 19: Calorimetric and Microbiological Evaluation of Bacteria

CHAPTER 1

LITERATURE REVIEW

Differential Scanning Calorimetry

Principles of DSC DSC is a thermal analysis technique that measures heat flow

difference between a sample and a reference as a function of temperature at a fixed

heating rate (Chowdhry and Cole, 1989; Hohne et al. 1996). DSC detects and monitors

thermally induced conformational transitions and phase transitions when components in

sample are heated. DSC data allow to determine transition temperature (Tm), heat

capacity (Cp), and heat of transition (∆H) (Chowdhry and Cole, 1989; Hohne et al., 1996;

Kaletunç, 2001).

There are two types in DSC: power compensated DSC and Heat flux DSC. In the

power compensated DSC, the sample and reference materials are held in a separate

chamber with its own heater. When a thermal event occurs in the sample, power or

energy is applied to or removed from chamber(s) to compensate for the energy change

(heat flow) occurring in the sample. The amount of power required to maintain the

3

Page 20: Calorimetric and Microbiological Evaluation of Bacteria

system in equilibrium condition is proportional to the heat absorbed or released by the

sample. In a heat flux DSC (Calvet DSC 111, Setaram, Lyon, France), which was used

for this research, the sample and reference materials are heated in a single chamber (Fig.

1.1). There are two heating elements in the Calvet DSC sensor of the chamber, one on

top of the block, the other underneath. Compared with conventional heat flux DSC, in

which temperature is measured only through the bottom of the crucible by a

thermocouple, the Calvet DSC has greater accuracy and sensitivity on the temperature

measurement because the heat flux transducer containing 24 thermocouple wires located

inside a thermosated calorimetric block, fully surround each crucible. Such a design

allows to measure almost all the heat that is exchanged with the sample. Additional

advantage of using Calvet DSC is that it can be applied for larger amount of sample.

s

Heat element

Temperature probe

e

1

Crucibl

1

Figure 1.1. Chamber of DSC 11

4

Figure 1.2. Diagram of the DSC 11

Page 21: Calorimetric and Microbiological Evaluation of Bacteria

When thermal event occurs in the sample crucible, a temperature differential (dT) is

created between the sample and reference area. Thermocouples surrounded in both

crucibles detect dT per time (Fig. 1.2). DSC program converts the detected dT to heat

flow (Q, J/sec) versus temperature (T, oC) using an equation;

LkAdTQ −=)(Heat flow

where k is thermal conductivity of crucible (Q/moC); A is area through which heat flows

(m2); and L is thickness of crucible (m). As a result, the changes in the sample that are

associated with absorption or evolution of heat cause a change in the differential heat

flow which is then recorded as a DSC curve (thermogram, Fig. 1.3).

Hea

t Flo

w

Figure

Tm

Endoderm

Exoderm

Temperature

5 1.3. Typical DSC curve (thermogram) of s

Tm

Tonset

tarch

Page 22: Calorimetric and Microbiological Evaluation of Bacteria

The value of the Q is assessed to calculate heat capacity (Cp) by an equation;

mrQC p ×

−=

where r is heating rate (oC/sec) and m is weight of the sample (g). The area under the

peak is directly proportional to the heat or enthalpic change (∆H = Cp ∆T) and its

direction indicates whether the thermal event on a sample is endothermic (denaturation or

melting) or exothermic (crystallization or aggregation). The characteristic peak of such a

plot provides the transition temperature (thermal stability, Tm) at which the thermal event

is half-complete (Fig. 1.3).

Bacterial thermal analysis by differential scanning calorimetry Bacteria are

composed of cellular components such as cell envelope, ribosomes, nucleic acids, and

proteins. Since the basic structures of those macromolecular components are

biopolymers, the components in whole cells may go through conformational transitions

upon exposure to heating by DSC. In DSC, the transitions recorded as endothermic (heat

absorption) or exothermic (heat loss) peaks in the thermogram.

The first application of DSC on bacterial thermal analysis was the study on the

physical properties of biomembranes. Steim et al. (1969) studied the physical properties

of lipids in cell membranes of Mycoplasma laidlwii using DSC of their whole cells, cell

membranes, and extracted lipids. They reported that both isolated cell membranes and

extracted membrane lipids showed an endothermic transition around 40oC in DSC

6

Page 23: Calorimetric and Microbiological Evaluation of Bacteria

thermograms. They claimed that the lipids in the membranes were as much stable as

extracted lipids because the enthalpies of lipid melting in both samples were not different.

However, they could not obtain distinguishable peaks in the DSC thermogram from pure

whole cell (Bach and Chapman, 1980).

The first successful DSC on whole cell was the study on heat inactivation and

spontaneous germination of bacterial spores. Maeda et al. (1974) observed that

germinated Bacillus megaterium spores had endothermic peaks at about 100oC and 130oC

in their DSC thermogram. For vegetative cells, Verrips and Kwast (1977) reported eight

unidentified endothermic peaks for whole cell DSC thermograms of Citrobacter freundii.

Maeda et al. (1974) believed that the first two peaks (at 50oC and 55oC) were related with

the thermal death because of the loss of viability occurring at the same range of

temperature.

Using DSC, Miles et al. (1986) studied heat resistance of bacterial species of nine

genera including Gram-negatives, Gram-positives and spore formers. Cells on agar

surfaces were collected into DSC pans and heated at rate of 10oC min-1 from 10 to 120oC.

Vegetative bacteria showed distinguished major peaks in the regions of 68-73, 77-84, 89-

99 and 105-110oC in DSC thermogram. The onset temperature of first peak, the largest

among the peaks, was the lowest (36-49oC) in psychrophilic bacteria (36oC for Vibrio

marinus, 42oC for Brochothrix thermoshacta, 47oC for Hafnia alvei and 49oC for

Pseudomonas fragi) and highest in thermophilic bacteria (67oC for Bacillus

stearothermophilus). Mesophilic bacteria such as Streptococcus faecalis and Escherichia

coli showed the onset temperature at range of 50 to 52oC. Therefore, Miles et al. (1986)

7

Page 24: Calorimetric and Microbiological Evaluation of Bacteria

claimed that the onset temperature of the first thermal denaturation is strongly related

with thermal tolerance of studied organisms. Later studies showed that the onset

temperature is correlated with the maximum growth temperatures of bacteria (Lepock et

al., 1990; Mackey et al., 1993; Mohacsi-Farkas et al., 1994; Teixeira et al., 1997). Miles

et al. (1986) also identified that the peak in the range of 89-99oC in each organism was

associated with the melting of DNA because it was reversible and its transition

temperature (Tm) was within the expected range of the melting of bacterial DNA.

Using the melting temperatures of the putative DNA peaks from the whole cell

DSC of 58 bacteria strains, Mackey et al. (1988) observed the relationship between the

transition temperatures of the DNA peaks in the whole cell thermograms and the

literature values of mole fraction of guanine + cytosine (G + C) base pairs in isolated

DNA (Table 1.1). Table 1.1 shows that the transition temperature of the putative DNA

peak was higher in bacterial strains containing greater content of guanine + cytosine (G +

C) base pairs in their DNA. It has been known that the thermal stability for bacterial

DNA increases with the G + C content due to more extensive hydrogen bonding (Jay,

1996). Using the transition temperature value of the DNA peaks in the whole cell DSC,

Mackey and coworkers (1988) also developed the model further to predict the mole

fraction of G + C in bacterial DNA (XGC) as Tm = (41.0 x XGC) + 73.8. The value of XGC

in the equation for each organism closely agreed with the XGC values based on the

spectroscopically determined transition temperatures of isolated bacterial DNA (De Ley,

1970).

8

Page 25: Calorimetric and Microbiological Evaluation of Bacteria

Bacterial Strain G + C content (mol%) Tm (oC)

Bacillus cereus ATCC 14579 35.7 88.6

Bacillus macerans ATCC 8244 52.2 95.8

Campylobacter coli NCTC 11366 32.3 86.2

Campylobacter jejuni NCTC 11351 31.6 83.8

Escherichia coli KL 16 51.6 94.3

Enterobacter agglomerans NCTC 9381 56.0 96.0

Lactobacillus bulgaricus NCDO 1489 50.0 91.7

Lactobacillus cremoris NCDO 543 42.0 91.0

Pseudomonas aeruginosa NCTC 10332 66.4 100.3

Table 1.1. The G + C content and the temperature recorded for peak associated with the melting of putative DNA from whole cell DSC (adapted from Mackey et al., 1988).

Mackey et al. (1991) investigated the identification of the origins of individual

transitions on the thermogram of E. coli NCTC 8164. For the whole cell DSC, the

pellets of the organism were heated from 5 to 130oC at 10oC min-1 heating rate.

Individual peaks observed in thermograms of whole cells of E. coli were assigned to cell

components by comparing the transition temperatures of isolated cell components with

corresponding transitions in whole cells (Table 1.2). Among the main thermogram peaks

obtained in the E. coli whole cells, the most prominent peaks (temperature range of

60~80oC) being associated with ribosome denaturation. The ribosomal denaturation by

the DSC was associated with the 30S and 50S ribosomal subunits in increasing order of

thermal stability.

9

Page 26: Calorimetric and Microbiological Evaluation of Bacteria

Cell component Mean transition temperature (Tm, oC)

30S ribosomal subunit 62-64

50S and 70S ribosomal subunits 69-80

Transfer RNA (tRNA) 75-76

DNA and cell wall 95, 105

Cell envelope 118-125

Table 1.2. Major transition temperatures in the thermograms of whole cells of E. coli NTCT 8164 and corresponding cell components (Mackey et al., 1991). Since the ribosome-associated components of the DSC thermogram were identified,

many DSC studies on bacteria have focused on the relationship between ribosome

stability and thermal resistance (Anderson et al., 1991; Mackey et al., 1993; Mohacsi-

Farkas et al., 1994; Teixeira et al., 1997). In those studies, a DSC instrument was used to

apply heat to the bacteria for determining the reduction of cell numbers by plate count

method as well as for generating thermogram. The studies showed strong correlations

between the temperature of ribosome-associated DSC thermogram events and the

temperature at which thermal death of bacteria occurs.

In thermal study on Listeria monocytogenes by Anderson et al. (1991), cell

suspensions in pans were heated to 60oC in the DSC with different holding times and

10

Page 27: Calorimetric and Microbiological Evaluation of Bacteria

removed, then survivors were counted by plate counting method. Anderson et al. (1991)

observed that the first major peak (ribosomal subunits) disappeared and viability

decreased by two orders of magnitude after 60oC for 5 min.

Mohacsi-Farkas et al. (1994) observed that the temperatures at which the loss of cell

viability started (55oC for E. coli, 52oC for Lactobacillus plantarum, and 58oC for L.

monocytogenes) matched with the transition temperatures of the first irreversible

endothermic peaks (ribosomal subunits). Similar results were reported for vegetative

cells of Bacillus stearothermophilus (Mackey et al., 1993) and Lactobacillus bulgaricus

(Teixeira et al., 1997). Those results suggest that ribosome damage is an important factor

in causing the loss of bacterial viability during heat treatment.

Thermal Processing Effect on Microorganisms

The effect on cell components Bacterial cells contain several targets for the action of

heat (Anderson et al., 1991; Russell, 2003). Therefore, it can be anticipated that the

extent of heat effect is related to the stability of macromolecules in cell wall, membrane

lipids, ribosomes, nucleic acids, and proteins.

Peptidoglycan represents the main component (50% of the weight) of the cell walls

of Gram-positive bacteria (Murray et al., 1965). The net-shaped structure of the

peptidoglycan layer may not be seriously affected by a mild heat treatment because it

contains polysaccharide chains cross-linked by tight peptide bridges to maintain the

11

Page 28: Calorimetric and Microbiological Evaluation of Bacteria

stability of the bacterial shape (Hammond et al., 1984; Novak and Juneja, 2001).

Peptidoglycan also plays a major role in heat resistance of spores of Gram-positive

bacteria, such as Bacillus subtilis and Clostridium botulinum (Ellar, 1978; Popham et al.,

1996). However, the effect of peptidoglycan on the thermal stability of intracellular

structures in vegetative cells has not been clearly elucidated. The main structure of the

Gram-negative cell wall is the outer membrane. Lipopolysaccharide (LPS) is the

predominant component (~40% weight) of the outer membrane and the remainder is

made up from phospholipids and proteins. LPS is held in the outer membrane by

relatively weak cohesive forces (ionic and hydrophobic interactions) and can be

dissociated from the cell surface with mild heat (Tsuchido et al., 1985). LPS consisting

of lipid A, core, and O antigen, is heat stable because those three structures are covalently

linked to each other (Wright and Tipper, 1979). Thermal studies of isolated outer

membrane components showed that the denaturation of LPS required much higher

temperature (>120oC) than the cell death temperature, while that for outer membrane

protein was around 70oC (Rodriguez-Torres et al., 1993; Phale et al., 1998).

It has been suggested that the primary cause of cellular heat injury is the damage of

membrane lipoprotein complexes or proteins that confer integrity of the cytoplasmic cell

membrane (Bowler et al., 1973). The damage leads to the dissipation of the

transmembrane H+ gradient and a decrease in intracellular pH (Weitzel et al., 1987; Piper

et al., 1997). Heat-induced damage of the membrane can be detected by measuring the

amount leakage of intracellular substances such as ions, nucleotides, and amino acids

(Russel and Harries, 1967). However, there was no relationship between the rate of the

12

Page 29: Calorimetric and Microbiological Evaluation of Bacteria

increase in leakage amounts and the loss of viability during mild heat (<58oC) treatment

in several studies, indicating that cytoplasmic membrane damage is not a major factor in

cell inactivation (Allwood and Russell, 1967; Russell and Harries, 1968). Mackey et al.

(1991) reported that the thermal transition of isolated membrane lipid of E. coli is the

range of 30 ~ 40oC in DSC. However, the membrane lipid transition was hardly detected

as a peak when whole bacterial cells were used for DSC (Anderson et al., 1991; Mackey

et al., 1991; Teixeira et al., 1997; Mohacsi-Farkas et al., 1999). The melting temperature

of the membrane lipids in thermophilic bacteria is proposed to be higher than that of

mesophilic bacteria since their membranes are rich in saturated fatty acids (Russell,

2003).

Ribosomes are large complexes of proteins and three rRNA (ribosomal ribonucleic

acid; 30S, 50S and 70S) subunits in prokaryotes. Ribosomes comprise a major part of the

bacterial cell, constituting 25% of the total cell mass. Approximately 65% of E. coli

ribosome consists of rRNA, with the rest consisting of ribosomal proteins. The

protein/RNA, RNA/RNA, and protein/protein interactions in the ribosomes stabilize

tertiary structures. These interactions in the ribosomal subunits can also be affected by

heat stress (Bonincontro et al., 1998). Heat inactivation of microorganisms was proposed

to be related to denaturation of ribosomal subunits, mainly 30S and 50S (Rosenthal and

Iandolo, 1970; Hurst, 1984). The loss of Mg2+, which stabilizes ribosomal subunits,

from membrane damage is the primary reason of thermal degradation of ribosome in

microorganisms (Hurst and Hughes, 1978; Hurst, 1984). After the loss of Mg, divalent

cation-inhibited nucleases become activated and catalyze the degradation of 30S and 50S

13

Page 30: Calorimetric and Microbiological Evaluation of Bacteria

RNA by cleavage of phosphodiester bonds, leading to irreversible ribosomal unfolding

(Datta and Niyogi, 1976). Mackey et al. (1991) isolated subunits (30S and 50S) of E. coli

ribosomes and compared them with whole cells in DSC thermograms. The denaturation

of the ribosomal subunits occurred at 50~80oC range in both thermograms. The 50S and

70S subunits, which have more rigid structures, were more stable than 30S subunits

during heat treatments. In a recent study on isolated E. coli ribosome, Bonincontro et al.

(1998) reported the DSC profile of thermal degradation of 50S was identical to that of

70S.

Heat treatment affects both double stranded DNA (dsDNA) and single stranded

DNA (ssDNA). The dsDNA damage is induced by direct heat which breaks the

hydrogen bonds between base pairs of DNA while ssDNA damage is mainly due to the

cleavage activity from endonucleases after heating (Russell, 2003). It has been known

that the denaturation temperatures of DNA are strongly related to base composition (Pace

and Campbell, 1967; Mackey et al., 1988). There is an important contribution from

intracellular cation concentration, shifting DNA denaturation to higher temperatures

because negatively charged phosphate (PO43-) backbones of dsDNA interact with cations

(Kumar, 1995). It has been suggested that the heat denaturation of DNA might not be a

major factor of vegetative cells or spore death because the event is only partially

irreversible and requires higher temperature (85-100oC) than bacterial death (Verrips et

al., 1977; Mackey et al., 1988; Mohacsi-Farkas et al., 1999). The DNA melting

temperature increases due to stabilizing interactions with other intracellular molecules

14

Page 31: Calorimetric and Microbiological Evaluation of Bacteria

such as cationic proteins and polyamines (Worcel and Burgi, 1972; Flink and Pettijohn,

1975).

Unlike DNA/DNA interactions, RNA usually exists as a single chain without a

complementary strand. RNA can hold back on itself to form double helical regions

(Saenger, 1984). The denaturation temperature of RNA is also strongly related with the

ratio of base pairs. Because of rotational freedom in the backbone of its non-base paired

regions, RNA can hold into tertiary structures involving irregular base pairing

(Gesteland, 1993). Among RNA structures, the thermal stability for tRNA (~79oC),

which has more complex tertiary structure, is higher than that for rRNA (~73oC) (Mackey

et al., 1991).

Thermal process leads denaturation and coagulation of bacteria proteins (Russell,

2003). Many of the thermal denaturations of proteins in microorganisms are irreversible

due to following aggregation and alterations of amino acid residues (Kurganov et al.,

1997). Thermal property of proteins in the cells largely depends on the presence of water

attached within groups or at surface of protein molecules having free charges and water

in the tertiary structure of protein (Earnshaw et al., 1995). It has been suggested that the

thermal resistance of cells is higher when the presence of the protein contacted water

level is low because more dipoles of the protein interact each other to stabilize the protein

complex (Warth, 1985). The environmental pH is also an important factor for the

thermal properties of cell proteins. The heat stability of proteins decreases if the pH

condition of heating environment is far below or above the isoelectric points of proteins

(Condon et al., 1992). In a recent spectrophotometric study on thermal stability of

15

Page 32: Calorimetric and Microbiological Evaluation of Bacteria

bacterial protein, Boer and Koivula (2003) reported that the thermal stability (Tm, ~65oC)

of purified Trichoderma reesei enzyme (cellobiohydrolase), which has optimal thermal

stability at pH 5, decreased by >10oC when pH of heating environment was adjusted with

3.5 or 8.0.

Many proteins in bacterial cells have been proposed to be stable at higher

temperatures than those known to support viability of microbial cells due to hydrophobic

interaction and binding with other components (Daniel and Cowan, 2000). It is

hypothesized that thermally or non-thermally induced complete denaturation of protein

molecules may not lead to cell death if the corresponding gene is undamaged and the

energy and building bocks are supplied to reform that proteins. However, the irreversible

denaturation of proteins such as RNA polymerase, ribosomal proteins and some enzymes

which are involved in protein synthesis, should cause loss of cell viability (Davis, 1990).

Mechanism of cell death The death of microorganisms during thermal processing has

been known as a two-step process; reversible damage occurs initially and is increasingly

converted into lethal events that result in cell death (Jung, 1986; Bowler and Manning,

1994). Figure 1.4 shows a possible sequence of lethal events during thermal death of

microorganisms and indicates that the plasma membrane is the primary target. However,

all of macromolecular components such as cell wall, enzymes and proteins, nucleic acids

can be directly affected to some degree by high temperatures. Since those components

are important structures for cell viability, the study on the irreversible heat damages of

the components has been highly recommended (Earnshaw et al., 1995; Russell, 2003).

16

Page 33: Calorimetric and Microbiological Evaluation of Bacteria

Thermal perturbation of plasma membrane

Inactivation of membrane proteins

Decreased order of lipid layer

Leakage of mono- and divalent ions

Failure of ion pump and nutrient transport

Loss of ion gradients

Calcium overload

Loss of coupling and inactivation of receptors

Disintegration of cell membrane

Activation of phospholipases, proteases and protein kinases

Breakdown of metabolic control and loss of cellular homeostasis

Cell death

Figure 1.4. A scheme for the possible sequence of events leading to the death of microorganism from heating (Bowler and Manning, 1994).

17

Page 34: Calorimetric and Microbiological Evaluation of Bacteria

Effect of chemical agents on microorganisms

Hurdle technology Homeostasis is an important adaptation mechanism of

microorganisms that maintains the stability of internal environment of them against

changes in living external environment. Disturbing the homeostatic mechanisms has

been regarded as main goal of hurdle technology in which food treatments are combined

to produce shelf-stable, minimally processed foods that have maintained the nutritional

qualities with extended shelf-life (Leistner and Gorris, 1995; Leistner, 2000). It has been

believed that effectiveness of the hurdle technology on microbial inactivation can be

improved when cells are exposed to heat after injured with chemicals (Karatzas et al.,

2000; Leistner, 2000). The physiological conditions of bacterial cells are known to be

affected by ethanol (Salton, 1963; Ingram, 1986), NaCl (Gutierrez et al., 1995; Poirier et

al., 1998) and acids (Abee and Wouters, 1999; Brul and Coote, 1999), which are utilized

during food processing.

Effect of ethanol The cellular membrane is a semipermeable barrier for survival of

bacteria; however, it is also the primary target of ethanol damage (Ingram, 1986).

Replacement of water molecules with ethanol can disrupt the hydrophobic core by

weakening hydrophobic interactions which maintain membrane integrity (Ingram, 1990).

Leakage of intracellular components through damaged cell membranes and end-product

inhibition of enzymes in glycolysis are considered as the basic mechanisms of ethanol

inhibition in microorganisms (Salton, 1963; Ingram, 1986). The presence of high

18

Page 35: Calorimetric and Microbiological Evaluation of Bacteria

concentration of ethanol in cells may change the dielectric properties of the intracellular

components; however, the result of these ethanol-induced changes on the cellular

structures and cell viability are not clearly defined. Most bacteria demonstrate a dose-

dependent inhibition of growth over a range of 1 to 10% (v/v) ethanol, and few organisms

such as ethanol producers and lactobacilli are viable at concentration above 10% (Ingram

and Buttke, 1984). The growth of E. coli strain is inhibited by ~5% of ethanol

concentration (Ingram, 1986).

Effect of NaCl Most food-borne bacteria are inhibited by ≥5% (w/v) of NaCl due to

plasmolysis in which water is drawn out of the cell and into the outside cell (Jay, 1996).

A rapid decrease in cell volume due to the loss of water from the cell was evidence for

the inactivation (Munns et al., 1983). Recent studies on E. coli indicated that cell death

observed in the presence of NaCl (>5% concentration, w/v) may be related to the toxic

effect of Na+ in the cell rather than a decrease in cell volume because solutes can be

accumulated in cells to maintain the internal osmotic pressure (Gutierrez et al., 1995;

Poirier et al., 1998; Shadbolt et al., 1999). E. coli is known to accumulate “compatible

solutes” such as betaine, trehalose, glycerol, sucrose, proline, mannitol, sorbitol and small

peptides, which can increase internal osmolarity against a hyperosmotic shock without

interfering functions of cellular enzymes. Poirier et al. (1998) hypothesized that the cell

volume decrease can also be limited due to counterbalance the external osmotic pressure

by increased Na+ and Cl- concentration inside of the cell. However, specific

macromolecular targets of the ionic stress of Na+ have not been identified.

19

Page 36: Calorimetric and Microbiological Evaluation of Bacteria

Effect of Acids Because neutral pH is optimum for the growth and survival of most

microorganisms, low pH values have long been considered as important factors to

inactivate bacteria (Jay, 1996). Bacteria can regulate their intracellular pH (pHi) at a

value close to neutrality against low pH environment using the pH homeostasis system

(Hill et al., 1995). Under high acidic conditions (<pH 3), however, the penetration of H+

ions across the cell membrane is faster than the removal of the ions by the pH

homeostasis system. As a result, the cell ceases essential biochemical activities (Bearson

et al., 1998). When the pH of food or solution is lower than pKa of a weak organic acid,

undissociated organic acids readily pass through cellular membrane and enter the cell.

Once weak organic acids reach the inside of bacterial cells, the cells are inactivated by

the release of both charged anions and protons from organic molecule (Brul and Coote,

1999). The charged form of the organic acid intensively lowers intracellular pH by

interfering with metabolic and anabolic processes of the cell (Abee and Wouters, 1999).

Among weak organic acids, acetic acid (CH3COOH) and lactic acid (CH3CHOHCOOH)

are more effective in the inactivation of bacteria because they have relatively small

molecular weight and easily diffuse into the cell (Hisao & Siebert, 1999).

High Hydrostatic Pressure (HHP)

HHP technology for food preservation Similar to thermal treatment, the

conformational and phase transitions of macromolecules in sample are affected in HHP

20

Page 37: Calorimetric and Microbiological Evaluation of Bacteria

treatment because of volume changes (Cheftel and Culioli, 1997). High pressure

treatment follows Le Chatalier’s principle in which a process associated with a reduction

in volume is favored by an increase in pressure and vice versa under equilibrium

(Earnshaw et al., 1995, Farkas and Hoover, 2000). HHP treatment also follows the

isostatic principle, in which pressure is transmitted in a uniform manner throughout the

sample (Cheftel, 1995). Therefore, the time required for reaching the pressure to all

components is instant and independent of the volume and shape of sample. This

principle represents a significant benefit of HHP, as compare to thermal treatment, where

a thermal gradient must be established in sample (Balny and Masson, 1993). Another

principle in HHP is that at constant temperature, elevated pressure levels increase the

degree of ordering of the molecules of a substance in sample. Thus, the melting point of

solids increases with the pressure (Heremans, 1992). Because of above advantages based

on its operation principles, high pressure processing technology of food is being

investigated as an alternative to thermal processing. This novel technology can be

potential to produce microbiologically safe food with enhanced quality, flavor and

textural properties in comparison to thermal processing (Mertens and Deplace, 1993;

Roberts and Hoover, 1996).

As early as 1889 Hite showed that pressures of 450 MPa or greater could eliminate

spoilage microorganisms and improve the storage quality of milk. Despite the early

imploring of the technology, the two main barriers that prevented rapid commercial

application of HHP were the lack of knowledge about the significant advantage of the

process over other existing preservation methods and difficulty in high pressure vessel

21

Page 38: Calorimetric and Microbiological Evaluation of Bacteria

manufacture and operation (Earnshaw et al., 1995); however, both of these problem areas

have been resolved. The recent research in HHP technology has intensively been

studying in two areas. Researchers have been focusing on the development of kinetic

data and advanced knowledge of the mechanism of HHP effect on food systems.

Engineering fields have been trying to solve the problems on the temperature distribution

within pressure vessels and compressibility differences within complex food systems

(Knorr, 1993).

Effects of HHP on microorganisms The effectiveness of hydrostatic pasteurization has

been reported for several foodborne pathogens, namely Salmonella spp., Escherichia coli

O157:H7, Camplylobacter jejuni, Vibrio parahaemolyticus, Listeria monocytogenes and

Staphylococcus aureus (Metrick et al., 1989; Shigehisa et al., 1991; Styles et al., 1991;

Patterson et al., 1995; Gervilla et al., 1997; Kalchayanand et al., 1998; Alpas et al., 1999).

Studies revealed that cell viability decreases with increasing pressure and time (Metrick et

al., 1989; Robey et al., 2001). The effect of HHP on bacterial cells is enhanced when the

cells are pressurized at ≥35oC (Kalchayanand et al., 1998; Alpas et al., 2000).

Resistance to high pressure varies among strains of the same species. Various strains

on foodborne pathogens were observed to be relatively resistant to pressure in

comparison to other strains (Styles et al., 1991; Patterson et al., 1995; Hauben et al.,

1997; Alpas et al., 1999; Benito et al., 1999). The destruction of ≥ 8 log cycles of some

strains of Escherichia coli O157:H7 and Staphylococcus aureus in phosphate buffer was

achieved by pressurization for 15 min at 20oC at 700 MPa (Patterson et al, 1995).

22

Page 39: Calorimetric and Microbiological Evaluation of Bacteria

However, strains of pressure-sensitive E. coli were reported to develop resistance to high

pressure by adaptation or mutation to survive at 800 MPa in the buffer (Hauben et al., 1997).

The composition of the pressurizing menstruum has a great impact on the effect of

HHP on microorganisms. Viability loss is lower in a food system than in phosphate buffer

(Metrick et al., 1989; Patterson et al, 1995). Proteins in culture medium or food were

reported to protect bacterial cells against pressure (Simpson and Gilmore, 1997; Park et al.,

2001). In addition, enriched media provide nutrients such as amino acids and vitamins to

pressure-injured cells to recover.

Cell inactivation by high pressure is strongly related with cell wall type and cellular

morphology. Gram-negative bacteria and rod-shaped cells showed more sensitivity to

pressure treatment than gram-positive bacteria and cocci-shaped cells (Styles et. al, 1991;

Cheftel, 1995; Ludwig and Schreck, 1997; Kalchayanand et al., 1998).

Studies have revealed that cell viability decreases with increasing pressure and time,

suggesting critical cellular activities have been irreversibly damaged (Hoover et al., 1989;

Metrick et al., 1989; Cheftel, 1995). The primary target of bacterial cells in HHP

treatment has been proposed to be the cytoplasmic membrane (Kalchayanand et al., 1998;

Farkas and Hoover, 2000). It was observed that bacterial cell viability is related with the

loss of the membrane integrity (Shigehisa et al., 1991) and the failure of active transport

system (Cheftel, 1995). The denaturation of enzymes, which includes the alteration of

molecular structures and change in active sites, has also been considered as a major factor

in pressure-induced cell injury (Suzuky and Suzuky, 1962; Mackey et al., 1995). Studies

using electron microscopy reported that the separation between the cell wall and

23

Page 40: Calorimetric and Microbiological Evaluation of Bacteria

cytoplasmic membrane, and destruction of ribosomes were shown in dead cells of

microorganisms after HHP treatment (Mackey et al. 1994; Isaacs et al., 1995). However,

there is little information in literature about the nature of irreversible changes occurring

in cells, which leads to cell death as a result of HHP treatment.

HHP in combination with other processing technologies High hydrostatic pressure

(HHP) has been shown to inactivate spoilage and pathogenic bacteria without altering the

food quality and has been recognized as an alternative to thermal processing (Roberts and

Hoover, 1996; Mertens and Deplace, 1993; Knorr, 1993). However, excessive high

pressure necessary to obtain a desirable reduction of pathogens is expected to alter the

conformational structure of high molecular weight compounds such as starch and protein.

As a result, normal texture and color of many foods can be adversely changed at high

pressure. In fish meat, lipid oxidation occurred since the peroxide value of fish oil increased

with increasing pressure (Ohshima et al., 1992). HHP treated tomato juice gave rancid

flavor due to the oxidation of free fatty acids (Porretta et al., 1995). The lightness of skim

milk color was significantly decreased after HHP at 600 MPa due to disintegration of casein

micelles into small fragments (Mussa and Ramaswamy, 1997). The HHP treated cheeses

had higher moisture, salt contents than raw or pasteurized cheeses (Trujillo et al., 1999). In

addition, very high pressures the process may not be economical for commercial use due

to the high cost of equipment and increased metal fatigue which leads to high

maintenance costs (Hoover et al., 1989; Mertens and Deplace, 1993). Microbiological

problems to be addressed include sublethal damage and following recovery in HHP

24

Page 41: Calorimetric and Microbiological Evaluation of Bacteria

inactivation of foodborne bacteria (Earnshaw, 1995; Patterson et al., 1995). The

problems can be significant issue to food industry when HHP is used as a single

preservation against food pathogens.

There have been studies employing the concept of “Hurdle technology” in which

HHP technique is combined with one or more suitable antimicrobial agent to produce

shelf-stable, minimally processed foods that have maintained the nutritional qualities

with extended shelf-life (Kalchayanand et al., 1998; Massachalck et al., 2001). Hurdle

technology has been applied to inactivate pathogenic bacteria by combining HHP with

CO2 (Hass et al., 1989), irradiation (Crawford et al., 1996; Paul et al., 1997), heat

(Patterson and Kilpatrick, 1998; Benito et al., 1999; Alpas et al., 2000), low pH (Alpas et

al., 2000), or lysozyme (Popper and Knorr, 1990; Masschalck et al., 2001).

HHP in combination with bacteriocins Some of antimicrobial peptides produced by

lactic acid bacteria, termed bacteriocins, have been used in foods as safe and natural

preservatives. In recent literature, there are studies employing combination of HHP

technique with bacteriocins (Kalchayanand et al., 1998; Ponce et al., 1998; Yuste et al.,

1998; Garcia-Graells et al., 1999; Masschalck et al., 2000, 2001). Kalchayanand and co-

workers (1998) reported a study based on application of combined HHP and bacteriocin

(mixture of pediocin and nisin) treatment on model systems involving various strains of

four pathogenic bacteria. The results in the Table 1.3 clearly indicate that bacteriocins

can provide an additional 1 to 5 log cycle reductions in bacterial populations in buffer

medium.

25

Page 42: Calorimetric and Microbiological Evaluation of Bacteria

Nisin is an antibacterial peptide produced by certain strains of Lactococcus lactis.

Nisin was approved as a food preservative in over 50 countries including European

Economic Community (EEC) and by FDA in US (Delves-Broughton, 1990; Yuste et al.,

1998). Nisin is effective against Gram-positive bacteria but shows very little activity

against Gram-negative bacteria which have nisin-impermeable barrier (outer membrane)

in their cell envelope (Delves-Broughton, 1990; Massachalck et al., 2001). Recent

studies showed that damaged cells of Gram-negative bacteria may be sensitive to nisin

(Kordel and Sahl, 1986; Kalchayanand et al., 1992; Masschalck et al., 2000). Moderate

HHP treatment has been reported to cause a number of morphological changes in

bacterial cells including cell lengthening, separation of cell membrane from the cell wall,

and pore formation in the cell wall (Mackey et al., 1994; Cheftel, 1995). Antimicrobial

molecules such as nisin may penetrate the damaged outer membrane of Gram-negative

bacteria. In a combined HPP and nisin treatment, the primary target in bacterial cell was

proposed to be the cytoplasmic membrane (Kalchayanand et al., 1998). However, the

mechanism of the inactivation of bacteria either HHP or nisin is still not clearly known.

26

Page 43: Calorimetric and Microbiological Evaluation of Bacteria

Log10 cfu/ml of survivors

Pathogens Unpressurized Pressurized Pressurized+bacteriocin

Staphylococcus aureus 582 8.8 4.0 <1.0

Listeria monocytogenes Scott A 8.9 5.1 <1.0

Salmonella typhimurium ATCC 14028 8.9 4.1 1.0

Escherichia coli O157:H7,932 7.9 3.4 2.1

Table 1.3. Survivors of four pathogens by pressurizing at 345 MPa for 10 min at 25oC in the absence and presence of bacteriocins (Kalchayanand et al., 1998).

27

Page 44: Calorimetric and Microbiological Evaluation of Bacteria

CHAPTER 2

EVALUATION BY DIFFERENTIAL SCANNING CALORIMETRY OF THE HEAT

INACTIVATION OF ESCHEICHIA COLI AND LACTOBACILLUS PLANTARUM*

ABSTRACT

Differential scanning calorimetry (DSC) was used to evaluate the thermal stability and

reversibility after heat treatment of transitions associated with various cellular

components of Escherichia coli and Lactobacillus plantarum. The reversibility and the

change in the thermal stability of individual transitions were evaluated by a second

temperature scan after pre-heating in the DSC to various temperatures between 40 and

130oC. Viability of bacteria subsequent to a heat treatment between 55 and 70oC in the

DSC was determined by both plate count and calorimetric data. The fractional viability

based on calorimetric data and plate count data showed a linear relationship. Viability

loss and the irreversible change in DSC thermograms of pre-treated whole cells were

highly correlated between 55 and 70oC. Comparison of DSC scans for isolated

ribosomes showed that the thermal stability of ribosomes from E. coli is greater than the

28

Page 45: Calorimetric and Microbiological Evaluation of Bacteria

thermal stability of L. plantarum ribosomes, consistent with the greater thermal tolerance

of E. coli observed from viability loss and DSC scans of whole cells.

Key Words: heat treatment, differential scanning calorimetry, ribosome denaturation, E.

coli, L. plantarum

* Adapted from Applied and Environmental Microbiology, 68:5379-5386 (2002).

29

Page 46: Calorimetric and Microbiological Evaluation of Bacteria

INTRODUCTION

The main goal of thermal processing is to inactivate the spoilage and pathogenic

microorganisms to produce a safe product with enhanced shelf life. An understanding of

the mechanism of microbial inactivation by heat is potentially useful for optimizing heat

treatments in order to eliminate foodborne disease and spoilage risk associated with

common and emerging strains while avoiding over processing of the food material.

Thermal inactivation of microorganisms is associated with irreversible denaturation of

membranes, ribosomes, and nucleic acids. However, the patterns of macromolecular

changes that induce the cell death of microorganisms during heat treatment are still not

clearly known.

Differential scanning calorimetry (DSC) is a thermal analysis technique that detects,

monitors, and characterizes thermally-induced conformational transitions and phase

transitions as a function of temperature. A number of overlapping transitions with a net

endothermic effect are observed when microorganisms are heated (Miles et al., 1986;

Anderson et al., 1991; Mackey et al., 1991; Mohacsi-Farkas et al., 1999; Kaletunç, 2001).

The observed transition peaks correspond to the denaturation of cellular components.

Mackey et al. (1991) investigated the origins of apparent individual transitions on the

thermogram of E. coli. Individual peaks observed in thermograms of whole cells of E.

coli were assigned to cell components by comparing the transition temperatures of

isolated cell components with corresponding transitions in whole cells (Fig. 2.1). It is

believed that a strong relationship exists between thermal death of bacteria and the first

30

Page 47: Calorimetric and Microbiological Evaluation of Bacteria

major peak in DSC thermograms (temperature range of 60~80oC) which is attributed to

ribosomal melting (Mackey et al., 1993; Teixeira et al., 1997). Several investigators have

shown correlations between the stability of ribosomes and cell viability for

Staphylococcus aureus (Allwood and Russel, 1967), Listeria monocytogenes (Stephens

and Jones, 1993), and Salmonella enterica serovar Typhimurium (Tolker-Nielsen and

Molin, 1996). Furthermore, a recent DSC investigation of pressure-treated E. coli NCTC

8164 demonstrated that lethality of cells and ribosome damage are closely related (Niven

et al., 1999). Irreversible denaturation of cellular DNA requires temperatures well above

the temperature of cell inactivation (Mackey et al., 1991). At temperatures that cause

ribosome denaturation, the DNA transition is reversible (Mohacsi-Farkas et al., 1999).

Previous DSC investigations of microorganisms employed scans to high

temperatures (at or above 100oC) resulting in inactivation of the microorganisms. Most

rescans did not display any peaks except for an endothermic transition attributable to

DNA (Miles et al., 1986; Anderson et al., 1991; Mackey et al., 1991; Mohacsi-Farkas et

al., 1999). Although DSC thermograms were compared to viability studies, no studies

examined the relationship between thermal stability differences in whole cells and in

isolated ribosomes or correlations between viability measures based on plate count and

calorimetric data. The objectives of this study include: i) comparison using calorimetry

of the thermal stability of two selected microorganisms, E. coli and L. plantarum, in

relation to the thermal stabilities of their ribosomes; ii) investigation of the reversibility

of individual transitions associated with various components of whole cells of E. coli and

L. plantarum, and iii) determination and comparison of the temperature dependence of

31

Page 48: Calorimetric and Microbiological Evaluation of Bacteria

cell viability for a linearly increasing temperature protocol from plate counts and

calorimetric data.

20 40 60 80 100 120 140

a1

a2

a3 b

c d

a1, a2, a3 --- Ribosome subunits b --- DNA

c --- DNA and cell wall d --- G- bacterial cell wall

Probable components of peaks

Heat Flow 0.2 mW

Temperature (oC)

Figure 2.1. DSC thermogram of whole cells of E. coli ATCC 14948

32

Page 49: Calorimetric and Microbiological Evaluation of Bacteria

MATERIALS AND METHODS

Source and preparation of organisms

E. coli ATCC 14948 and L. plantarum ATCC 10241 were obtained from the Culture

Collection, Department of Microbiology at the Ohio State University. A loopful of each

organism was revived in 10 ml Trypticase soy broth (Difco laboratories, Detroit, MI)

supplemented with 0.3 % (w/w) yeast extract for E. coli or MRS broth (Difco) for L.

plantarum and incubated at 37oC for 18 hours. Each culture was stored frozen (-80oC) in

30 % (v/v) sterile glycerol. A loopful of each stock culture was transferred to 10 ml

Trypticase soy or MRS broth and incubated 10 hrs at 37oC before use.

Each culture was inoculated (1 % v/v) into a broth containing Trypticase soy or MRS

broth. Cultures were incubated at 37oC. The growth phase was determined by measuring

absorbance (A640), using a Beckman Du-50 spectrophotometer, and matching appropriate

viable counts from a standard growth curve. The cells were grown to late exponential

growth phase, as determined from the growth curve. The final concentration of cells in

the medium was 1.3 ± 0.1 x 109 cfu ml-1 for E. coli and 9.0 ± 0.1 x 108 cfu ml-1 for L.

plantarum. Cells in the broth were harvested by centrifugation (Beckman J2-21

centrifuge) at 10 000 g for 10 min at 4oC. The supernatant was discarded and the pellets

were washed with sterile distilled water and centrifuged for a second time before

transferring into DSC crucibles.

33

Page 50: Calorimetric and Microbiological Evaluation of Bacteria

Calorimetry of whole cells

Pellets of whole cells were transferred into the empty sample crucible and were

weighed (56 ± 0.3 mg wet weight). The dry material content of the pellets was

determined by freeze drying (Freezone 4.5, Freeze dry system, Model 77510, Labconco,

Missouri) as 19 ± 0.3 % for E. coli and 20 ± 0.5 % for L. plantarum on a wet basis. The

standard deviations were calculated based on twelve freeze dried pellets for each

bacterium.

A differential scanning calorimeter (DSC 111, Setaram, Lyon, France) was used to

record thermograms of microorganisms heated at a 3oC min-1. All DSC measurements

were conducted using fluid-tight, stainless steel crucibles. For each DSC run, the

reference crucible was filled with ~45 µl (~80 % of sample wt) of distilled water. A DSC

run was performed with unsealed, empty sample and reference crucibles to record an

empty crucible baseline. Crucibles were sealed using aluminum o-rings and were

refrigerated at 4oC prior to DSC runs. The sample and reference crucibles were placed in

the DSC and equilibrated at 1oC using a liquid nitrogen cooling system. After heating in

the DSC, samples were cooled rapidly by liquid nitrogen and rescanned to ascertain the

reversibility of thermograms. Samples were reweighed after DSC measurements to

check for loss of mass during heating. Thermograms of samples showing signs of

leakage were discarded.

34

Page 51: Calorimetric and Microbiological Evaluation of Bacteria

Heat pre-treatment in the DSC of whole cell pellets

Heat pre-treatment was performed in the DSC. Throughout the text an unheated

sample will be referred to as an untreated sample. The pellet was sealed in the sample

crucible, heated to the pre-treatment temperature and maintained at the pre-treatment

temperature for 60 seconds, followed by rapid cooling to 1oC. The sample was rescanned

from 1 to 130oC at 3oC min-1 to assess the reversibility of thermally-induced transitions in

bacterial cells. The reversibility of the transitions was evaluated by performing partial

scans between 40 and 130oC 5oC intervals. Additional pre-treatment runs were

conducted at 57.5oC for L. plantarum and at 57.5, 62.5, and 64oC for E. coli due to sharp

decreases in viability observed over the temperature range of 50-70oC.

Measurement of cell viability after heat pre-treatment

Heat pre-treatment prior to viability measurements was conducted in the DSC as

described in section 3. The crucible containing a pellet was capped (not sealed) using an

aluminum ring and screw cap. The reference crucible was filled with distilled water (~80

% of sample wt). The crucibles were refrigerated (4oC) until use. Pellets in crucibles

were heated to pre-treatment temperatures between 50 and 70oC as specified in the

previous section at a 3oC min-1 heating rate in the DSC. After rapid cooling, a portion

(40 or 50 mg) of heat treated pellet from the sample crucible was transferred using a

sterile loop to a (1.5 ml) sterile polyethylene tube. Sterile peptone water was added to

make a final volume of 1 ml with 1/25 or 1/20 (w/v) ratio. After careful suspension in

the tube, the cells were serially diluted and plated into Trypticase soy agar or MRS agar

35

Page 52: Calorimetric and Microbiological Evaluation of Bacteria

to determine viable counts. After 36 hours incubation at 37oC, viable counts of each

sample were obtained by calculation of the dilution ratio. The level of the lowest

detection was 2.5 x 101 or 2 x 101 cfu g-1 in pellet. An untreated sample was used as a

control.

Preparation and calorimetry of intact ribosomes

The protocol described by Mackey et al. (1991) with modification of buffer solutions

was applied to prepare the intact ribosomes for both bacteria. The cell pellets obtained by

centrifugation of 3.5 l of late exponential phase cultures were washed and resuspended in

20 mM HEPES buffer at pH 7.5, containing 6 mM MgCl2 and 50 mM NH4Cl. The cell

suspension was disrupted by passing twice or three times through a previously cooled

French press (AMINco SLM Instruments, Inc. Urbana, IL). Deoxyribonuclease (RNase

free) (Sigma) was added (0.4 mg ml-1) and the material was centrifuged (Beckman L85-

55M Ultracentrifuge) at 32 500 g for 30 min. The supernatant (cell-free extract) was

centrifuged at 150 000 g for 3.5 h to obtain a pellet of crude intact ribosomes. The water

content of the ribosome pellet was determined to be 65.7 % for E. coli ribosome and 64.9

% for L. plantarum ribosome on a wet basis. Pellets of intact ribosomes were placed in

the DSC sample crucible. The reference crucible was filled with HEPES buffer equal to

the amount of buffer in the sample. The crucibles were heated from 1 to 140oC at a 4oC

min-1 in the DSC.

36

Page 53: Calorimetric and Microbiological Evaluation of Bacteria

Calorimetry of intact ribosomes in different pH conditions

After freeze drying, dried pellets were weighed (~2 mg) and transferred into DSC

crucible. The pellets were mixed with ~36 mg of 50 mM potassium phosphate buffer

(pH 6, 5, 4 or 3, Fisher Chemicals, Fair Lawn, NJ). The reference crucible was filled

with potassium phosphate buffer equal to the amount of buffer in the sample. The

crucibles were heated from 1 to 140oC at a 4oC min-1 in the DSC.

Data analysis

DSC thermograms were corrected for differences in the empty crucibles by

subtracting an empty crucible baseline. Total heats corresponding to the envelope of

endothermic peaks (enthalpy, J g-1) between approximately 45-130oC for E. coli and 45-

110oC for L. plantarum were determined by integrating the temperature vs. heat flow

curve using software provided by the instrument manufacturers. A curved baseline using

three-temperature points was utilized to calculate the apparent enthalpy of both whole

cells and the intact ribosomes. Use of a curved baseline which takes into account the

apparent heat capacity change before and after the transition(s) of interest is explained in

Chapter 3. Peak temperatures for the thermally induced transitions were also determined.

37

Page 54: Calorimetric and Microbiological Evaluation of Bacteria

e

Growth of the cells to the end of exponential growth stag

Centrifugation to obtain cell pellets Intact ribosome isolation DSC

38

Plate counting

Calorimetric curvedata

MicrobiologicalViable count data

t

Analysis of the effects of heat treatmen

Heat treatment in DSC

Figure 2.2. Experimental scheme of calorimetric and microbial analysis
Page 55: Calorimetric and Microbiological Evaluation of Bacteria

RESULTS

Thermograms of E. coli and L. plantarum whole cells

Figure 2.3 shows the DSC thermograms for untreated E. coli and L. plantarum

pellets. The major peak in the DSC thermograms of both bacteria was observed over a

temperature range of 40 to 80oC. Several differences exist between the DSC

thermograms of E. coli and L. plantarum. The first peak, a1, (Tm, 56 oC), which is

proposed to be the denaturation of the smallest ribosomal subunit (30S) in E. coli

(Mackey et al., 1991), is not observed in the thermogram of L. plantarum as a separate

peak or shoulder. The major peak, peak a2, appears at a higher temperature in the E. coli

thermogram (70oC) in comparison to the L. plantarum thermogram (63oC). A peak (peak

b) similar to the peak reported by Mackey et al. (1991) as the melting of DNA in E. coli

exists, although at slightly different temperatures, in thermograms of both E. coli (94oC)

and L. plantarum (93oC). Similarly, peak c, a peak suggested by Mackey et al. (1991) to

be related to denaturation of DNA with a cell wall component appears at 102.5oC in E.

coli thermogram and at 100oC in L. plantarum thermogram. Figure 2.3 also shows that

peak d (Tm, 118oC) which appears in the E. coli thermogram, is absent from the L.

plantarum thermogram. Also apparent from Figure 2.3 is a difference in apparent heat

capacity of the live and inactivated cells (difference between the baseline before and after

the transition) of about 0.6 J g-1 K-1 for both organisms.

39

Page 56: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120

a

dc

ba

a1

3

2

Heat Flow

0.2 mW

Temperature (oC)

Figure 2.3. Thermograms of whole cells of E. coli ( ▬ ) and L. plantarum ( ••• ) obtained by DSC (1 to 150oC with 3oC min-1 heating rate).

40

Page 57: Calorimetric and Microbiological Evaluation of Bacteria

Thermograms of isolated ribosomes

Intact ribosomes from both bacteria were isolated and DSC thermograms of

ribosomes suspended in HEPES buffer at pH 7.5 were collected and compared with those

of whole cells. Two endothermic transitions were observed for E. coli ribosomes (Fig.

2.4). The L. plantarum ribosome thermogram displayed an endothermic peak with a

shoulder on the ascending side of the peak. Comparison of denaturation peaks for

ribosomes suspended in HEPES buffer at pH 7.5 showed that the peak temperature of the

transition was higher for E. coli ribosome (74.3oC) than for L. plantarum ribosome

(70.7oC), indicating higher thermal stability. The area of the peak which corresponds to

the enthalpy of ribosome denaturation also was different for each ribosome. E. coli

ribosomes display a higher enthalpy of denaturation (26.6 J g-1 dry ribosome) than L.

plantarum ribosomes (23.5 J g-1 dry ribosome). A heat capacity change as a result of

ribosome denaturation (difference between the baseline before and after the transition) is

observed for both E. coli (0.4 J g-1 K-1) and L. plantarum (0.33 J g-1 K-1).

DSC thermograms of intact ribosomes are compared to those of whole cells (Figs.

2.5a and 2.5b, thermograms A and B). For E. coli, the transition temperature (75oC) of

the major peak in the thermogram of the whole cell pellet washed in HEPES buffer

coincides with the transition temperature (74.3oC) of the ribosome denaturation peak of

the isolated intact ribosomes suspended in HEPES buffer (Fig. 2.5a). However, for L.

plantarum, while a single peak is observed for ribosomes suspended in HEPES buffer at

pH 7.5, the whole cell thermogram shows dual transition temperatures (63.5 and 68.2oC).

41

Page 58: Calorimetric and Microbiological Evaluation of Bacteria

1

2

3

4

5

6

20 40 60 80 100 120 140

App

aren

t hea

t cap

acity

(J/g

o C)

Temperature (oC)

Figure 2.4. Thermograms of isolated intact ribosomes of E. coli ( ▬ ) and L. plantarum ( ••• ) obtained by DSC (1 to 140oC with 4oC min-1 heating rate).

42

Page 59: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120 140

A

B

0.5 mWHeat Flow

(a)

Temperature (oC)

20 40 60 80 100 120 140

A

B

Heat Flow0.5 mW

(b)

Temperature (oC)

Figure 2.5. Thermograms of whole cells (A) and isolated intact ribosomes (B) of

E. coli (a) and L. plantarum (b) obtained by DSC after HEPES buffer (pH 7.5) wash.

43

Page 60: Calorimetric and Microbiological Evaluation of Bacteria

Thermograms of isolated ribosomes in acidic condition

The effect of low pH (levels 6 to 3) on the thermal characteristics of isolated E. coli

and L. plantarum ribosomes was evaluated using DSC (Table 2.1, Figs. 2.6a,b).

Decrease in transition peak of their ribosomes was accompanied with increase in acidic

heating condition. Both peak temperature and enthalpy value (J/g) of the transition are

higher in E. coli ribosome thermogram than in L. plantarum ribosome thermogram at any

pH level (Table 2.1). Figure 2.6b shows that the size of the peak (enthalpy) of the L.

plantarum ribosome was apparently decreased (~13%) at pH 5 while that of E coli

ribosome remained same (Fig. 2.6a). The thermogram of L. plantarum ribosomes

suspended in the buffer at pH 4 (Fig. 2.6b) displays a profile (transition temperatures,

61.8 and 67.4oC) which strongly resembles the major peak in the thermogram of L.

plantarum whole cells (Fig. 2.5b thermogram A).

Isolated E. coli ribosome Isolated L. plantarum ribosomeBuffer pH Transition

temperature (oC) Enthalpy (J g-1

dry wt) Transition

temperature (oC) Enthalpy (J g-1

dry wt)

6.0 70.6 28.52 69.8 17.87

5.0 67.6 28.51 66.1 15.51

4.0 65.3 23.27 61.8, 67.4 14.57

3.0 61.0 11.86 59.8 9.75

Table 2.1. Transition temperature and apparent enthalpy values for E. coli and L. plantarum ribosomes after DSC in different pH.

44

Page 61: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120 140

Heat Flow 0.2 mW

DSC in pH 3 buffer

DSC in pH 4 buffer

DSC in pH 5 buffer

DSC in pH 6 buffer

(a)

Temperature (oC)

20 40 60 80 100 120 140

Heat Flow 0.2 mW

DSC in pH 3 buffer

DSC in pH 4 buffer

DSC in pH 5 buffer

DSC in pH 6 buffer

(b)

Temperature (oC)

Figure 2.6. DSC thermogram of isolate ribosome of E. coli (a) and L. plantarum (b)

at different pH of phosphate buffer.

45

Page 62: Calorimetric and Microbiological Evaluation of Bacteria

Effect of heat pre-treatment on the DSC profiles of E. coli and L. plantarum

Pre-treatment of bacteria at various temperatures was performed in the DSC with a

partial scan to a pre-defined temperature. The reversibility of transitions after pre-

treatment was evaluated with a full second DSC scan (1 to 130oC) following the partial

scan. Thermograms displaying changes in major conformational transitions of both

organisms as a function of heat treatment are shown in Figures 2.7 and 2.8. In general,

pre-treatments below 70oC resulted in significant changes in both the temperature and the

area of the major peak observed in the thermograms of both bacteria indicating

irreversible effect of pre-treatment. It is apparent from Figures 2.7 and 2.8 that the major

peak in the DSC thermogram is obliterated after a pre-treatment temperature to 85oC for

E. coli (Fig. 2.7 thermogram F) and 75oC for L. plantarum (Fig. 2.8 thermogram F). Pre-

treatment resulted in changes in the shapes of existing peaks as well as in the appearance

of new peaks. Thermograms of un-treated bacteria are composed of several overlapping

transitions. A new peak, which was previously partially obscured, may seem to appear if

the overlapping transition disappears due to the heat treatment. For E. coli, peak a3

which is only partially visible due to the overlapping peak a2 in the control and 50-65oC

pre-treatment thermograms becomes visible after pre-treatment at 70oC (Fig. 2.7).

In Figures 2.7 and 2.8, only the full scans following pre-treatment which resulted in

major irreversible changes are displayed. For E. coli, peak b, which is associated mainly

with DNA denaturation (Mackey et al., 1991), was reversible after heat treatment at

115oC. For L. plantarum, peak b shifted to a lower temperature following a heat

treatment at 95 oC (Fig. 2.8 thermogram G) and was not reversible after the cell pellet

46

Page 63: Calorimetric and Microbiological Evaluation of Bacteria

was pre-heated to 100oC (Fig. 2.8 thermogram H). After a partial scan to 115oC, the

transition due to an outer-cell wall component of E. coli (peak d) was absent. No

evidence of native cellular components was observed in the thermogram of the 130oC

heat-treated E. coli pellet (Fig. 2.7 thermogram H). However, for L. plantarum there was

no evidence of native cellular components detected in thermograms of pellets heat-treated

at 100oC.

47

Page 64: Calorimetric and Microbiological Evaluation of Bacteria

40 60 80 100 120 140

a

a

a

1

2

3 b

c dA

H

G

F

E

D

C

B

Heat Flow0.2 mW

Temperature (oC)

Figure 2.7. Effect of heat pre-treatment on the thermogram of E. coli. Control (A), pre-treatment temperatures: 50oC (B), 60oC (C), 65oC (D), 70oC (E), 85oC (F), 115oC (G), 130oC (H). Thermograms are offset for clarity.

48

Page 65: Calorimetric and Microbiological Evaluation of Bacteria

40 60 80 100 120 140

a

a

2

3 b

A

H

G

F

E

D

C

B

Heat Flow0.2 mW

c

Temperature (oC)

Figure 2.8. Effect of heat pre-treatment on the thermogram of L. plantarum. Control (A), pre-treatment temperatures: 50oC (B), 55oC (C), 60oC (D), 65oC (E) 75oC (F), 95oC (G), 100oC (H). Thermograms are offset for clarity.

49

Page 66: Calorimetric and Microbiological Evaluation of Bacteria

Comparison of DSC thermograms and viability of E. coli and L. plantarum after heat pre-treatment DSC thermograms of pellets of each microorganism were compared to each control

thermogram after heat-treatment at different temperatures (Figs. 2.9 and 2.10). The area

under the curve of the second scan (apparent enthalpy, J g-1) was evaluated by

integration. Noticeable reductions in the apparent enthalpy value occurred with heat-

treatment up to 65oC (~53 %) for E. coli and up to 60oC (~58 %) for L. plantarum. The

viability of each microorganism treated in DSC under conditions identical to the

corresponding DSC experiment was determined and plotted on Figures 2.9 and 2.10. For

E. coli, the viable cell counts of the culture pellet displayed a slight change with heat

treatment up to 60oC. Heat treatment of E. coli to higher temperature resulted in 6-log10

unit reductions for 65oC and 7-log10 unit reductions for 70oC treatments. These

reductions were accompanied by a decrease in the area of the peak (a2) corresponding to

the denaturation of ribosomes in the thermogram (Fig. 2.9 thermograms E and F). For L.

plantarum, irreversible denaturation of ribosomes was observed in the thermogram

following a 57.5oC heat treatment (Fig. 2.10 thermogram C) with a viability loss of 2.3-

log10 units.

As described in Chapter 3, by assuming that the apparent enthalpy is proportional to

number of viable cells after correction for residual apparent enthalpy, the fractional

viability can be defined as the reduced apparent enthalpy, [(∆H-∆Hf)/(∆H0-∆Hf)], where

∆H is the apparent enthalpy after a pre-treatment, ∆Hf is the residual apparent enthalpy

after treatment resulting in no viability and ∆H0 is the apparent enthalpy of untreated

50

Page 67: Calorimetric and Microbiological Evaluation of Bacteria

cells. Fractional viability values calculated from calorimetric data [(∆H-∆Hf)/(∆H0-∆Hf)]

and plate count data (N/N0) are plotted in Figure 2.11. A linear relationship between the

reduced apparent enthalpy and the fraction of survivors is observed, except for the points

corresponding to high temperature treatment.

51

Page 68: Calorimetric and Microbiological Evaluation of Bacteria

2

4

6

8

10

12

30 40 50 60 70 80 90

A

A

B

C

C

B

ED

F

E

Heat Flow0.2 mW

a1

a 2

a3

F

Log

(CFU

/g)

Temperature (oC)

Figure 2.9. Viable counts (--∙--) and DSC thermograms of E. coli for control (A) and after heat pre-treatment at 60oC (B), 62.5oC (C), 64oC (D, thermogram not shown), 65oC (E) and 70oC (F). Thermograms are offset for clarity.

52

Page 69: Calorimetric and Microbiological Evaluation of Bacteria

2

4

6

8

10

12

30 40 50 60 70 80 90

A

A

B

B

C

D

D

E

E

F

F

Heat Flow

0.2 mW

2

3

C

a

a

Log

(CFU

/g)

Temperature (oC)

Figure 2.10. Viable counts (--∙--) and DSC thermograms of L. plantarum for control (A) and after heat pre-treatment at 55oC (B), 57.5oC (C), 60oC (D), 65oC (E) and 70oC (F). Thermograms are offset for clarity.

53

Page 70: Calorimetric and Microbiological Evaluation of Bacteria

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

N /N 0

[(∆H

-∆H

f)/(∆

H0-∆

Hf)]

Figure 2.11. Correlation between fractional apparent enthalpy and fractional viability for E. coli (o) and L. plantarum (+).

54

Page 71: Calorimetric and Microbiological Evaluation of Bacteria

DISCUSSION

This study aims to assess bacterial resistance to heat treatment by i. comparing the

thermal resistance of bacteria and the thermal stability of isolated intact ribosomes; ii.

evaluating the reversibility of thermal transitions of various cellular components

following heat treatment; and iii. developing a relationship between fractional viability

calculated from plate count data and calorimetric data.

The primary low temperature features (40~80oC) of whole cell DSC profiles of

bacteria are believed to correspond to the thermal unfolding of ribosomes (Mackey et al.,

1991). Using ribosomes isolated from E. coli, Mackey et al. (1991) showed that an

endotherm with three overlapping peaks appearing between 47 and 85oC for E. coli

whole cells is associated with ribosome denaturation. In present study, although E. coli

whole cells display a ribosomal denaturation endotherm consisting of three peaks, two

peaks (a2 and a3) are observed for ribosomal denaturation in L. plantarum thermogram

(Fig. 2.3). It has been reported that the 30S ribosomal subunit is less thermally stable

than the larger ribosomal subunit (Mackey et al., 1991; Stephens and Jones, 1993;

Bonincontro et al., 1998) suggesting peak a1 may be attributed to denaturation of the 30S

ribosomal subunit. It is also apparent from Figure 2.3 that the peaks a2 and a3 are shifted

to lower temperatures in comparison to the corresponding peaks for E. coli. The lower

peak temperatures of peaks a2 and a3 of the L. plantarum thermogram suggest that the

relative stabilities of L. plantarum ribosomes are lower than those of E. coli ribosomes.

The pH of the L. plantarum medium is reduced due to lactic acid production during L.

plantarum growth. It is possible that the increased acidity in the medium of L. plantarum

55

Page 72: Calorimetric and Microbiological Evaluation of Bacteria

may influence the stability of the ribosomal subunits directly or indirectly. Another

factor that may influence ribosome stability is altered intracellular cation concentrations,

in particular Mg2+is required for ribosome integrity (Hurst, 1984). A loss of the peak (a1)

is observed in the thermogram of acid-treated E. coli (Chapter 4. Figs. 4.4 and 4.5).

Mohacsi-Farkas et al. (1994) reported that the ribosomal denaturation peak of L.

plantarum shifted to lower temperatures as the pH of the suspending medium decreased

below 5. Furthermore, their results show that while a low temperature (Tm, 57oC)

endothermic transition appears when cells are suspended in buffer at pH 6.8 and 5, this

transition is not observed on the thermograms for whole cells suspended in buffer at pH

4.6 and lower. Because the pH of the growth medium for L. plantarum was measured to

be 4.4 in the present study, our results are in agreement with the previous data showing

an absence of peak a1 and lower transition temperatures for peaks a2 and a3 induced by

low pH. Ribosomes also were reported to be destabilized by loss of Mg2+ from cells

(Hurst and Hughes, 1978; Rheinberger et al., 1988; Anderson et al., 1991). The lack of a

visible peak, a1, in the L. plantarum thermogram may indicate denaturation of the 30S

ribosomal subunit as a result of Mg2+ loss (Tomlins and Ordal, 1976; Hurst, 1984).

Alternatively, peak a1 may be present but obscured by the other ribosomal peaks a2 and

a3 because their transition are shifted to lower temperatures.

The comparison of ribosomal denaturation for each bacterium at pH 7.5 shows that

both the transition temperature and the apparent enthalpy for E. coli are higher than those

for L. plantarum indicating a higher thermal stability and a greater energy requirement to

disrupt the structure of the E. coli ribosome (Fig. 2.4). Similar behavior is observed for

56

Page 73: Calorimetric and Microbiological Evaluation of Bacteria

whole cells suggesting that both the thermal stability and the energy required to inactivate

the bacterial cells are higher for E. coli.

E. coli ribosomes show similar thermal stabilities when in whole cells (75.1oC) and

when isolated (74.3oC) in terms of transition temperatures of corresponding peaks.

However, for L. plantarum the transition shapes and the thermal stabilities of whole cells

washed with HEPES buffer at pH 7.5 and of isolated ribosomes are similar only when the

ribosomes are suspended in potassium phosphate buffer at pH 4. Although an effect on

ribosomes is indicated in both cases, the strong resemblance in shape of observed peaks

in whole cells and isolated ribosomes may have different causes and must be explored

further. The higher thermal stability of isolated L. plantarum ribosomes when suspended

in HEPES buffer at pH 7.5 containing 6mM MgCl2 (Fig. 2.5b thermogram B) in

comparison with isolated ribosomes suspended in potassium phosphate buffer without

Mg2+ at pH 4 (Fig. 2.6b) may be attributed to stabilization of ribosomes by magnesium

ions in vitro (Noll and Noll, 1976). Similar behavior also was observed with E. coli

whole cells where the transition temperature of the major peak was 70oC following a

water wash but 75.1oC following a HEPES buffer wash. Anderson et al. (1991) note that

the ionic composition and concentration of the buffer affect the thermal stability and

shape of the ribosomal denaturation peak which in turn may affect the thermal resistance

of bacteria.

In the whole cell thermograms, there are likely more transitions than are observable

as discrete peaks. Some of these transitions may occur within the same temperature

range and be obscured by the larger ribosome denaturation peaks. Comparison of

57

Page 74: Calorimetric and Microbiological Evaluation of Bacteria

thermograms of ribosomal denaturation and whole cells (Figs. 2.7 and 2.8) shows that the

difference in heat capacity between the native and denatured states, as shown by the

difference in pre- and post-transition baselines, is 1.5 times greater for whole cells. A

positive heat capacity is typically observed for denaturation of proteins. A typical

globular protein of ~15 kDa, the change in the heat capacity is on the order of 0.4-0.67 J

g-1 K-1 (Gomez et al., 1995). Given the larger heat capacity observed for whole cells in

comparison to denaturation of ribosomes, it is probable that other cellular components

contribute to the endothermic transitions attributed to the denaturation of ribosomes.

Anderson et al. (1991) indicate that the small number of peaks observed in whole cell

thermograms can be due to a larger number of transitions including protein unfolding and

denaturation.

Another visible difference between the E. coli and L. plantarum thermograms is a

high temperature endothermic transition (peak d) only observed in the DSC thermogram

of E. coli whole cells. Mackey et al. (1991) observed a peak corresponding to peak d in

the thermogram of the cell envelope fraction and proposed this peak to be the result of

cell envelope denaturation. These investigators hypothesized that a cell wall associated

thermo-stable protein may account for the appearance of this peak. Other DSC studies in

our laboratory showed that the peak was observed in thermograms of Pseudomonas

fluorescens but was absent from the thermograms of Staphylococcus aureus and

Leuconostoc mesenteroides (unpublished results) suggesting the origin of this peak is a

cellular component of Gram-negative bacteria. In most Gram-negative bacteria the outer-

cell wall layer exists as a true unit membrane. The outer-cell wall membrane contains

58

Page 75: Calorimetric and Microbiological Evaluation of Bacteria

lipid, phospholipid, polysaccharide, and protein. The lipid and polysaccharide form a

specific lipopolysaccharide (LPS) layer. Rodriguez-Torres et al. (1993) using DSC

reported that LPS show endothermic transitions above 120oC, with the specific

temperature depending on the linkage type. Reversibility studies demonstrate that this

peak in E. coli is denatured by heat treatment above 110oC.

The transition temperatures associated with DNA denaturation are not considerably

different for both microorganisms, with that for L. plantarum is being slightly lower

(93oC) than that for E. coli (94oC). Although the thermal stabilities of the DNA for both

microorganisms are similar, their reversibility subsequent to heat treatment differs

significantly. There is no indication of a DNA peak in the thermogram of the L.

plantarum pellet pre-heated to 100oC (Fig. 2.8 thermogram H). The peak is preserved in

the thermogram of the E. coli pellet heated up to 125oC, although the apparent enthalpy is

reduced and peak is shifted to lower temperatures as the heat pre-treatment temperature is

increased. The change in energy required for denaturation of DNA subsequent to heat

treatment indicates partial refolding upon cooling or folding to a different state (Cantor

and Schimmel, 1980). Furthermore, the appearance of a previously obscured, reversible

peak (Tm at 88.4oC) in the thermogram of 95oC pre-heated L. plantarum may be due to

partial reversibility of denatured DNA (Fig. 2.8 thermogram G). Mackey et al. (1988)

showed that there is a strong correlation between guanine + cytosine (G + C) content of

DNA and the Tm of the putative DNA peak determined from a DSC scan of whole cells.

The G + C content of L. plantarum (44~46 mol %) (Kandler and Weiss, 1986) is lower

than that of E. coli (51.6 mol %) (Mackey et al., 1988). Using the empirical relation

59

Page 76: Calorimetric and Microbiological Evaluation of Bacteria

between G + C content and Tm reported by Mackey et al. (1988) gives a predicted DNA

transition temperature of 92.4-93oC for L. plantarum and 94.8oC for E. coli. Mackey et

al. (1988) also reported 94.3oC DNA transition temperature determined using DSC for E.

coli. The experimental values for DNA melting in this study are in close agreement with

the literature data, including the expectation of a lower DNA peak temperature for L.

plantarum.

DSC curves can be exploited further to determine the fractional viability of

microorganisms based on calorimetric data as described in Chapter 3. For both E. coli

and L. plantarum, Figures 2.9 and 2.10 reveal that as the severity of the heat treatment

increases, the observed peak temperature of ribosomal denaturation increases, implying

sequential damage to the ribosomal subunits and/or the existence of a range of thermal

resistance in the microorganism population. It is apparent that a loss of viability of cells

of both organisms occurs when the microorganisms are subjected to heat pre-treatment in

the range of 50~70oC. The viability loss is related to the apparent enthalpy change of

ribosomal subunits monitored by DSC because preheating to 55~70oC affected the peaks

associated with ribosome subunits but had no apparent influence on the thermally-

induced transitions of other cellular structures. Both the putative ribosomal peaks in the

thermogram and cell viability of L. plantarum were noticeably reduced by pre-heating

from 55oC to 70oC (Fig. 2.10). A similar pattern was observed in the DSC profiles of E.

coli, although the reductions in ribosomal peaks and cell viability occurred at higher

temperature (Fig. 2.9). However, as discussed in Chapter 3, the peak area corresponding

to only the ribosome transition within the whole cell thermogram can not be determined

60

Page 77: Calorimetric and Microbiological Evaluation of Bacteria

accurately because the baseline is not well defined due to overlapping transitions. Instead

the total peak area corresponding to the total apparent enthalpy must be used. With

increasing treatment temperature, the total apparent enthalpy (between approximately 50-

130oC for E. coli and 50-110oC for L. plantarum) decreases gradually compared to the

peak area for the untreated control. It is apparent from Figures 2.7 and 2.8 that residual

transitions remain even after the cells are inactivated implying that the total area under

the thermogram includes contributions related to both cell death and additional

macromolecular transitions. After subtracting the contributions due to enthalpy

associated with inactive cells, a reduced apparent enthalpy value can be defined to

determine the fraction of survivors in terms of calorimetric data. A plot of reduced

apparent enthalpy versus the fractional survivors from plate count data (Fig. 2.11) gives a

linear relationship. As I have shown in Chapter 3, these data can be interpreted in terms

of D and z values of microorganisms which are subjected to linearly increasing

temperature. In this Figure, the points close to a viability value 1 represent low

temperature treatment, while the points close to 0 represent high temperature treatments.

It is apparent that for both microorganisms at very low temperature viability calculated

from both plate count and calorimetric data are in close agreement. However, as the

treatment temperature increases a disparity appears between the plate count and

calorimetric data for both microorganisms. The disparity between the viability data

derived from the two methods is larger for L. plantarum than for E. coli. As the

temperature of the treatment increases further the disparity decreases. It is expected that

as the temperature of the heat treatment increases the number of injured microorganisms

61

Page 78: Calorimetric and Microbiological Evaluation of Bacteria

will increase. The injured microorganisms die during a complete DSC scan following

the partial scan without having a chance to repair. However, the plate count method

provides favorable conditions for injured cells to recover. We speculate that the disparity

between viabilities calculated from calorimetric data and from plate count data at

intermediate treatment temperatures may be due to the microorganisms injured during

pre-treatment. L. plantarum may have a greater tendency toward injury than E. coli. This

speculation may also explain the lag period typically observed in the semi-log survival

curves of microorganisms and needs to be explored further.

In this study, the patterns of the temperature induced changes in ribosomes, cell

envelope components, and DNA of E. coli (Gram-negative) and L. plantarum (Gram-

positive) bacteria are compared by DSC. The results indicate that more intensive heat

treatment is needed to inactivate E. coli in comparison to L. plantarum. Mohacsi-Farkas

and co-workers (1999) also reported a higher heat-inactivation temperature for E. coli in

comparison to L. plantarum. The thermal tolerance of microorganisms may depend on

the growth conditions as well as cell structure. Both the thermal stability and enthalpy of

ribosome denaturation are influenced by low pH in vitro and in vivo for E. coli and L.

plantarum. The calorimetric evaluation of the thermal stability and enthalpy change of

isolated ribosomes as a function of pH in present study shows that the values for E. coli

are higher than those for L. plantarum indicating a higher thermal stability and a greater

energy requirement to disrupt the structure of the E. coli ribosome at low pH conditions.

We have demonstrated a correlation between viability calculated from calorimetric

data and from plate count data. Calorimetric data provide unique information by direct

62

Page 79: Calorimetric and Microbiological Evaluation of Bacteria

measurement of the energy required to inactivate microorganisms. Evaluating and

quantifying differences in thermograms of whole cells and isolated components, permits

ranking of the relative thermal stabilities of the various cellular components and

identification of those most susceptible to thermal disruption.

REFERENCES

Allwood, M. C., and Russel, A. D. 1967. Mechanism of thermal injury in Staphylococcus aureus. Appl. Microbiol. 15:1266-1269. Anderson, W. A., Hedges, N. D., Jones, M. V. and Cole, M. B. 1991. Thermal inactivation of Listeria monocytogenes studied in differential scanning calorimetry. J. Gen. Microbiol. 137:1419-1424. Bonincontro, A., Cinelli, S., Mengoni, M., Onori, G., Risuleo, G. and Santucci, A. 1998. Differential stability of E. coli ribosomal particles and free RNA towards thermal degradation studied by microcalorimetry. Biophys. Chem. 75:97-103. Cantor, C. R. and Schimmel, P. R. 1980. Biophysical Chemistry Part III. The behavior of biological macromolecules, p. 1222-1223. W.H. Freeman and Company, San Francisco. Gomez, J., Hilser, V. J., Xie, D. and Freire, E. 1995. The heat capacity of proteins. Proteins: Structure, Function, and Genetics 22:404-412. Hurst, A. 1984. Reversible heat damage, p. 303-318. In A. Hurst, and A. Nasim (ed.), Repairable Lesions in Microorganisms. Academic Press, London.

63

Page 80: Calorimetric and Microbiological Evaluation of Bacteria

Hurst, A., and Hughes, A. 1978. Stability of ribosomes of Staphylococcus aureus S6 sublethally heated in different buffers. J. Bacteriol. 133:564-568. Kaletunç, G. 2001. Thermal analysis of bacteria using differential scanning calorimetry, p. 227-235. In F. Bozoglu, T. Deak, and B. Ray (ed.), Novel Process and Control Technologies in the Food Industry. IOS press, Amsterdam. Kandler, O. and Weiss, D. 1986. Regular, nonsporing, gram-positive rods, p. 1208-1234. In Peter H. A. Sneath (ed.), Bergey's manual of systematic bacteriology, vol. 2. Williams & Wilkins, Baltimore, Md. Lee, J. and Kaletunç, G. 2002. Calorimetric determination of inactivation parameters of microorganisms. J. Appl. Microbiol. 93:178-189. Mackey, B. M., Parsons, S. E., Miles, C. A. and Owen, R. J. 1988. The relationship between base composition of bacterial DNA and its intracellular melting temperature as determined by differential scanning calorimetry. J. Gen. Microbiol. 134:1185-1195. Mackey, B. M., Miles, C. A., Parsons, S. E. and Seymour, D. A. 1991. Thermal denaturation of whole cells and cell components of Escherichia coli examined by differential scanning calorimetry. J. Gen. Microbiol. 137:2361-2374. Mackey, B. M., Miles, C. A., Seymour, D. A. and Parsons, S. E. 1993. Thermal denaturation and loss of viability in Escherichia coli and Bacillus stearothermophilus. Lett. Appl. Microbiol. 16:56-58. Miles, C. A., Mackey, B. M. and Parsons, S. E. 1986. Differential Scanning Calorimetry of Bacteria. J. Gen. Microbiol. 132:939-952. Mohacsi-Farkas, Cs., Farkas, J., Meszaros, L., Reichart, O. and Andrassy, E. 1999. Thermal denaturation of bacterial cells examined by differential scanning calorimetry. J. Therm. Anal. Calor. 57:409-414.

64

Page 81: Calorimetric and Microbiological Evaluation of Bacteria

Mohacsi-Farkas, Cs., Farkas, J. and Simon, A. 1994. Thermal denaturation of bacterial cells examined by differential scanning calorimetry. Acta Aliment. 23:157-168. Niven, G. W., Miles, C. A. and Mackey, B. M. 1999. The effect of hydrostatic pressure on ribosome conformation in Escherichia coli: an in vivo study using differential scanning calorimetry. Microbiology 145:419-425. Noll, M. and Noll, H. 1976. Structural dynamics of bacterial ribosomes. V. Magnesium-dependent dissociation of tight couples into subunits: measurement of dissociation constants and exchange rates. J. Mol. Biol. 105:111-130. Rheinberger, H., Geigenmuller, U., Wedde, M. and Neirhaus, K. H. 1988. Parameters for preparation of E. coli ribosomes and ribosomal sub-units active in tRNA binding. Method. Enzymol. 164:658-662. Rodriguez-Torres, A., Ramos-Sanchez, M. C., Orduna-Domingo, A., Martin-Gil, F. J. and Martin-Gil, J. 1993. Differential scanning calorimetry investigations on LPS and free lipids A of the bacterial cell wall. Res. Microbiol. 144(9):729-740. Stephens, P. J., and Jones, M. V. 1993. Reduced ribosomal thermal denaturation in Listeria monocytogenes following osmotic and heat shocks. FEMS Microbiol. Lett. 106:177-182. Teixeira, P., Castro, H., Mohacsi-Farkas, C. and Kirby, R. 1997. Identification of sites of injury in Lactobacillus bulgaricus during heat stress. J. Appl. Microbiol. 83 (2):219-226. Tolker-Nielsen, T. and Molin, S. 1996. Role of ribosome degradation in the death of heat-stressed Salmonella typhimurium. FEMS Microbiol. Lett. 142:155-160. Tomlins, R. H. and Ordal, Z. J. 1976. Thermal injury and inactivation in vegetative bacteria, p. 153-190. In F. A. Skinner, and W. B. Hugo (ed.), Inhibition and inactivation of vegetative microbes. Academic Press, London.

65

Page 82: Calorimetric and Microbiological Evaluation of Bacteria

CHAPTER 3

CALORIMETRIC DETERMINATION OF INACTIVATION PARAMETERS OF

MICROORGANISMS*

ABSTRACT

This study aimed to apply differential scanning calorimetry (DSC) to evaluate the thermal

inactivation kinetics of bacteria. The apparent enthalpy (∆H) of Escherichia coli cells

was evaluated by a temperature scan in a DSC after thermal pre-treatment in the

calorimeter to various temperatures between 56 and 80oC. Conventional semi-

logarithmic survival curve analysis was combined with a linearly increasing temperature

protocol. Calorimetrically determined D and z values were compared to those obtained

from plate count data collected under isothermal conditions to validate the new approach.

The calculated D values using both apparent enthalpy and viability data for cells heat

treated in the DSC were similar to the D values obtained from isothermal treatment.

Temperatures for 1 through 10-log microbial population reductions, calculated from plate

count and enthalpy data were in agreement within 0.5-2.4oC at a 4oC min-1 heating rate.

66

Page 83: Calorimetric and Microbiological Evaluation of Bacteria

This novel calorimetric method provides an approach to obtain accurate and reproducible

kinetic parameters for inactivation. The calorimetric method here described is time

efficient and is conducted under conditions similar to food processing conditions.

Key Words: thermal inactivation, differential scanning calorimetry, linearly increasing

temperature, apparent enthalpy, kinetic parameter

* Adapted from Journal of Applied Microbiology, 93:178-189 (2002).

67

Page 84: Calorimetric and Microbiological Evaluation of Bacteria

INTRODUCTION The processing temperature and time necessary to produce a safe product are

determined using the D value (the time needed to reduce the population by one log) and z

value (temperature change required for a one log reduction in D value) for a target

microorganism. In general, D and z values for microorganisms are calculated from semi-

logarithmic survival curves produced as a function of time under isothermal conditions at

several temperatures. The thermal processing design is based on the thermal resistance of

target bacteria which is described by D and z values determined under isothermal

conditions. However, in industrial applications, processing temperature cannot be

reached instantly but requires a “come up time” in which a significant reduction of

microbial population may occur as the temperature rises (Peleg, 1999). Furthermore, the

reaction rate and D value are affected by the temperature and the rate of heat transfer

(Teixeira, 1992). Therefore, it is important to determine the D and z values under

conditions similar to those used in processing.

Kinetic parameters for chemical and biochemical reactions can be evaluated

analytically using differential and integral methods from data collected under non-

isothermal conditions (Deindoerfer and Humphrey, 1959; Rhim et al., 1989; Nunes et al.,

1991). There are several studies in the literature that model microorganism inactivation

during increasing temperature protocols (Reichart, 1979; Thompson et al., 1979a,b; Van

Impe et al., 1992). Reichart (1979) calculated the D and z values for Saccharomyces

cerevisae, Escherichia coli, and Bacillus stearothermophilus using the linear portion of

the survival curve generated as a function of rising temperature with decreasing heating

68

Page 85: Calorimetric and Microbiological Evaluation of Bacteria

rate. Reichart (1977) reported that similar D and z values were obtained from analysis of

viable count data produced under isothermal and increasing temperature conditions. On

the other hand, Thompson and co-workers (1979a,b), from their investigation on the

effect of heating rate on the inactivation of Clostridium perfringens and Salmonella

typhimurium reported that the D value for C. perfringens at 60oC increased from 5.3 min

to 20 min when the heating rate increased from 0oC h-1 (isothermal treatment) to 15oC h-1,

while the D value for S. typhimurium at 50oC did not change significantly when the

heating rate changed from 6oC h-1 to 12.5oC h-1. In all cases, they report D values higher

than the D50 obtained with isothermal treatment.

Some investigators have used differential scanning calorimetry to achieve heat

treatment under controlled conditions of linearly increasing temperature and to determine

the thermally-induced transitions with the ultimate purpose of evaluating the relationship

between the stability of cellular components and cell injury or death (Miles et al., 1986;

Mackey et al., 1988; Lepock et al., 1990; Mackey et al., 1991; Belliveau et al., 1992;

Mackey et al., 1993). An equation describing the rate of microorganism inactivation as a

function of linearly increasing temperature was derived by Miles et al. (1986) and was

used to determine the temperature at which the maximum death rate occurred for

vegetative cells (Miles et al. 1986) and spores (Belliveau et al. 1992). Miles and Mackey

(1994) developed the model further to predict the number of surviving microorganisms as

a function of temperature at a constant heating rate. These investigators used the

resulting equation to predict the survival of Listeria monocytogenes heated to different

final temperatures in minced beef and compared the predictions with experimental

69

Page 86: Calorimetric and Microbiological Evaluation of Bacteria

results. They also calculated the temperatures required to reduce viability by 7D at 0.1,

1, and 10oC min-1 heating rate using the published D and z values determined under

isothermal conditions. The results demonstrated that the temperatures required to

inactivate L. monocytogenes increased with the heating rate. Miles and Mackey (1994)

stated that the derived equation can also be used to calculate the D and z values under

linearly increasing temperature protocols.

Although several investigators determined the viable counts of microorganisms

following heat treatment in the DSC, no D and z values were reported. In this study, the

mathematical model proposed by Miles and Mackey (1994) was utilized to determine the

D and z values using both the viable count and calorimetric data obtained with linearly

rising temperature in DSC. The D and z values for E. coli K12 determined from the

calorimetric data were compared to the corresponding values from plate count data

obtained after heat treatment in the DSC and after isothermal treatment. The close

agreement between the calorimetrically determined microorganism inactivation kinetic

parameters and those of determined from plate count data demonstrates the advantage of

this unique approach in obtaining reproducible and accurate results in a short time.

70

Page 87: Calorimetric and Microbiological Evaluation of Bacteria

MATERIALS AND METHODS

Source and preparation of organisms

E. coli K12 was obtained from the Culture Collection, Department of Microbiology

at the Ohio State University. A loopful of organism was revived in 10 ml Trypticase soy

broth (Difco laboratories, Detroit, MI) supplemented with 0.3 % (w/w) yeast extract

(Difco laboratories, Detroit, MI) (TSBYE) and incubated at 37oC for 18 hours. The

culture was stored frozen (-80oC) in 30 % (v/v) sterile glycerol. A loopful of the stock

culture was transferred to 10 ml TSBYE and incubated 10 hrs at 37oC before use.

A growth curve for E. coli was generated from viable count data to determine the

time required to reach the late exponential growth phase. Culture was inoculated (1 %

v/v) into TSBYE and incubated at 37oC until reaching late exponential growth phase,

when the final concentration of cells in the medium was 1.0 ± 0.1 x 109 cfu ml-1. After

reaching the late exponential phase, the growth medium was centrifuged (Beckman J2-

21 centrifuge) at 10 000 g for 10 min at 4oC to separate the cells as pellets. The pellet

was washed with 150 ml of sterile distilled water before transfer into DSC crucibles. A

differential scanning calorimeter (DSC 111, Setaram, Lyon, France) was used for

collection of all thermograms. A portion (~100 mg) of the pellet was transferred into a

tared (1.5 ml) polyethylene tube, weighed, freeze dried (Freezone 4.5, Labconco Freeze

Dry System), and reweighed to determine the percentage of dry matter in the pellet. The

amount of moisture in the E. coli pellet (wt/wt) used in the DSC experiments was

determined to be 83 ± 0.3 % by freeze drying.

71

Page 88: Calorimetric and Microbiological Evaluation of Bacteria

Apparent enthalpy value of E. coli K12 from DSC

A DSC thermogram with empty stainless steel sample and reference crucibles was

collected to measure the empty crucible baseline. Temperature calibration was

confirmed using an indium sample in the stainless steel crucible at a constant heating

rate of 4oC min-1. Pellets of cells were carefully transferred into the sample crucible and

weighed (70 ± 0.3 mg wet weight). When the reference crucible is left empty an artifact

due to the heat capacity imbalance between crucibles is observed at the initiation of

temperature scanning. A known quantity of water, similar in mass to the moisture in

sample, was placed in the reference crucible to eliminate the artifact. The reference

crucible was filled with 58 ± 0.2 mg (~83 % of sample wt) of distilled water. Both

crucibles were sealed using aluminum o-rings. The sealed crucibles were refrigerated

(4oC) until used for DSC. The sample and reference crucibles were placed in the DSC

and equilibrated at 1oC. Pellets were pre-heated in the DSC to 56.7, 58.7, 60.7, 62.7,

64.7, 66.7, 68.7, 70.7, 75.7, or 80.7oC with 4oC min-1 heating rate. After pre-heating in

the DSC, samples were cooled immediately by liquid nitrogen, equilibrated at 1oC, and

rescanned to 140oC at 4oC min-1. Samples were reweighed after DSC measurements to

check for loss of mass during heating. Thermograms of samples showing signs of

leakage were not used. A control scan was recorded by heating the pellets from 1oC to

140oC at 4oC min-1. DSC thermograms of samples were corrected for differences in the

empty crucibles by subtraction of an empty crucible baseline. Peak areas (apparent

enthalpies, J g-1) corresponding to the contributions of survivors were determined from

the apparent heat capacity vs. temperature graph using software provided by the

72

Page 89: Calorimetric and Microbiological Evaluation of Bacteria

instrument manufacturer. Apparent enthalpy values were also corrected for the amount

of dry matter in the pellets based on the freeze-dryer results. Because the moisture

content of the pellet was not known prior to the DSC measurement, the reference

crucible was filled with water equal to 83 % of the sample weight. When the moisture

content was determined, the difference between the sample moisture content and the

amount of water in the reference crucible was calculated. The enthalpy contributions of

excess water, based on the measured difference in water content, was calculated by

integrating the equation describing the specific heat of water as a function of

temperature over the temperature limits used to calculate the apparent enthalpy. The

apparent enthalpy value was corrected by adding (subtracting) the enthalpy contribution

of the water if the moisture content of the sample was lower (greater) than the amount of

water in the reference crucible.

The number of E. coli K12 survivors after DSC

Weighed pellets of cells (~70 mg wet weight) were transferred into sterile DSC

crucibles using sterile loops. A stainless steel cap, an aluminum o-ring and a screw cap

was placed in each crucible without sealing. The reference crucible was filled with

distilled water (~83 % of sample wt). The capped crucibles were kept under

refrigeration (4oC) until used for DSC. Pellets in crucibles were heat-treated to 56.7,

58.7, 60.7, 62.7, 64.7, 66.7, 68.7, 70.7oC at 4oC min-1 heating rate using the DSC

instrument. After cooling, a portion (50 mg) of the heated pellet from each crucible was

transferred to a (1.5 ml) sterile polyethylene tube using a sterile loop. Sterile peptone

73

Page 90: Calorimetric and Microbiological Evaluation of Bacteria

water was added to make a final volume of 1 ml with 1/20 (w/v) ratio. After careful

suspension in the tube, the cells were serially diluted and plated onto Trypticase soy agar

to determine viable counts. After 36 hours incubation at 37oC, viable counts of each

sample were obtained by calculation of dilution ratios. The level of the lowest detection

was 2 x 101 cfu g-1 in pellet.

The number of E. coli K12 survivors after isothermal heat treatment

Weighted cell pellets (70 mg wet weight) were transferred into thin-walled

polypropylene reaction tubes containing smug-fitting snap caps. The tubes were

submerged in a temperature-controlled water bath stabilized at 56, 58, 60, 62, or 64oC.

The temperature was continuously monitored by a thermocouple placed in the water

bath next to the tubes. The temperature variance was ~±0.02oC during treatment. The

tubes were removed at time intervals; 10, 20, 30, 40 min for 56oC, 2, 5, 10, 20, 30 40

min for 58oC, 1, 2, 4, 5, 8, 11, 14 min for 60oC, 1, 2, 3, 4 min for 62oC, 0.5, 1, 1.5, 2 min

for 64oC, and were cooled in an ice-water bath. The cells in the tube were serially

diluted and plated onto Trypticase soy agar to determine viable counts using the

procedure described above.

DSC data analysis

DSC thermograms were corrected for differences in the empty crucibles by

subtracting an empty crucible baseline. Total heats corresponding to the envelope of

endothermic peaks (enthalpy, J g-1) between approximately 40 and 130oC were

74

Page 91: Calorimetric and Microbiological Evaluation of Bacteria

determined by integrating the temperature vs. heat flow curve using software provided

by the instrument manufacturer. A curved baseline taking into account the variation in

heat capacity before and after the transition passing through three designated points on

the thermogram was used to calculate the apparent enthalpy of whole cells. Data points

at three temperatures were selected to determine the baselines for all DSC curves as

shown in Figure 3.4. The initial temperature point was on the pre-transition baseline

(40oC). The final point was on the post-transition baseline (130oC). An intermediate

point was selected at a temperature below the onset of the final peak which corresponds

to transitions in the cell envelope (108oC).

THEORY

Miles and Mackey (1994) developed a model to predict the viability of

microorganisms which have subjected to linearly increasing temperature. Their

derivation combining a first order inactivation model for microorganisms with a linearly

increasing temperature protocol was adapted to analyze the calorimetric data as described

below.

According to first order inactivation kinetics, the ratio of survivors as a function of

treatment time at a constant temperature can be described as;

75

Page 92: Calorimetric and Microbiological Evaluation of Bacteria

Dt

NN

−=0

10log (1)

where, N is the number of survivors at time t and N0 is the initial number of viable cells.

The z value can be calculated from the D values determined at least at two temperatures

as the slope of a line of log D vs. temperature;

z

TTDD e

e

−=10log (2)

where, T is the temperature and De is the D value at an arbitrary temperature Te.

When the temperature increases at a constant linear heating rate, r; from an initial

temperature T0:

rtTT += 0 (3)

where, T0 is the initial temperature

the number of survivors is given by an equation, derived by Miles et al. 1986:

⎭⎬⎫

⎩⎨⎧

⎥⎦

⎤⎢⎣

⎡⎟⎠

⎞⎜⎝

⎛ −−⎟

⎞⎜⎝

⎛ −−=

zTT

zTT

rDz

NN ee

e

)(303.2exp

)(303.2expexp 0

0

(4)

When

20 >−zTT

,

76

Page 93: Calorimetric and Microbiological Evaluation of Bacteria

equation 4 reduces to

⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛ −−

≈z

TTrDz

NN e

e

)(303.2expexp

0

(5)

Equation 5 is converted into linear form by taking the logarithm of both sides twice,

ee

TzrD

zTzN

N 303.2ln303.2lnln0

−+=⎥⎦

⎤⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛− (6)

slope intercept

The slope (2.303/z) of ln[-ln(N/N0)] vs. T curve is used to calculate a z value. The z value

and the intercept from the graph are used to calculate a De value for the microorganism at

a given heating rate, r. Values for D at any temperature are obtained from equation 2.

The value N/N0 represents the fraction of survivors as a result of heat treatment. The

apparent enthalpy, ∆H, is the area under the DSC thermogram, ∆H = ∫ Cp dT.

Assuming that ∆H is proportional to the number of survivors after correction for the

residual area observed for killed cells and macromolecular transitions, ∆Hf, we can write

an expression for the fraction of survivors in terms of the DSC observable;

f

f

HHHH

NN

∆−∆

∆−∆≈

00

(7)

77

Page 94: Calorimetric and Microbiological Evaluation of Bacteria

Substitution of equation 7 into equation 6 yields;

eef

f TzrD

zTzHH

HH 303.2ln303.2lnln0

−+=⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

∆−∆

∆−∆− (8)

By analogy to the viability data, the z value is determined from the slope (2.303/z) of

ln[-ln(∆H-∆Hf)/(∆H0-∆Hf)] vs. T (Eqn. 8 and Fig. 5). Once De is determined from the

intercept, D at any temperature can be obtained from equation 2.

The temperature for any given log reduction in survivors, n, is calculated by rearranging

equation 6 as;

⎥⎦

⎤⎢⎣

⎡+⎟

⎠⎞

⎜⎝⎛+= )303.2ln(ln

303.2n

zrDzTT e

e (9)

where n = -log N/N0

78

Page 95: Calorimetric and Microbiological Evaluation of Bacteria

e

C

Growth of the E. coli cells to the end of exponential stag

Centrifugation to obtain cell pellets

Heat-treatment in DSC

i

g g

a a

Apparent enthalpy

data

s

79

Heat-treatment in sothermal condition

DS

Plate countin

Viable count dat

Applied to the equation

e

Calorimetric D and z value Isothermal D and z valu

Viable count dat

Plate countin

Figure 3.1. Experimental scheme of calorimetric and microbial analysis
Page 96: Calorimetric and Microbiological Evaluation of Bacteria

RESULTS

Heat treatment of E. coli with linearly rising temperature in DSC

DSC thermograms displaying the thermally-induced transitions of E. coli K12

control cells and the cells immediately subsequent to thermal pre-treatment are shown in

Figure 3.2. The pre-treatment was accomplished by partial scanning to temperatures

between 56.7 to 80.7oC followed by rapid cooling to the initial temperature. The

scanning rate was 4oC min-1 for both partial and subsequent complete scans. It is

apparent from Figure 3.2 that as the final temperature of the partial scan increases, the

shape and size of some transitions in the second scan change compared to the control

scan; while, other transitions remain unchanged. These observations indicate irreversible

and reversible changes in cellular components as a result of thermal pre-treatment. An

increase in pre-heating temperature results in a decrease in the area of the transition peak

(apparent enthalpy, ∆H, J g-1) observed in the 50~85oC region, an indication of

irreversible changes. The onset temperature of the first major transition also was

determined from the thermograms. A closer examination of Figure 3.2 reveals that the

onset temperature of the transition increases up to 70.7oC as the temperature of the pre-

treatment increases to 70.7oC. Further increases up to 80.7oC in the pre-treatment

temperature do not result in an increase in the onset temperature although the apparent

enthalpy calculated continues to decrease.

The DSC thermograms of E. coli K12 also exhibit a significant difference between

the apparent specific heat capacities of the live and inactivated cells. These heat capacity

differences are apparent as differences between the pre- and post transition baselines.

80

Page 97: Calorimetric and Microbiological Evaluation of Bacteria

Figure 3.2. Apparent specific heat capacity versus temperature curves of control and heat-treated E. coli. Curves are displayed by 0.1 Jg-1 K-1 for clarity.

20 40 60 80 100 120 140

Temperature (oC)

control

56.7 o C pre-heated

58.7 o C pre-heated

60.7 o C pre-heated

62.7 o C pre-heated

64.7 o C pre-heated

70.7 o C pre-heated

68.7 o C pre-heated

66.7 o C pre-heated

80.7 o C pre-heated

75.7 o C pre-heated

0.5 J g -1 K -1

81

Page 98: Calorimetric and Microbiological Evaluation of Bacteria

Evaluation of D and z values of E. coli K12 from viable counts data after a heat

After a partial scan in the DSC at 4 C min to the temperature specified in Table

3.1, the E. coli pellets were cooled rapidly and the pellet was removed from the crucible.

The number of survivors was determined using plate counting. If the assumption of first

order inactivation kinetics is valid, according to equation 6 ln[-ln(N/N0)] should be a

linear function of temperature. It is apparent from Figure 3.5 that over the temperature

range of 58.7 to 68.7 C a linear relationship exists between ln[-ln(N/N0)] and temperature

with regression coefficient r2 = 0.974. The number of survivors after heating to 56.7oC

with 4 C min in the DSC was equal to the initial number of viable cells, while no

survivors were detected when the cells were heated to 70.7 C. The D value at 60 C and z

value were calculated from the intercept and the slope of the line displayed in Figure 4

and were found to be 5.9 min and 3.8 C, respectively.

Evaluation of D and z values of E. coli K12 from calorimetric data after a heat

Following a partial scan to the temperature specified in Table 3.1 and rapid cooling

to 1 C, the E. coli pellets were scanned to monitor the thermally-induced transitions

associated with the bacterial cells surviving the partial scan. The DSC thermograms were

normalized to yield the apparent specific heat capacity (J g K ) as a function of

temperature. The area under the curve of the second scan (apparent enthalpies, J g ) was

evaluated by integrating the apparent heat capacity vs. temperature curve. A curved

baseline using three-temperature points was utilized to calculate the apparent enthalpy.

treatment to different final temperatures in DSC

o -1

o

o -1

o o

o

treatment to different final temperatures in DSC

o

-1 -1

-1

82

Page 99: Calorimetric and Microbiological Evaluation of Bacteria

Treatment temperature (oC) *Viable counts (cfu g-1) †Apparent enthalpy (J g-1)

Control 3.6 x 1011 4.32

56.7 3.6 x 1011 4.28

58.7 3.4 x 1011 4.19

60.7 2.6 x 1011 3.89

62.7 2.1 x 1011 3.60

64.7 8.2 x 109 3.15

66.7 7.9 x 105 2.78

68.7 3.0 x 102 2.70

70.7 <2.0 x 101 2.67

*Plate count data after heat treatment in the DSC. †Calorimetric data after heat treatment in the DSC.

Table 3.1. Viability and apparent enthalpy values for E. coli K12 cell pellets after pre-treatment in the DSC.

83

Page 100: Calorimetric and Microbiological Evaluation of Bacteria

D value (min) z value (oC)

Heating medium for E. coli

Temperature

(oC)

Isothermal

Non-

isothermal

Isothermal

Non-

isothermal

Reference

Nutrient broth 56 4.5 4.9 Chambers et al. (1957)

Ringer solution 55 4.0 Lemcke and White (1959)

Sucrose solution (0.99 aw) 57.2 1.2 Goepfert et al (1970)

Tryptic soy broth 52 25.6 Stiles et al. (1973)

Milk solution (10 %) Milk solution (51 %)

58 58

1.40 13.5

4.67.9

Dega et al. (1972)

Glucose solution (0.5 %)

55 56

57.2 60

3.54

0.29

6.7

3.52 1.6

0.27

3.68

3.58

Reichart (1979)

Cell pellets

58 60 62 64

5.90 1.80 0.50 0.24

4.23 Present work

84

Table 3.2. D and z values reported for E. coli from isothermal and non-isothermal heat treatments.

Page 101: Calorimetric and Microbiological Evaluation of Bacteria

A representative curve and baseline for the E. coli control pellet heated from 1 to 140oC

are displayed in Figure 3.4. The apparent enthalpy value was found to be 4.32 J g-1 wet

weight. Because this sample was not subjected to a partial scan, the measured apparent

enthalpy corresponds to the ∆H0 value. The peak areas of heat-treated E. coli pellets (∆H)

were determined using the same analysis procedure applied to the second scan. The

apparent enthalpy values were corrected to account for moisture content differences

between the sample and reference crucibles. The apparent enthalpy of maximally treated

cells, ∆Hf, was calculated from the second scan after a partial scan to 70.7oC. The criteria

for selecting 70.7oC pre-treatment to calculate ∆Hf are discussed later in discussion

section. As expected from Equation 8, the ln[-ln(∆H-∆Hf)/(∆H0-∆Hf)] vs. temperature

graph produced a straight line (r2 = 0.988) from which the D value at 60oC and the z value

were calculated to be 6.1 min and 5.2oC, respectively (Fig. 3.6).

Evaluation of D and z values of E. coli K12 from viable counts data after isothermal heat treatment The number of survivors in E. coli pellets subjected to isothermal heat treatment at

58, 60, 62, and 64oC were determined. D values for each isothermal treatment were

determined from the slope of ln(N/N0) vs. time plots (Table 3.2). According to Equation

2, log10(D/D58) values were plotted against temperature and the z value was calculated to

be 4.23 oC from isothermal treatment studies.

85

Page 102: Calorimetric and Microbiological Evaluation of Bacteria

DISCUSSION

DSC thermograms of microorganisms display several endothermic or net

endothermic transitions which may be a combination of exothermic and endothermic

events (Verrip and Kwast, 1977; Miles et al., 1986; Lepock et al., 1990; Anderson et al.,

1991; Mackey et al., 1991; Kaletunç, 2001). Several main peaks of the E. coli

thermogram can be assigned to thermally induced transitions of particular cellular

components by comparison to the transition temperatures of isolated cell components

(Mackey et al., 1991). It has been reported that a strong relationship exists between

thermal death and the peaks observed at 50~85oC for vegetative microorganisms (Miles

et al., 1986; Teixeira et al., 1997; Niven et al., 1999). Mackey et al. (1991) and Niven et

al. (1999) proposed that the temperature region over which the major reduction due to

pre-heat treatment in the area under the peaks is observed in the DSC thermogram is the

result of denaturation of the main ribosomal subunit. To calculate the enthalpy associated

with a certain peak, the baselines in the temperature regions well below and well above

the transition region should be apparent. In the DSC thermograms of microorganisms,

the existence of several overlapping transitions resulting from various cellular

components does not allow one to evaluate the enthalpy of the individual transitions. Pre

and post-transition baselines can be readily defined to determine total area under the

thermal transition curve. Therefore, we have measured the total apparent enthalpy of all

transitions. Miles et al. (1986) used a similar approach to determine the total enthalpy of

denaturation for a number of bacteria and reported that there was a significant difference

86

Page 103: Calorimetric and Microbiological Evaluation of Bacteria

among bacteria, with enthalpy of denaturation values based on the dry matter content

ranging from 9 to 20 J g-1 dry matter. The total enthalpy of denaturation for the E. coli

K12 used in this study is 23 J g-1 dry matter.

A significant portion of the difference observed between the apparent specific heat

capacities of the live and inactivated cells occurs between the baselines prior to and after

the major peak which is attributed to disruption of the ribosome structure. As the

preheating temperature increases, the amount of inactivated cells during preheating

increases resulting in a visible reduction in the apparent heat capacity difference between

the pre- and post transition baselines in the rescan thermogram (Fig. 3.2). A similar

observation was reported by Miles et al. (1986) as the rescan thermogram displayed

slightly higher specific heat capacities at low temperatures compared to the initial scan.

Denaturation of proteins typically is accompanied by a positive heat capacity change.

For a typical globular protein of ~15 kDa, the change in the heat capacity is on the order

of 1.5-2.5 kcal K-1 mol-1 (0.4-0.67 J g-1 K-1) and may have a profound effect on the

overall energetics (Gomez et al. 1995). The apparent heat capacity change for E. coli

pellets upon inactivation is determined to be of the same order of magnitude, 0.6 J g-1 K-1

(Fig. 3.2).

In the absence of a heat capacity change over the transition, the enthalpy

calculation is carried out by constructing a linear baseline connecting the segments of the

thermogram before and after the peak and by evaluating the area between the peak and

the baseline. The construction of the baseline is critical for accurate evaluation of the

enthalpy value for a first order transition coupled with a heat capacity change as typically

87

Page 104: Calorimetric and Microbiological Evaluation of Bacteria

observed in thermograms of microorganisms. Upon rescanning of the samples which

were heated to 140oC, we observed a superimposition of baselines between the two scans

above 130oC (Fig. 3.3); implying that the heat capacity of the heat damaged cells is

independent of any thermal treatment. Therefore, a curved baseline was used between

the segment of the thermogram prior to the major thermally-induced transition (~40oC)

and the segment of the thermogram after the last peak (~130oC) (Fig. 3.4). The total peak

area was determined from the first scan of the control sample. For heat treated samples,

the second scan after a partial scan to pre-treatment temperature was used to calculate the

total apparent enthalpy. The heat flow data after correction for the difference between

the empty crucibles was used to calculate the total apparent enthalpy. The resultant

enthalpy value is further corrected for the difference in the amount of water between the

sample and reference crucibles.

A close examination of Figure 3.2 reveals that the peaks attributed to the ribosomal

subunits in the DSC thermogram (observed between 50 and 85oC) disappeared when the

E. coli pellet was heated to 80.7oC, while the thermal transitions associated with other

cellular components appear to be unaffected by pre-heating to the same temperature. The

absence of viability revealed in the plate counts of cells preheated to 70.7oC in the DSC

suggests that the total area under the thermogram peaks includes contributions directly

related to the cell death as well as contributions due to additional macromolecular

transitions. Macromolecular changes may include conformational changes and phase

transitions that manifest themselves as a combination of overlapping endothermic and/or

exothermic events resulting in a net endothermic profile. Residual peak area which

88

Page 105: Calorimetric and Microbiological Evaluation of Bacteria

Hea

t flo

w (m

W)

Temperature (oC)

Figure 3.3. A typical DSC thermogram for whole cells of E. coli K12 after empty crucible baseline subtraction. initial scan (a), rescan after cooling (b).

89

Page 106: Calorimetric and Microbiological Evaluation of Bacteria

Hea

t flo

w (m

W)

Temperature (oC)

Figure 3.4. DSC thermogram for whole cells of E. coli K12 displaying curve base line used to determine the apparent enthalpy value.

90

Page 107: Calorimetric and Microbiological Evaluation of Bacteria

corresponds to apparent enthalpy changes associated with macromolecular transitions is

observed even after the cells are completely inactivated. The residual apparent enthalpy

(∆Hf) determined from samples pre-treated at 70.7oC was subtracted from the total

apparent enthalpy calculated from each DSC rescan to compensate for the contributions

of inactive cells. After pre-treatment to 70.7oC no viability was detected in the plate

count data. The analysis of each rescan thermogram showed that the onset temperature

of the major peak increases with increasing pre-treatment temperature up to 70.7oC and is

unchanged for pre-treatment temperatures of 70.7, 75.7, and 80.7oC. Furthermore, the

continuous decrease in apparent enthalpy stabilizes around 70.7oC (Table 3.1), but

continues to decrease to 1.86 J g-1 at 75.7oC and to 1.69 J g-1 at 80.7oC. Consequently,

the pre-treatment temperature used to calculate the residual apparent enthalpy (∆Hf) can

be determined using either plate count data or calorimetric data. Therefore, the fractional

enthalpy change associated with pretreatment is used to estimate the reduction in the

viable cell population. It is apparent from equation 7 that the fractional enthalpy

calculation depends on the choice of ∆Hf. Thus it is important to determine the value of

(∆Hf) carefully.

The corrected apparent enthalpies (∆H) of cells determined from second scans after

an initial scan to various temperatures in the DSC together with ∆H0 and ∆Hf values were

used to construct a fractional survivor versus temperature graph according to equation 8

(Fig. 3.6). Similar to the DSC based plate count data (Fig. 3.5), a linear relationship

(r2=0.988) is observed between the reduced apparent enthalpy [(∆H-∆Hf)/(∆H0-∆Hf)] and

the temperature. It is apparent that the assumption of equation 7, in which the fractional

91

Page 108: Calorimetric and Microbiological Evaluation of Bacteria

-4

-3

-2

-1

0

1

2

3

4

55 60 65 70

Temperature (oC)

ln[-

ln(N

/N0

)]

Figure 3.5. Temperature dependence of fractional survivor population determined from plate count data after heat pre-treatment of E. coli cells in the DSC.

92

Page 109: Calorimetric and Microbiological Evaluation of Bacteria

-3

-2

-1

0

1

2

3

55 60 65 70

Temperature (oC)

ln[-

ln((∆H

-∆H

f)/(∆

H0

-∆H

f)]

Figure 3.6. Temperature dependence of fractional survivor population determined from calorimetric data after heat pre-treatment of E. coli cells in the DSC.

93

Page 110: Calorimetric and Microbiological Evaluation of Bacteria

survivor population from plate count viability and apparent enthalpy data are

proportional, is valid.

Microbial survival curves typically are parameterized in terms of D and z values,

where D represents the time required to reduce the population by one log unit and z

represents the temperature change required to reduce D by one log value. There are a

number of D and z values for E. coli, reported in the literature (Table 3.2). It is apparent

that even the D and z values determined from isothermal treatment are influenced by the

medium in which the bacteria are suspended. Furthermore, thermal resistance may

depend on the strain of bacteria. In this study, E. coli pellets with 1011 cfu mg-1 bacterial

concentration were prepared in order to produce a measurable and reproducible heat flow

signal in the DSC for calorimetric analysis of bacterial death. E. coli pellets of 70 mg are

used for comparison of D and z values obtained from both plate count and apparent

enthalpy data of microorganisms treated in the DSC under a 4oC min-1 heating rate.

Isothermal heat treatments in a water bath were applied to approximately 70 mg E. coli

pellets in order to be able to compare the D and z values from isothermal and

nonisothermal temperature protocols. For comparison, our D and z values calculated

under isothermal treatment conditions are included in Table 3.2. Although of the same

order of magnitude, D values calculated in this study are clearly higher than the values

reported in the literature. This could be due to a higher equilibration time required for the

pellet to reach a constant temperature compared to a microorganism sample in a capillary

tube or to differences between our strain and those for which D and z values are reported

in the literature. Very high cell densities, such as in the pellet in this study, are associated

94

Page 111: Calorimetric and Microbiological Evaluation of Bacteria

with increased heat resistance of spore-forming and nonsporing microorganisms (Hansen

and Riemann, 1963; Beaman et al., 1981; Jay, 1996). Also, it is known that the values of

D and z depend on the composition of the medium, and the washing solution, and the

stage of growth (Strange and Shon, 1964; Hoffman et al., 1966; Tomlins and Ordal,

1976). It should be emphasized that the purpose of the study is to compare the D and z

values of the E. coli samples prepared from the same strain, with same microorganism

concentration, and amount of sample using different heating strategies and also to

demonstrate that calorimetric data can be utilized to evaluate the thermal resistance of

microorganisms.

Comparison of D values in Figure 3.7 reveals that D values obtained from data

collected under linearly increasing temperature protocols were higher than those obtained

under isothermal conditions. Several factors may lead to such results. The D values

determined from isothermal kinetic data may be lower than the actual values, because at

all of the isothermal treatment temperatures (58, 60, 62, 64oC) an initial lag time

unaccounted for in the standard model was observed. Because the number of points from

the log-linear portion of the survival curve were much larger compared to the number of

points in the initial lag time portion, calculated D values from the fitted data were

dominated by the linear portion of the survival curve which would be expected to lead to

an underestimation of D values compared to the actual ones.

If the heating rate is higher than the time necessary for the samples to reach

thermal equilibrium a thermal gradient may be established in the sample. The thermal

gradient is a function of sample dimensions, thermal diffusivity, and the heating rate.

95

Page 112: Calorimetric and Microbiological Evaluation of Bacteria

0

2

4

6

8

10

55 60 65 70 75

Temperature(oC)

D v

alue

(min

)

Figure 3.7. Comparison of D values calculated from the calorimetric and viability data obtained under non-isothermal heat treatment in the DSC and D values obtained from isothermal heat treatment. Calorimetric data ( ), Viability data ( ), Isothermal data ( ).

96

Page 113: Calorimetric and Microbiological Evaluation of Bacteria

Depending on the magnitude of the temperature gradient, a distribution of active, injured,

and inactive cells may be present in the pellet at any time. Miles and Mackey (1994)

calculated a 3oC thermal gradient between the surface and the center of a cylindrical

container of 0.01m radius filled with Listeria monocytogenes inoculated beef with a

thermal diffusivity of 1.4 x 10-7 m2 s-1. The DSC crucibles used in this study were

cylindrical shape with a radius of 0.0025 m. The pellets had approximately 83 ± 0.3 %

by weight water. Assuming a thermal diffusivity of 1.3 x 10-7 m2 s-1, similar to that of a

starch gel with 82 % moisture (Andrieu et al., 1989), the temperature gradient between

the center and the surface of the crucible calculated to be 0.8oC. If such a temperature

correction is applied to the data in Figure 3.7, both of the curves describing D values

obtained under linearly increasing temperature conditions approach the D values obtained

under isothermal conditions. This correction becomes more significant in the steep part

of each curve. A close examination of Figures 3.7 and 3.8 reveals that even an order of

magnitude change in D value at low temperatures (Fig. 3.7) results in a reduction in

microbial population of less than 0.5 log units (Fig. 3.8). At higher temperatures, where

significant microbial population reduction occurs, the differences among the D values

obtained by isothermal plate count, nonisothermal plate count, and nonisothermal

calorimetric data are insignificant.

The D values obtained using linearly increasing temperature protocols may be

influenced by heat transfer limitations as well as by stress adaptation of microorganisms

as a result of sublethal heat treatments during constant heating rate schemes. If slow

heating rates are employed, temperature adaptation may occur during the heating process

97

Page 114: Calorimetric and Microbiological Evaluation of Bacteria

which may lead to an increase in the thermal tolerance of microorganisms and higher D

values compared to those obtained under isothermal conditions. The increase in heat

resistance under linearly increasing temperature protocols was reported to be dependent

on the heating rate (Tsuchido et al., 1982; Mackey and Derrick, 1987). The study on

Salmonella Typhimurium by Mackey and Derrick (1987) reveals that microorganism

survival under isothermal conditions subsequent to a 0.6oC min-1 heating rate was greater

in comparison to survival after heating at 10oC min-1 heating rate. The impact of

sublethal heat treatment on the thermotolerance of food pathogens has been studied

extensively (Bunning et al., 1990; Farber and Brown, 1990; Linton et al., 1992; Murano

and Pierson, 1992). Linton et al. (1992) reported that a 10 min heat treatment of L.

monocytogenes at 48oC, increased the D value at 55oC, by more than two-fold. However,

at heating rates typical of domestic or commercial cooking practices (0.5-5.0oC min-1),

the sample temperature is expected to increase 5oC or more in 10 minutes. Miles and

Mackey’s study (1994) showed that the temperatures required to reduce viable numbers

of L. monocytogenes by 7D increased as the heating rate increased from 0.1 to 10oC min-

1. A similar temperature increase of 5oC was observed when the heating rate increased

from 0.1 to 1oC min-1 and from 1 to 10oC min-1. Because heat adaptation of a

microorganism is more likely to occur at the slowest heating rate of 0.1oC min-1, it is

expected that a smaller increase in sample temperature required to kill 7D when the

heating rate is increased from 0.1 to 1oC min-1 in comparison to a heating rate increase

from 1 to 10oC min-1. Heat adaptation may become a greater concern for isothermal

98

Page 115: Calorimetric and Microbiological Evaluation of Bacteria

0

1

2

3

4

5

6

7

8

9

10

55 60 65 70 75Temperature (oC)

n, L

og r

educ

tion

Figure 3.8. Predicted log reductions in cell population as a function of temperature using a 4 oC min-1 heating rate (equation 9) using apparent enthalpy data ( ), viability data ( ), and isothermal data for holding time of 5 sec ( ).

99

Page 116: Calorimetric and Microbiological Evaluation of Bacteria

treatments, because as the sample temperature approaches the set temperature, the driving

force decreases causing an increase in the time required to reach the set temperature. The

temperatures required to reduce the number of viable E. coli by 0.1 through 10 log units

in the DSC employing a constant heating rate of 4oC min-1 are predicted using equation 9

(Fig. 3.8). The difference between the predicted temperatures using plate count and

enthalpy data varies between 0.5-2.4oC over the 10-log unit reduction in survivor

population. For comparison, corresponding curves derived from the isothermal data for

5 sec exposure time is also displayed in Figure 3.8. The calculation of the temperatures

for isothermal treatment was carried out using the following equation,

⎟⎟⎠

⎞⎜⎜⎝

⎛−=

ee nD

tzTT 10log (10)

where t is the time in minutes.

The plots in Figure 3.8 suggest that the log population reduction of the E. coli cells

treated at isothermal conditions with a 5 sec holding time is equivalent to the log

reduction of cells calculated from the enthalpy data obtained with a linearly increasing

temperature treatment at 4oC min-1.

DSC provides information on the thermal and thermodynamic stability of materials.

The thermal analysis of both spores (Belliveau et al. 1992) and vegetative cells (Miles et

al. 1986) demonstrated a clear relationship between the onset of cellular component

denaturation observed in the DSC and the thermal resistance of the organisms. The

maximum death rate of the organisms occurred above the onset of thermal denaturation

while the maximum growth temperatures were below the onset of thermal denaturation.

100

Page 117: Calorimetric and Microbiological Evaluation of Bacteria

It is also apparent from our results on E. coli that the onset of thermal denaturation of the

control samples is around 58oC. The plate count results for E. coli after partial scanning

to various temperatures with 4oC min-1 scanning rate clearly shows that a decrease in

viable counts occurs after a partial scan to 58.7oC (Table 3.1).

Although a first order rate equation is generally employed to describe the survival

curve of bacteria and to evaluate the processing parameters necessary to inactivate

microorganisms, it has been demonstrated that survival curves may show deviations from

the log-linear model displaying an initial lag time, a tail after a linear semi-logarithmic

survival curve, or other nonlinear semi-logarithmic behavior (Miles et al. 1994; Peleg,

2000). Calorimetric data can be fitted to mathematical models other than linear semi-

logarithmic models to determine the inactivation parameters of microorganisms.

This study focused on the utilization of calorimetric data to evaluate thermal

inactivation of bacteria under a linearly increasing temperature protocol. The basis for

thermodynamic study of microorganism inactivation is that the relevant initial and final

states can be defined and the energetic differences between these states can be measured

using calorimetric instrumentation. The amount of thermal energy (apparent

enthalpy, ∆H) associated with denaturation of cellular components before and after

application of heat treatment with linearly rising temperature using DSC was related to

the number of viable cells of E. coli K12 and was used to calculate the fraction of

surviving cells. An equation based on first order inactivation kinetics is used to calculate

D and z values from the fraction of surviving cells exposed to heat treatment using a

linearly increasing temperature protocol. The results suggest that the apparent enthalpy

101

Page 118: Calorimetric and Microbiological Evaluation of Bacteria

data obtained from DSC can be used to evaluate D and z values, as well as, the linear

temperature rise necessary to reduce a microbial population by a chosen factor. While

this calorimetric approach requires careful data collection and analysis, the significantly

shorter time required coupled with comparable accuracy, make this method competitive

with the plate count technique.

REFERENCES

Anderson, W.A., Hedges, N.D., Jones, M.V. and Cole, M.B. 1991. Thermal inactivation of Listeria monocytogenes studied in differential scanning calorimetry. J. Gen. Microbiol. 137:1419-1424. Andrieu, J., Laurent, M., Puaux, J.P. and Oshita, S. 1989. Thermal properties of unfrozen and frozen food gels determined by an automatic flash method. In Proceedings of the Fifth International Congress on Engineering and Food. ed. Spiess, W.E.L., Schubert, H. pp. 447 - 455. Elsevier Applied Science. Beaman, T.C., Greenamyre, J.T., Corner, T.R., Pankratz, H.S. and Gerhardt, P. 1981. Bacterial spore heat resistance correlated with water content, wet density, and protoplast/sporoplast volume ratio. J. Bacteriol. 150:870-877. Belliveau, B.H., Beaman, T.C., Pankratz, H.S. and Gerhardt, P. 1992. Heat killing of bacterial spores analyzed by differential scanning calorimeter. J. Bacteriol. 174:4463-4474.

102

Page 119: Calorimetric and Microbiological Evaluation of Bacteria

Bunning, V.K., Crawford, R.G., Tierney, J.T. and Peeler, J.T. 1990. Thermotolerance of Listeria monocytogenes and Salmonella typhimurium after sublethal heat shock. Appl. Environ. Microbiol. 56:3216-3219. Chambers, C.W., Tabak, H.H. and Kabler, P.W. 1957. Effect of Krebs cycle metabolites on the viability of Escherichia coli treated with heat and chlorine. J. Bacteriol. 73:77-84. Dega, C.A., Goepfert, J.M. and Amundson, C.H. 1972. Heat resistance of salmonellae in concentrated milk. Appl. Microbiol. 23:415-420. Deindoerfer, F.H. and Humphrey, A.E. 1959. Analytical method for calculating heat sterilization times. Appl. Microbiol. 7:256-264. Farber, J.B. and Brown, B.E. 1990. Effect of prior heat shock on heat resistance of Listeria monocytogenes in meat. Appl. Environ. Microbiol. 56:1584-1587. Goepfert, J.M., Iskander, I.K. and Amundson, C.H. 1970. Relation of the heat resistance of salmonellae to the water activity of the environment. Appl. Microbiol. 19:429-438. Gomez, J., Hilser, V.J., Xie, D. and Freire, E. 1995. The heat capacity of proteins. Proteins: Structure, Function, and Genetics 22:404-412. Hansen, N.H. and Riemann, H. 1963. Factors affecting the heat resistance of nonsporing organisms. J. Appl. Bacteriol. 26:314-333. Hoffman, H., Valdina, J., and Frank, M.E. 1966. Effect of high incubation temperature upon the cell wall of Escherichia coli. J. Bacteriol. 38:1635-1637. Hohne, G.W.H., Hemminger, W. and Flammersheim, H.-J. 1996. Differential Scanning Calorimetry: An Introduction for Practitioners. Berlin; New York: Springer-Verlag. Jay, J.M. 1996. High temperature food preservation and characteristics of thermophilic microorganisms. In Modern Food Microbiology 5th ed. pp. 347-369. New York: Chapman and Hall.

103

Page 120: Calorimetric and Microbiological Evaluation of Bacteria

Kaletunç, G. 2001. Thermal analysis of bacteria using differential scanning calorimetry. In Novel Process and Control Technologies in the Food Industry ed. Bozoglu, F., Deak, T. and Ray, B. pp. 227-235. Amsterdam: IOS press. Lemcke, R.M. and White, H.R. 1959. The heat resistance of Escherichia coli from cultures of different ages. J. Appl. Bacteriol. 22:193-201. Lepock, J.R., Frey, H.E. and Inniss, W.E. 1990. Thermal analysis of bacteria by differential scanning calorimetry: relationship to protein denaturation in situ to maximum growth temperature. Biochim. Biophis. Acta 1055:19-26. Linton, R.H., Webster, J.B., Pierson, M.D., Bishop, J.R. and Hackney, C.R. 1992. The effect of sublethal heat shock and growth atmosphere on the heat resistance of Listeria monocytogenes Scott A. J. Food Prot. 55:84-87. Mackey, B.M. and Derrick, C.M. 1987. Changes in the heat resistance of Salmonella typhimurium during heating at rising temperatures. Lett. Appl. Microbiol. 4:13-16. Mackey, B.M., Miles, C.A., Parsons, S.E. and Seymour, D.A. 1991. Thermal denaturation of whole cells and cell components of Escherichia coli examines by differential scanning calorimetry. J. Gen. Microbiol. 137:2361-2374. Mackey, B.M., Miles, C.A., Seymour, D.A. and Parsons, S.E. 1993. thermal denaturation and loss of viability in Escherichia coli and Bacillus stearothermophilus. Lett. Appl. Microbiol. 16:56-58. Mackey, B.M., Parsons, S.E., Miles, C.A. and Owen, R.J. 1988. The relationship between base composition of bacterial DNA and its intracellular melting temperature as determined by differential scanning calorimetry. J. Gen. Microbiol. 134:1185-1195. Miles, C.A. and Mackey, B.M. 1994. A mathematical analysis of microbial inactivation at linearly rising temperatures: calculation of the temperature rise needed to kill Listeria monocytogenes in different foods and methods for dynamic measurements of D and z values. J. Appl. Bacteriol. 77:14-20.

104

Page 121: Calorimetric and Microbiological Evaluation of Bacteria

Miles, C.A., Mackey, B.M. and Parsons, S.E. 1986. Differential scanning calorimetry of bacteria. J. Gen. Microbiol. 132:939-952. Murano, E.A. and Pierson, M.D. 1992. Effect of heat shock and growth atmosphere on the heat resistance of Escherichia coli O157:H7. J. Food Prot. 55:171-175. Niven, G.W., Miles, C.A. and Mackey, B.M. 1999. The effect of hydrostatic pressure on ribosome conformation in Escherichia coli: an in vivo study using differential scanning calorimetry. Microbiology 145:419-425. Nunes, R.V., Rhim, J.W. and Swartzel, K.R. 1991. Kinetic parameter evaluation with linearly increasing temperature profiles: internal methods. J. Food Sci. 56:1433-1437. Peleg, M. 1999. On calculating sterility in thermal and non-thermal preservation methods. Food Res. Int. 32: 271-278. Peleg, M. 2000. Microbial survival curves – the reality of flat “shoulders” and absolute thermal death time. Food Res. Int. 33:531-538. Reichart, O. 1979. A new experimental method for the determination of the heat destruction parameters of microorganisms. Acta Aliment. 8:131-155. Rhim, J.W., Nunes, R.V., Jones, V.A. and Swartzel, K.R. 1989. Determination of kinetic parameters using linearly rising temperature. J. Food Sci. 54:446-450. Strange, R.E. and Shon, M. 1964. Effect of thermal stress on viability and ribonucleic acid of Enterobacter aerogens in aqueous suspensions. J. Gen. Microbiol. 34:99-114. Stiles, M.E., Roth, L.A. and Clegg, L.F.L. 1973. Heat injury and resuscitation of Escherichia coli. Can. Inst. Food Sci. Technol J. 6:226-229. Teixeira, A. 1992. Thermal processing calculations. In Handbook of food processing ed Heldman D.R.and Lund, L.B. pp. 563-619. New York: Marcel Dekker.

105

Page 122: Calorimetric and Microbiological Evaluation of Bacteria

Teixeira, P., Castro, H., Mohacsi-Farkas, C. and Kirby, R. 1997. Identification of sites of injury in Lactobacillus bulgaricus during heat stress. J. Appl. Microbiol. 83(2):219-226. Thompson, D.R., Willardsen, R.R., Busta, F.F. and Allen, C.E. 1979a. Clostridium perfringens population dynamics during constant and rising temperatures in beef. J. Food Sci. 44:646-651.

Thompson, W.S., Busta, F.F., Thompson, D.R. and Allen, C.E. 1979b. Inactivation of salmonellae in autoclaved ground beef exposed to constantly rising temperatures. J. Food Prot. 42:410-415. Tomlins, R.I. and Ordal, Z.J. 1976. Thermal injury and inactivation in vegetative bacteria. In Inhibition and Inactivation of Vegetative Microbes ed. Skinner, F.A. and Hugo, W.B. pp.153-190. New York: Academic Press. Tsuchido, T., Hayashi, M., Takano, M. and Shibasaki, I. 1982. Alteration of thermal resistance of microorganisms in a non-isothermal heating process. J. Antibacterial Antifungal Agents 10:105-109. Van Impe, J.F., Nicolai, B.M., Martens, T., De Baerdemaeker, J. and Vandewalle, J. 1992. Dynamic mathematical model to predict microbial growth and inactivation during food processing. Appl. Environ. Microbiol. 58:2901-2909. Verrips, C.T. and Kwast, R.H. 1977. Heat resistance of Citrobacter freundii in media with various water activities. Eur. J. Appl. Microbiol. 4:225-231.

106

Page 123: Calorimetric and Microbiological Evaluation of Bacteria

CHAPTER 4

EVALUATION BY DIFFERENTIAL SCANNING CALORIMETRY OF THE

EFFECTS OF ETHANOL, NaCl, ACETIC ACID AND pH ON ESCHERICHIA COLI

ABSTRACT

The influence of chemical (acids, ethanol or NaCl) treatment on the cellular components

of Escherichia coli was evaluated using differential scanning calorimetry (DSC). Cell

viability was assessed using plate count. DSC thermograms showed that the transition of

ribosomal subunits in the cells was affected by mild chemical treatments where less than

0.4 log reduction of cell viability occurred. The thermal stability (Tm) for ribosomal

subunits denaturation decreases as chemical agent concentration increases. The total

apparent enthalpy (∆H) also decreases, mainly due to reduction of ribosomal subunit

peak as the concentration increases. Unlike in ethanol and NaCl treatments, the transition

of DNA was irreversibly affected after the treatment with inorganic (HCl) or organic

(CH3COOH) acid. The number of surviving cells received chemical treatments was

lower than that of non-treated cells after mild heat treatment (at 60, 62.5 and 65oC)

107

Page 124: Calorimetric and Microbiological Evaluation of Bacteria

indicating the conformational changes in cellular components by chemical treatments

may have sensitized bacteria to heat. Greater heat sensitivity for acid-treated cells might

be due to the chemically-induced irreversible damages on DNA as well as the damages

on ribosomal subunits.

Key Words: chemical treatment, differential scanning calorimetry, thermal stability,

ribosome, Escherichia coli

108

Page 125: Calorimetric and Microbiological Evaluation of Bacteria

INTRODUCTION

Thermal processing is the main choice of food preservation for inactivation of

pathogenic and spoilage bacteria in order to produce a safe product with enhanced shelf

life. High temperatures employed during thermal processing adversely affect texture,

flavor, nutrient value of food products. Therefore, mild heating in conjunction with

antimicrobial agents have been utilized to preserve the nutritional qualities while

maintaining extended shelf-life. This approach is known as hurdle technology (Abee and

Wouters, 1999). Hurdle technology is based on the reduced thermal resistance observed

for bacteria treated with chemical or physical means prior to or during heat treatment

(Karatzas et al., 2000; Leistner, 2000). The most commonly employed hurdles to reduce

the intensity of heat treatment include controlling water activity (aw), acidity, and use of

preservatives (Cameron et al., 1980; Adams et al., 1989; Membre, 1997; Casadei et al.,

2001). The physiological conditions of bacterial cells such as pH, aw, and ionic

interactions, which maintain metabolic reactions and homeostasis, are shown to be

affected by ethanol (Ingram and Buttke, 1984; Ingram, 1986), salt (Csonka, 1989; Poirier

et al., 1998) and acids (Abee and Wouters, 1999; Brul and Coote, 1999).

Differential scanning calorimetry (DSC) has been used to characterize the

conformational transitions of cellular components of bacteria during heat treatment

(Mackey et al., 1991; Kaletunç, 2001). Thermal stability of a cellular component can be

evaluated from the peak temperature of the corresponding transition from DSC

thermograms of whole cells (Miles et al., 1986; Anderson et al., 1991; Mackey et al.,

109

Page 126: Calorimetric and Microbiological Evaluation of Bacteria

1991; Bellivieau et al., 1992; Kaletunç, 2001; Lee and Kaletunç, 2002b). DSC has been

demonstrated to evaluate the effect of high hydrostatic pressure on bacteria by comparing

the DSC thermograms of pressure-treated cells with thermograms of untreated cells

(Niven et al., 1999; Alpas et al., 2003; Kaletunç et al., 2004). It has also been used to

investigate the survival of bacterial cells after heat treatment under linearly increasing

temperature using DSC instrument (Chapters 2 and 3).

To prevent the repair of the microbial homeostasis after food processing, chemical

hurdles of choice should target the various cellular components such as membrane,

nucleic acids, and proteins (Leistner, 1992; Leistner and Gorris, 1995). Therefore, to

achieve the optimal design of the hurdle technology, the investigations of the effect of

chemicals on the cellular components need to be investigated to elucidate the irreversible

changes in macromolecular components occurring of cells, which leads to cell injury and

death as a result of chemical treatments.

In the present study, the influence of acids, ethanol or NaCl treatment on the

cellular components of E. coli and the viability of the chemically treated cells during

subsequent heat treatment were evaluated using DSC. Plate count method was performed

to determine the effect of treatments on the viability of the E. coli.

110

Page 127: Calorimetric and Microbiological Evaluation of Bacteria

MATERIALS AND METHODS

Source and maintenance of organism

E. coli ATCC 14948 was obtained from the Culture Collection, Department of

Microbiology at the Ohio State University. A loopful of the organism was revived in 10

ml Trypticase soy broth supplemented with 0.3% (w/w) yeast extract (TSBYE) and

incubated at 37oC for 18 hours. The culture was stored frozen (-80oC) in 30% (v/v)

sterile glycerol. A loopful of this stock culture was transferred in 10 ml TSBYE and

incubated 10 hrs at 37oC before use.

DSC profiles of E. coli after chemical treatments

Revived E. coli culture was inoculated (1% v/v) into TSBYE. Cultures were

incubated at 37oC until the end of exponential growth phase, when the final concentration

of cells in a medium was 1.0 ± 0.1 x 109 cfu ml-1. E. coli cells in the broth was treated by

adding 6, 10, 12 and 15% (vol/vol) ethanol (95%, Pharmco Inc., Brookfield, CT), 6 and

10% (wt/wt) sodium chloride (NaCl, Sigma), pH 3.0 and 4.0 using hydrochloric acid

(36%, wt/vol; Fisher Scientific), or 0.2, 0.5, 1.0 and 2.0% acetic acid (glacial, Fisher

Scientific). After 1 h of treatment at 37oC, a portion (1 ml) of the treated and untreated

(control) cultures were pour plated into Trypticase soy agar to determine viable cell

counts in media. Remained cultures were centrifuged at 10 000 g for 10 min at 4oC.

Supernatants were discarded. Pellets were washed with 100 ml of sterile distilled water

and centrifuged again before transferring into DSC crucibles.

111

Page 128: Calorimetric and Microbiological Evaluation of Bacteria

Calorimetry

A differential scanning calorimeter (DSC 111, Setaram, Lyon, France) was used to

record the thermograms of the untreated and treated E. coli cells. All DSC measurements

were conducted using fluid-tight, stainless steel crucibles. A DSC run was performed

with unsealed-empty sample and reference crucibles to measure the instrument baseline.

Pellets of cells were weighed (56 ± 0.3 mg wet weight) and carefully transferred into the

sample crucible. The water content of the pellets was determined using freeze dryer

(Freezone 4.5, Freeze dry system, Model 77510, Labconco, MO) as 80 % on wet basis.

For each DSC run, reference crucible was filled with ~45 µl (~80 % of sample wt) of

pure distilled water. Both crucibles were sealed using aluminum rings and covered with

screw caps. The sealed crucibles were refrigerated at 4oC until used for DSC. The

sample and reference crucibles were placed in the DSC and equilibrated at 1oC using

liquid nitrogen cooling system.

Samples were heated in the DSC at 3oC min-1 from 1 to 150oC. After heating,

samples were rapidly cooled by liquid nitrogen and rescanned to observe the reversibility

of thermograms. Samples were re-weighed after measurements to check for loss of mass

during heating and the result of samples showing signs of leakage were discarded.

Viability after heat treatment in DSC following chemical treatment

Weighed cell pellets (~70 mg wet weight) of chemically-treated cultures and control

were carefully transferred into sterile DSC empty sample crucibles using sterile loops.

112

Page 129: Calorimetric and Microbiological Evaluation of Bacteria

Each crucible was capped (not sealed) using aluminum ring and screwed cap. Empty

reference crucible was filled with ~56 µl (~80% of sample wt) of pure distilled water.

The capped crucibles were kept in refrigerator (4oC) until used for DSC. Pellets in

crucible were heat-treated to 60, 62.5, 65oC with 3oC/min heating rate using DSC

instrument. After cooling, a portion (50 mg) of the heated pellet from each crucible was

transferred to a (1.5 ml) sterile polyethylene tube using a sterile loop. Sterile peptone

water was added to make a final volume of 1 ml with 1/20 (w/v) ratio. After careful

suspension in the tube, the cells were serially diluted and plated onto Trypticase soy agar

to determine viable counts. After 36 hours incubation at 37oC, viable counts of each

sample were obtained by calculation of dilution ratios. The level of the lowest detection

was 2 x 101 cfu g-1 in pellet.

DSC data analysis

DSC thermograms were corrected for differences in the empty crucibles by

subtracting an empty crucible baseline. Total heats corresponding to the endothermic

peaks of whole cells (enthalpy, J g-1) between approximately 40 and 130oC were

determined by integrating the temperature vs. heat flow curve using software provided by

the instrument manufacture. A curved baseline taking into account the variation in heat

capacity before and after the transition passing through three designated points on the

thermogram was used to calculate the apparent enthalpy of whole cells. Data points at

three temperatures were selected to determine the baselines for all DSC curves. The

initial temperature point was on the pre-transition baseline (40oC). The mid-point was

113

Page 130: Calorimetric and Microbiological Evaluation of Bacteria

selected at a temperature below the onset of the final peak which corresponds to

transitions in the cell envelope (~108oC). The final point was on the post-transition

baseline (130oC).

Growth of E. coli cells to the end of exponential growth stage

Chemical shock by adding acids, ethanol or NaCl for 1 h Plate counting forviability

a

DSC for calorimetric dat

Centrifugation to obtain cell pellets

114

Analysis of the survivability of chemically- treated E. coli

cells in heat treatment

Analysis of the effects ofchemicals on E. coli cells

s

Figure 4.1. Experimental scheme of calorimetric and microbial analysi

Heat-treatment in DSC

Plate counting

Page 131: Calorimetric and Microbiological Evaluation of Bacteria

RESULTS

Effect of ethanol on thermal transitions and viability of E. coli

Figure 4.2 shows the DSC thermogram for untreated (A) and ethanol treated (B, C, D,

and E) E. coli pellets. Individual endothermic peaks (a-d) were shown to be associated

with components such as ribosomal subunits (a1, a2 and a3), DNA (b), DNA with cell wall

(c), and outer membrane of Gram-negative organisms (d) (Mackey et al, 1991; Lee and

Kaletunç, 2002b). Thermograms in Figure 4.2 show that ribosomal subunits, which are

composed of RNA and ribosomal proteins, in the E. coli cells were affected by ethanol

treatment, while other transitions remain same. The onset temperature and thermal

stability for ribosomal subunits denaturation decreased as ethanol concentration increase

(Table 4.1). The total apparent enthalpy also decreased, mainly due to reduction of

ribosomal subunit peak as ethanol concentration increased. After treatment with 12%

ethanol, the transition temperature and enthalpy of the peaks due to DNA (peak b) and

DNA with cell wall component (peak c) were markedly decreased. However, ethanol

concentration did not affect the plate count data (<1 log reduction) up to 12% ethanol

addition. The decrease in transition temperature and enthalpy associated with outer

membrane component (peak d) was apparent after 15% ethanol treatment (Table 4.1).

115

Page 132: Calorimetric and Microbiological Evaluation of Bacteria

Transition temperature (Tm,, oC) Enthalpy (J/g) Concentration in broth (v/v)

Viability [log (N/N0)]

Total Apparent enthalpy (∆H, J/g)

Onset temperature

(oC) Peak a2 Peak a3 Peak b Peak c Peak d Peak b+c Peak d

Non-treatment - 4.19 54.6 69.8 77.9 93.8 102.4 117.7 0.47 0.19

Ethanol

6% 0.01 4.12 47.1 69.4 76.7 93.2 101.5 117.6 0.47 0.19

10% 0.09 3.80 43.6 67.5 75.7 93.2 101.9 117.5 0.43 0.19

12% 0.35 3.76 42.2 64.0 - 93.3 101.3 116.7 0.41 0.11

15% 4.79 2.85 36.4 62.4 - 92.8 100.9 115.9 0.38 0.11

NaCl 6% 0.24 4.21 46.6 67.2 76.6 93.6 101.3 117.4 0.43 0.17

10% 1.15 4.20 46.1 67.3 76.7 94.7 101.4 117.7 0.44 0.16

HCl pH 4.0 0.15 3.20 37.8 65.7 - 100.3 117.6 0.32 0.15

pH 3.0 0.22 2.61 33.2 61.1 - 93.9 116.2 0.39 0.14

Acetic acid 0.2% (pH 4.9) 0.09 4.15 42.7 67.8 71.7 - 93.9 102.4 117.7 0.39 0.13

0.5% (pH 4.2) 0.21 3.89 38.7 66.1 70.6 - 101.6 117.6 0.29 0.14

1.0% (pH 3.9) 0.38 3.01 38.3 60.2 68.2 - 97.9 117.7 0.36 0.10

2.0% (pH 2.8) 5.31 0.82 - - - 89.8 116.8 0.28 0.12

116

Table 4.1. Effects of chemicals on viability and DSC transitions of E. coli.

Page 133: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120 140

a2

a1

a3 bc d

A

B

C

Heat flow

0.4 mW

D

E

Temperature (oC)

Figure 4.2. DSC thermogram of E. coli pellet after ethanol treatment. Control (A), ethanol concentration in the treatment: 6% (B), 10% (C), 12% (D), 15% (E). Thermograms are offset for clarity.

117

Page 134: Calorimetric and Microbiological Evaluation of Bacteria

Effect of salt on thermal transitions and viability of E. coli

Sodium chloride (NaCl) was added to assess the effect of salt on E. coli cells in

growth medium. Except for the transitions in ribosomal subunits, no apparent differences

of major peaks were found between thermograms of control and salt-added cells (Fig.

4.3). Salt treatments resulted in absence of peak a1 and changes in transition

temperatures of peak a2 and a3 while total apparent enthalpy remain unchanged (Fig. 4.3

and Table 4.1). After 6% NaCl (wt/wt) treatment, the transition temperatures of

ribosomal peaks were lowered by 2oC for peak a2 and 1oC for peak a3. In addition,

decrease in onset temperature (~8oC) of the peak a2 is shown in thermograms of the

treated cells. NaCl treatment of cells to higher concentration (10%) resulted in ~1-log10

unit reduction (Table 4.1); however, except for increase in transition temperature (~1oC)

of DNA peak (peak b), the increase in NaCl concentration to 10% did not affect the

transitions of cellular components in DSC thermograms of the salt-treated cells (Fig. 4.3

thermograms B and C, Table 4.1).

Effect of HCl on thermal transitions and viability of E. coli

HCl was used to adjust the pH of environment to 4 and 3. The thermograms (B and

C) in Figure 4.4 show that HCl affected all of the cellular components in DSC

thermogram at both pH 4 and 3. The substantial decrease in onset temperature and

ribosomal transition (a2) temperature of the cells were observed as pH of the medium

decreased. Specifically, the temperature of a2 decreased 4oC after pH 4 treatment and

9oC after pH 3 treatment.

118

Page 135: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120 140

a2

a1

a3b

c d

A

B

C

Heat flow

0.2 mW

Temperature (oC)

Figure 4.3. DSC thermogram of E. coli pellet after NaCl treatment. Control (A), NaCl concentration in the treatment: 6% (B), 10% (C). Thermograms are offset for clarity.

119

Page 136: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120 140

a2

a1

a3 b

c d

A

B

C

Heat flow

0.2 mW

Temperature (oC)

Figure 4.4. DSC thermogram of E. coli pellet after inorganic acid (HCl) treatment. Control (A), pH in the treatment: pH 4 (B), pH 3 (C). Thermograms are offset for clarity.

120

Page 137: Calorimetric and Microbiological Evaluation of Bacteria

The acid treatment resulted in the absence of peak b in thermogram B and shift of

peak c in thermogram C (Fig. 4.4 and Table 4.1). In addition, marked reductions in the

total apparent enthalpy occurred after the acid treatment at pH 4 (~24%) and at pH 3

(~38%). Although noticeable changes in peaks for various cellular components were

observed, loss of viability prior to DSC was minimal among control, pH 4- and pH 3-

treated cultures (<0.3 log reductions) (Table 4.1).

Effect of acetic acid on thermal transitions and viability of E. coli

Acetic acid (CH3COOH) was used to evaluate the effect of organic acid on E. coli.

Peak a2 appeared to include two overlapping endothermic transitions for acetic acid

treated cells (Fig. 4.5). The transition temperature for a2 peak decreased as acetic acid

concentration increased (Table 4.1). The absence of peak b in thermogram C and the

shifts of peak c in thermograms D,E were observed after acetic acid treatments (Fig 4.5).

Total apparent enthalpy decreased with increasing acetic acid concentration. The

viability loss of acetic acid treated cells prior to DSC measurement did not differ from

control cells until concentration of acetic acid was 1% (<0.4 log unit reduction of

viability) in the medium. When the medium acidity was dropped to pH 2.8 by 2% acetic

acid concentration, the viable count of the culture declined 5-log cycles (Table 4.1). The

transition due to ribosomal subunits of cells was absent after treatment with 2% acetic

acid (Fig. 4.5 Thermogram E).

121

Page 138: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120 140

a2

a1

a3 b

c d

A

B

C

Heat flow

0.4 mW

D

E

Temperature (oC)

Figure 4.5. DSC thermogram of E. coli pellet after organic acid (acetic acid) treatment. Control (A), acetic acid concentration (v/v) in the treatment: 0.2% (B), 0.5% (C), 1.0% (D) and 2.0% (E). Thermograms are offset for clarity.

122

Page 139: Calorimetric and Microbiological Evaluation of Bacteria

Effect of heat treatment on viability of chemically treated E. coli cells

E. coli cultures exposed to 12% ethanol, 6% NaCl, pH 4, pH 3, 0.5% acetic acid,

and 1% acetic acid treatments for 1 hr were heat treated in DSC to 60, 62.5 and 65oC.

Chemical treatments alone caused 0.2 to 0.4 log reductions in viability of E. coli cells.

Figure 5 compares viability of E. coli cells after chemical and heat treatment with the

cells after heat treatment alone. Application of chemical stresses such as HCl (pH 3) and

acetic acid (1%) prior to heat treatment markedly (log 3.7 for HCl and 3.8 for acetic acid)

reduced the viability of E. coli cells at the lowest heat treatment temperature of 60oC

compared to that of untreated (0.1 log) and ethanol (1.0 log) or NaCl (0.5 log) treated

cells. For 1.0% acetic acid treated cells, ~8 log reduction after 62.5oC treatment was

achieved. Viable cells were not observed after 65oC treatment for cells treated with 1%

acetic acid or by HCl at pH 3. After heat treatment in DSC at 65oC, a log reduction of

7.9 for 0.5% acetic acid treatment, 6.5 for HCl treatment at pH 4, 6.0 for 12% ethanol

treatment, and 5.7 for 6% salt treatment were observed while the control cells decreased

4.7 log units.

123

Page 140: Calorimetric and Microbiological Evaluation of Bacteria

2

3

4

5

6

7

8

9

10

11

12

13

Before DSC 60 62.5 65

Control

Ethanol 12%

Salt 6%

pH4 usingHCl

pH3 usingHCl

Acetic acid0.5%

Acetic acid1.0%

Log

(cfu

/g) i

n pe

llet

Temperature (oC)

Figure 4.6. Survival of untreated and chemically treated E. coli after heat treatment under linearly increasing temperature.

124

Page 141: Calorimetric and Microbiological Evaluation of Bacteria

DISCUSSION

The studies on effect of chemicals on bacteria have been focused on the

cytoplasmic membrane with minimal effort on the influence of chemical treatment on

intracellular structures such as ribosomes, nucleic acids and proteins. Modifications of

cell permeability and leakage of ions, failure of proton motive force in which ATP is

synthesized, and inhibition of membrane-associated enzyme activity were considered as

potential damages to bacteria due to interaction of chemicals with the membrane (Weitzel

et al., 1987; Maillard, 2002).

Chemicals, such as ethylenediamine tetraacetic acid and hydrogen peroxide (H2O2),

were reported to chelate metal ions associated with ribosomes leading to dissociation of

larger ribosomal subunits to smaller subunits or disintegration of each subunit (Nakamura

and Tamaoki, 1968). Except for above chelators and specific antibiotics such as

erythromycin (on 50S subunit) and tetracycline (on 30S subunit), ribosomal subunits

have not been considered as major target sites for most chemicals (Hugo, 1999; Maillard,

2002). However, the marked reductions in size and transition temperature of ribosomal

subunit transitions in DSC thermograms of ethanol (15%), HCl (<pH 4), and acetic acid

(>0.5%) treated E. coli cells in this study indicate ribosomal structure was irreversibly

altered by high concentrations of these chemicals. The disappearance of first peak (a1),

which is proposed to be the denaturation of 30S ribosomal subunit in E. coli (Mackey et

al., 1991), in thermogram of all of chemically treated E. coli whole cells in present DSC

study indicates the smaller and flexible ribosomal subunit might be very susceptible

125

Page 142: Calorimetric and Microbiological Evaluation of Bacteria

component for chemical treatment. In Chapter 2, I discussed that ribosome thermal

stability is altered by low pH condition during cell growth as well as loss of intracellular

metals like Mg2+ ions during heat treatment.

The pH of inside cell was reported to be lower during chemical treatment due to

accumulation of excessive H+ ions in the cell (Maillard, 2002). Chemicals containing

hydrocarbons such as ethanol and acetic acid can induce membrane damage leads to the

disintegration of proton motive force which synthesizes ATP to maintain H+ level of

microbial cell (Quintas et al., 2000; Minamino et al., 2003). In addition, a recent study

showed that monovalent ions in NaCl generate the dissipation of the proton motive force

in bacterial membrane (Minahk and Morero, 2003). Therefore, the effect of chemical

treatments on ribosomal subunits in present study might be result from the interaction of

chemicals with ribosomal subunits and/or the interaction of chemicals with cytoplasmic

membrane. The occurrences of above interactions by a chemical might be concentration

dependent.

DSC thermograms in Figure 1 show that the shape and size of ribosome transitions

(peaks a1,2,3) changed after various levels of ethanol treatments; while, other transitions

remain unchanged. There were minimal changes in viability and the apparent enthalpy

(J/g) in DSC thermograms of the treated E. coli cells up to 12% ethanol concentration,

while those values were considerably reduced in 15% ethanol treated cells (Table 4.1). In

the studies of E. coli (Chapters 2 and 3), heat-treatment under linearly increasing

temperature slightly decreased the apparent enthalpy value of the whole cell DSC until

the temperature reached to a certain level where a log reduction in viable counts

126

Page 143: Calorimetric and Microbiological Evaluation of Bacteria

occurred. Ethanol is known to have a similar effect to heat stress in initiating response

mechanisms by the microbial membrane (Piper, 1995; Casadei et al., 2001). Both

stresses cause induction of membrane-bound heat shock proteins which enhance the

proton efflux to help homeostasis of the increased membrane permeability that result

from the stresses (McElhaney, 1985; Piper, 1995). Therefore, the cell viability might be

protected by above response mechanism in the presence of ethanol concentration up to

12%. Because ethanol generally penetrate cytoplasmic membrane via lipid-bilayer

portion, the increase in the length of the fatty acid chains in the membrane of

microorganisms, which provide a more fortified hydrophobic barrier than normal fatty

acid chain, are also considered defense mechanism against ethanol (Ingram and Buttke,

1984; Broadbent and Lin, 1999). However, since the above result was observed after

growth of cells in the presence of ethanol, further study on the change in fatty acid

composition in the same condition as this study will be needed to verify the occurrence of

the mechanism. In present study, E. coli cells in the end of logarithmic growth phase

were treated by ethanol for 1 hour prior to DSC.

DSC thermograms show that there is no apparent change in transition of cellular

components of E. coli with the exception of ribosomal subunits after NaCl treatments

(Fig. 4.3 and Table 4.1). The concentrations of NaCl higher than 10% did not cause

further reduction of the ribosomal transition temperature and area in E. coli cells in spite

of decreased viability (Table 4.1). High concentration of intracellular cations has been

reported to protect DNA from denaturation at high temperatures because removal of

counterions (positively charged mono- or di-valent ions), which interact with negatively

127

Page 144: Calorimetric and Microbiological Evaluation of Bacteria

charged phosphate backbones of the double stranded DNA, is increasingly unfavorable

due to surrounded bulk cations (Manning et al., 1978; Record et al., 1978). Therefore,

slight increase in the transition temperature of peak c in the thermogram of 10% NaCl

treated cells indicating the thermal stability of DNA structure might be enhanced by bulk

Na+ ions which surround the DNA.

E. coli cells are known to accumulate “compatible solutes” such as betaine and

trehalose which can increase internal osmotic pressure against a hyperosmotic shock

without interfering functions of cell components (Gutierrez et al., 1995; Abee and

Wouters, 1999). In a study on NaCl sensitivity of E. coli, Nakamura et al. (1992)

reported that the addition of NaCl to suspensions of mutant cells that do not have sodium

efflux system results in the loss of viability while wild-type cells show resistance.

Therefore, it is possible that E. coli cell inactivation at the higher concentration of NaCl

may be mainly due to leakage of more Na+ ions into the cell and inhibition of certain

regulatory pathway by interacting with key enzymes and proteins as well as with ions. It

is reported that excessive Na+ cause the failure of respiratory electron transport system

which is important factor for viability of cell (Allakhverdiev et al., 1999). One log unit

viability loss observed with addition of 10% NaCl (w/v) did not cause any change in the

total apparent enthalpy requirement for the cell death in subsequent heat treatment in

DSC in spite of thermal stability loss of 2oC in ribosomal subunits. The result may

indicate that NaCl-induced changes may be entropy driven rather than enthalpy driven.

Figure 4.4 and 4.5 reveals that the peaks attributed to the ribosomal subunits in the

DSC thermogram greatly shifted to lower temperature when the E. coli was subjected to

128

Page 145: Calorimetric and Microbiological Evaluation of Bacteria

HCl or acetic acid treatment. The considerable reduction of total area under the

thermogram peaks (apparent enthalpy) of the HCl and acetic acid treated cells includes

the absence or size reduction of peaks associated with DNA transition (peaks b and c,

which remained unchanged after ethanol or salt treatment) as well as the decrease in area

associated with ribosomal subunit transitions. The absence of peak b in the thermograms

of Figures 4.4B and 4.5C possibly indicates that double stranded DNA structure was

affected by HCl (pH 4) or 0.5% acetic acid (pH 4.2) treatment. Exposure of bacterial

cells to acidic pH was known to induce DNA damage due to the reduction of divalent

ions such as Mg2+ and Ca2+ which stabilize DNA molecules (Hickey and Hirshfield,

1990; Juenja et al., 1995). Peak c, proposed to be due to denaturations of cell wall and

part of DNA (Mackey et al., 1991), becomes apparent due to the obliteration of peak b

which obscured peak c in the thermogram of untreated E. coli cells. The peak with 94oC

in Figure 4.4C and the peaks with 98oC (in Fig. 4.5D) or 90oC (in Fig. 4.5E) might

indicate that the thermal stability of the component(s) related in peak c was (were)

reduced after acid treatments. The possible reasons of the difference in thermal stability

of two peaks were from the following differences between two DNAs: the structure of

DNA (chromosome or plasmid), interaction with components such as cationic proteins

and polyamines in the cell (Worcel and Burgi, 1972; Flink and Pettijohn, 1975).

The survivability of the chemically treated cells during heat treatments was much

lower than those of untreated cells (Fig. 4.6). The result suggests that there is

relationship between reduction in the ribosomal subunits transition in DSC thermogram

and increase in heat sensitivity of chemically treated cells. The result also agrees with the

129

Page 146: Calorimetric and Microbiological Evaluation of Bacteria

findings of previous DSC studies that heat resistance of bacteria is strongly related to the

onset temperature and the thermal stability of the main ribosomal subunit peak (Miles et

al., 1986; Mohacsi-Farkas et al., 1994; Lee and Kaletunç, 2002b). In present study, DSC

data show the reductions in the onset temperatures (by 8~12oC) and the transition

temperatures (peaks a2, by 3~10oC) of major ribosomal subunit occurred in the

thermograms of chemically treated cells that were used in heat treatments (Table 4.1).

Among chemically treated cells, the onset temperature and thermal stability of ribosome

were much lower in HCl- and acetic acid-treated cells than in ethanol- and NaCl-treated

cells. In addition, the reduction of total apparent enthalpy was greater in those acid

treated cells due to greater size reductions in ribosomal and DNA transitions of their

thermograms. Above results might be correlated to higher thermal sensitivity of acid

treated cells in subsequent heat treatment. Among acid treated cells, the highest

inactivation rate during the heat treatment up to 65oC for 1% acetic acid treated cells, as

judged by the steepest slope of the plots in Figure 4.6, indicates that the hurdle effect of

acetic acid is greater than that of HCl.

In conclusion, DSC thermograms for E. coli revealed conformational changes in

cellular components after chemical treatments. Mild treatments affect the thermal

stability of ribosomal subunits in the cell, thereby increasing the sensitivity of bacteria to

heat treatment. The heat sensitivity is greater for acid-treated cells because more cellular

components were irreversibly affected after the treatments. These hurdle effects should

be considered when current thermal processing technologies are modified.

130

Page 147: Calorimetric and Microbiological Evaluation of Bacteria

REFERENCES Abee, T. and Wouters, J.A. 1999. Microbial stress response in minimal processing. Int. J. Food Microbiol. 50:65-91. Adams, M.R., O’Brien, P.J. and Taylor, G.T. 1989. Effect of ethanol content of beer on the heat resistance of a spoilage Lactobacillus. J. Appl. Bacteriol. 66:491-495. Allakhverdiev, S.I., Nishiyama, Y., Suzuki, I., Tasaka. Y. and Murata, N. 1999. Genetic engineering of the unsaturation of fatty acids in membrane lipids alters the tolerance of Synechocystis to salt stress. Proc. Natl. Acad. Sci. USA 96: 5862 5867. Alpas, H., Lee, J., Bozoglu, F. and Kaletunç, G. 2002. Differential scanning calorimetry of pressure-resistant and pressure-sensitive strains of Staphylococcus aureus and Escherichia coli O157:H7. Int. J. Food Microbiol. 87:229-237. Anderson, W.A., Hedges, N.D., Jones, M.V. and Cole, M.B. 1991. Thermal inactivation of Listeria monocytogenes studied in differential scanning calorimetry. J. Gen. Microbiol. 137:1419-1424. Bearson, B.L., Wilson, L. and Foster, J.W. 1998. A low pH-inducible, PhoPQ-dependent acid tolerance response protects Salmonella typhimurium against inorganic acid stress. J. Bacteriol. 180:2409-2417. Belliveau, B.H., Beaman, T.C., Pankratz, H.S. and Gerhardt, P. 1992. Heat killing of bacterial spores analyzed by differential scanning calorimeter. J. Bacteriol. 174:4463-4474. Bowler, K. and Manning, R. 1994. Membrane as the critical targets in cellular heat injury and resistance adaptation. In Temperature adaptation of biological membranes ed. Cossins, A.R. pp. 185-203. Portland Press: London. Broadbent, J.R. and Lin, C. 1999. Effect of heat shock or cold shock treatment on the resistance of Lactococcus lactis to freezing and lyophilization. Cryobiology 39:88-102.

131

Page 148: Calorimetric and Microbiological Evaluation of Bacteria

Brul, S. and Coote, P. 1999. Preservative agents in foods – Mode of action and microbial resistance mechanisms. Int. J. Food Microbiol. 50:1-17. Cameron, M.S., Leonard, S.J. and Barret, E.L. 1980. Effect of moderately acidic pH on heat resistance of Clostridium sporogenes spores in phosphate buffer and in buffered pea puree. Appl. Environ. Microbiol. 39:943-949. Casadei, M.A., Ingram, I., Hitchings, E., Archer, J. and Gaze, J.E. 2001. Heat resistance of Bacillus cereus, Salmonella typhimurium and Lactobacillus delbrueckii in relation to pH and ethanol. Int. J. Food Microbiol. 63:125-134. Csonka, L.N. 1989. Physiological and genetic response of bacteria to osmotic stress. Microbiol. Rev. 53:121-147. Flink, I. and Pettijohn, D.E. 1975. Polyamines stabilize DNA folds. Nature London 253:62-63. Gutierrez, C., Abee, T. and Booth, I.R. 1995. Physiology of the osmotic stress response in microorganisms. Int. J. Food Microbiol. 28:233-244. Hickey, E.W. and Hirshfield, I.N. 1990. Low-pH-induced effects on patterns of protein synthesis and on internal pH in Escherichia coli and Salmonella typhimurium. Appl. Environ. Microbiol. 56:1038-1045. Hill, C., O’Driscoll, B. and Booth, I. 1995. Acid adaptation and food poisoning microorganisms. Int. Food Microbiol. 28:245-254. Hisao, C. and Siebert, K.J. 1999. Modeling the inhibitory effects of organic acids on bacteria. Int. J. Food Microbiol. 47:189-201. Hou, M-H., Lin, S-B., Yuann, J-M.P., Lin, W-C., Wang, A.H.-J. and Kan, L-S. 2001. Effects of polyamines on the thermal stability and formation kinetics of DNA duplexes with abnormal structure. Nucleic Acids Res. 24:5121-5128.

132

Page 149: Calorimetric and Microbiological Evaluation of Bacteria

Hugo, W.B. 1999. Disinfection Mechanisms, In principles and practice of disinfection, preservation, and sterilization, 3rd ed. ed. Russell, A.D., Hugo, W.B. and Ayliffe, G.A.J. pp. 258-283. Blackwell Scientific Publication: Oxford. Ingram, L.O. 1986. Microbial tolerance to alcohols: Role of the cell membrane. Trends Biotechnol. 4:40-44. Ingram, L.O. 1990. Ethanol tolerance in bacteria. Crit. Rev. Biotechnol. 9:305-319. Ingram, L.O. and Buttke, T. 1984. Effects of alcohols on microorganisms. Adv. Microbial Physiol. 25:253-300. Juneja, V.K., Marmer, B.S., Phillips, J.G. and Miller, A.J. 1995. Influence of the intrinsic properties of food on thermal inactivation of spores of nonproteolytic Clostridium botulinum: development of a predictive model. J. Food Safety. 15:349-364. Kaletunç, G. 2001. Thermal analysis of bacteria using differential scanning calorimetry. In Novel Process and Control Technologies in the Food Industry ed. Bozoglu, F., Deak, T. and Ray, B. pp. 227-235. Amsterdam: IOS press. Kaletunç, G., Lee, J., Alpas, H. and Bozoglu, F. 2004. Evaluation of structural changes induced by high hydrostatic pressure in Leuconostoc mesenteroides. Appl. Environ. Microbiol. 70:1116-1122. Karatzas, A.K., Bennik, M.H.J., Smid, E.J. and Kets, E.P.W. 2000. Combined action of S-carbone and mild heat treatment on Listeria monocytogenes Scott A. J. Appl. Microbiol. 89:296-301. Lee, J. and Kaletunç, G. 2002a. Calorimetric determination of inactivation parameters of microorganisms. J. Appl. Microbiol. 93:178-189. Lee, J. and Kaletunç, G. 2002b. Evaluation by differential scanning calorimetry of the heat inactivation of Escherichia coli and Lactobacillus plantarum. Appl. Environ. Microbiol. 68:5379-5386.

133

Page 150: Calorimetric and Microbiological Evaluation of Bacteria

Leistner, L. 1992. Food preservation by combined methods. Food Res. Int. 25:151-158. Leistner, L. 2000. Basic aspects of food preservation by hurdle technology. Int. J. Food. Microbiol. 55:181-186. Leistner, L. and Gorris, L.G.M. 1995. Food preservation and hurdle technology. Trends Food Sci. Technol. 6:41-46. Mackey, B.M., Miles, C.A., Parsons, S.E., and Seymour, D.A. 1991. Thermal denaturation of whole cells and cell components of Escherichia coli examined by differential scanning calorimetry. J. Gen. Microbiol. 137:2361-2374. Maillard, J,-Y. 2002. Bacterial target sites for biocide action. J. Appl. Microbiol. Sym. Suppl. 92:16S-27S. Manning, G.S. 1978. The molecular theory of polyelectrolyte solutions with applications to the electrostatic properties of polynucleotides. Q Rev Biophys. 11:179-246. McElhaney, R.N. 1985. The effect of membrane lipids on permeability and transport in prokaryotes. In Structure and properties of cell membranes ed. Benga, G. p. 20. vol. 2. CRC Press, Boca Raton, FL. Membre, J.-M., Majchrzak, V. and Jolly, I. 1997. Effects of temperature, pH, glucose and citric acid on the inactivation of Salmonella typhimurium in reduced calorie mayonnaise. J. Food Prot. 60:1497-1501. Miles, C.A., Mackey, B.M. and Parsons, S.E. 1986. Differential scanning calorimetry of bacteria. J. Gen. Microbiol. 132:939-952. Minahk, C.J. and Morero, R.D. 2003. Inhibition of entercin CRL35 antibiotic activity by mono- and divalent ions. Lett. Appl. Microbiol. 37:374-379.

134

Page 151: Calorimetric and Microbiological Evaluation of Bacteria

Minamino, T., Imae, Y., Oosawa, F., Kobayashi, Y. and Oosawa, K. 2003. Effect of intracellular pH on rotational speed of bacterial flagellar motors. J. Bacteriol. 185:1190-1194. Mohacsi-Farkas, Cs., Farkas, J. and Simon, A. 1994. Thermal denaturation of bacterial cells examined by differential scanning calorimetry. Acta Aliment. 23:157-168. Munns, R., Greenway, H., Setter, T.L. and Kuo, J. 1983. Tugor pressure, volumetric elastic modulus, osmotic volume and ultrastructure of Chlorella emersonii grown at high and low external NaCl. J. Exp. Bot. 34:144-155. Nakamura, H., Hase, A. and Funatsuki, K. 1992. Biological actions of acridines: Salt sensitivity of Escherichia coli which is determined by the acrA gene. Memo. Konan Univ. Sci. Ser. 39:213-223. Nakamura, K. and Tamaoki, T. 1968. Reversible dissociation of Escherichia coli ribosomes by hydrogen peroxide. Biochim. Biophys. Acta 161:368-376. Niven, G.W., Miles, C.A. and Mackey, B.M. 1999. The effect of hydrostatic pressure on ribosome conformation in Escherichia coli: an in vivo study using differential scanning calorimetry. Microbiology 145:419-425. Poirier, I., Marechal, P.-A., Evrard. C. and Gervais, P. 1998. Escherichia coli and Lactobacillus plantarum responses to osmotic stress. Appl. Microbiol. Biotechnol. 50:704-709. Piper, P.W. 1995. The heat shock and ethanol stress responses of yeast exhibit extensive similarity and functional overlap. FEMS Microbiol. Lett. 134:121-127. Quintas, C., Lima-Costa, E. and Loureiro-Dias, M.C. 2000. The effect of ethanol on the plasma membrane permeability of spoilage yeasts. Food Technol. Biotechnol. 38:47-51. Record, M. Th., Jr. 1975. Effects of Na+ and Mg++ ions on the helix-coil transition of DNA. Biopolymers 14:2137-2158.

135

Page 152: Calorimetric and Microbiological Evaluation of Bacteria

Record, M. Th., Jr., Anderson, C.F. and Lohman, T.M. 1978. Thermodynamic analysis of ion effects on the binding and conformational equilibria of proteins and nucleic acids: the roles of ion association or release, screening, and ion effects on water activity. Q Rev Biophys. 11:103-178. Shadbolt, C.T., Ross, T. and McMeekin, T.A. 1999. Nonthermal death of Escherichia coli. Int. J. Food. Microbiol. 49:129-138. Salton, M.R.J. 1963. the relationship between the nature of the cell wall and the Gram strain. J. Gen. Microbiol. 30:233-235. Weitzel, G., Pilatus, U. and Rensing, L. 1987. The cytoplasmic pH, ATP content and total protein synthesis rate during heat-shock protein inducing treatments in yeast. Exp. Cell Res. 170:64-79. Worcel, A. and Burgi, E. 1972. On the structure of the folded chromosome of Escherichia coli. J. Mol. Biol. 71:127-147.

136

Page 153: Calorimetric and Microbiological Evaluation of Bacteria

CHAPTER 5

EVALUATION OF VIABILITY AND STRUCTURAL CHANGES INDUCED BY

HIGH HYDROSTATIC PRESSURE IN ESCHEICHIA COLI

ABSTRACT

The effect of high hydrostatic pressure (HHP) on the cell structures of Escherichia coli

was determined using differential scanning calorimetry (DSC) and electron microscopy

(EM). The changes in the structures were compared with viability. The cells were

pressurized between 200 and 700 MPa at 35oC for 5 min. DSC studies of whole cells

showed a decrease in apparent enthalpy above 200 MPa pressure treatments. The major

contribution to enthalpy decrease was due to reduction in the transition attributed to the

denaturation of ribosomes. The enthalpy and the thermal stability of the DNA transition

were affected by HHP treatments above 300 MPa. Linear relationship between the

fractional viability based on calorimetric data and plate count data was obtained. In EM

studies, integrity of cell envelope was maintained in pressure- or heat-inactivated cells;

however, the leakage of cell wall substances and the formation of empty space between

cell envelope and internal structure were observed in pressure- inactivated cells.

137

Page 154: Calorimetric and Microbiological Evaluation of Bacteria

Key Words: high hydrostatic pressure, differential scanning calorimetry, electron

microscopy, Escherichia coli, inactivation

138

Page 155: Calorimetric and Microbiological Evaluation of Bacteria

INTRODUCTION

Because the interest among consumers in natural or minimally processed foods is

increasing, the inactivation of pathogenic and spoilage microorganisms using a treatment

alternative to thermal processing is under consideration by food industry. High

hydrostatic pressure (HHP) treatment has the potential to produce the microbiologically

safe food without impairing the nutrient content of a food (Mertens and Deplace, 1993;

Roberts and Hoover, 1996). As early as 1889 Hite showed that pressures of 450 MPa or

greater could eliminate spoilage microorganisms and improve the preservation of milk.

The effectiveness of hydrostatic pasteurization on the destruction of several foodborne

pathogens such as Salmonella spp., Escherichia coli O157:H7, Vibrio parahaemolyticus,

Listeria monocytogenes and Staphylococcus aureus has been reported (Metrick et al., 1989;

Styles et al, 1991; Patterson et al, 1995; Kalchayanand et al, 1998; Alpas et al, 1999). Cell

viability decreases with increasing pressure, time, and temperature suggesting critical

cellular activities have been irreversibly damaged (Hoover et al., 1989; Metrick et al.,

1989; Alpas et al., 2000; Robey et al., 2001).

In HHP treatment, the primary target in bacterial cell is proposed to be the

cytoplasmic membrane (Kalchayanand et al., 1998; Farkas and Hoover, 2000). Studies

showed that bacterial cell viability is related to the loss of the membrane integrity

(Shigehisa et al., 1991; Casadei et al., 2002). The denaturation of membrane bound

ATPases, which includes the alteration of molecular structures and change in active sites,

has also been considered as a major factor in pressure-induced cell inactivation (Suzuky and

139

Page 156: Calorimetric and Microbiological Evaluation of Bacteria

Suzuky, 1962; Mackey et al., 1995; Wouters et al., 1998). Studies using electron

microscopy (EM) reported that the empty space between the cell envelope and inner cell

body, compacted amorphous clusters in ribosome or protein region and fibril formation in

DNA region were observed in transmission electron microscopy (TEM) sections on

pressure inactivated Salmonella Thompson (Mackey et al. 1994) and Lactobacillus

viridescens (Park et al., 2001). In a scanning electron microscopy (SEM) study on

Pseudomonas fluorescens, Lopez-Caballero et al. (2002) reported that pressure induced

rough and wrinkled cell surface is related to the destruction of cell wall which also leads

the leakage of intra cellular materials.

Differential scanning calorimetry (DSC) has been used to detect thermally induced

conformational transitions in bacterial cells and spores (Miles et al., 1986; Mackey et al.,

1991; Belliveau et al., 1992; Mohacsi-Farkas et al., 1999; Lee and Kaletunç, 2002a,b).

Peak temperatures of observed endothermic transitions, in DSC thermograms, correspond

to the thermal stabilities of cellular components of microorganism (Mackey et al., 1991;

Kaletunç, 2001). In addition, DSC measurement provides information about amount of

energy (enthalpy, ∆H) associated with the transition. DSC also has been utilized to

evaluate the effects of treatments other than heat such as pH (Mohacsi-Farkas et al.,

1994) and pressure (Niven et al., 1999; Alpas et al., 2003) by comparing the thermograms

of cells before and after treatment. Niven and coworkers (1999) investigated the effect of

high hydrostatic pressures (HHP) on ribosome conformation in whole E. coli NCTC 8164

cells using DSC. They observed reduction in ribosome-associated enthalpy was

correlated with loss of cell viability due to pressure treatment between 50 to 250 MPa.

140

Page 157: Calorimetric and Microbiological Evaluation of Bacteria

The effect of maximum pressure treatment (250 MPa), however, was not enough to

inactivate (less than a log reduction) the E. coli cells in their study. Furthermore, the

information in literature about the irreversible changes of specific macromolecular

components occurring in cells, which leads to cell death as a result of HHP treatment is

limited.

The objectives of this study were (1) to investigate the stability of cellular

components of pressure treated E. coli using DSC, (2) to determine the fraction of

survivors as a function of HHP (up to 700 MPa) using plate count and calorimetric data,

and (3) to compare the changes in cellular components of pressure-inactivated cells with

that of heat-inactivated cells using the DSC and the electron microscopy.

MATERIALS AND METHODS

Preparation of organisms for pressurization

E. coli K12 was obtained from the Culture Collection, Department of

Microbiology at the Ohio State University. A loopful of organism was revived in 10 ml

Trypticase soy broth (Difco laboratories, Detroit, MI) supplemented with 0.3 % (w/w)

yeast extract (Difco laboratories, Detroit, MI) (TSBYE) and incubated at 37oC for 18

hours. The culture was stored frozen (-80 oC) in 30 % (v/v) sterile glycerol. A loopful of

the stock culture was transferred to 10 ml TSBYE and incubated 10 hrs at 37oC before

141

Page 158: Calorimetric and Microbiological Evaluation of Bacteria

use. Revived E. coli K12 culture was inoculated (1% v/v) into TSBYE. Cultures are

incubated at 37oC until the late exponential growth phase, when the final concentration of

cells in a medium was 1.0 ± 0.1 x 109 cfu ml-1. Cell suspensions (200 ml) were removed

from the incubator and placed in 3-mil-thick sterile polyethylene stomacher bag (Fisher

scientific, Pittsburgh, PA) (10 x 15 cm). Air was removed from the bag prior to heat

sealing. The bag was placed inside a second stomacher bags (Fisher scientific, Pittsburgh,

PA) (18 x 30 cm) to prevent contamination of the high-pressure unit if the primary package

was to fail and heat sealed under vacuum. Prepared bags were kept at 4oC prior to

pressurization. Control samples were prepared for plate count, SEM, TEM, and DSC

analysis. One of the control samples was placed in hot water bath (65oC) until no viable

cells detected. The cellular morphology and components of cells in this heat-inactivated

broth or pellet was compared with that of pressure-inactivated cells using the electron

microscopy or DSC.

High Hydrostatic Pressure Processing

A High Pressure Processing Unit (ABB Quintus Food Processor QFP-6 Cold

Isostatic Press, Columbus, Ohio) was used for pressure treatments. The hydrostatic

pressurization unit was capable of operating up to 900 MPa (8,874 atm). A

water/propylene glycol (Houghton-Safe 620-TY, Houghton Int., Inc., Valley Forge, Pa)

mixture (1:1, vol/vol) was used as the pressure transmitting fluid. The liquid can be

heated to the desired temperature prior to pressurization by an electric heating system

around the chamber. The rate of pressure increase was about 400 MPa per min. The

142

Page 159: Calorimetric and Microbiological Evaluation of Bacteria

pressure level, time, and temperature of pressurization were controlled and maintained

during the pressurization cycle.

The bags containing cell suspensions were pressurized at 100, 200, 300, 400, 500,

600 or 700 MPa pressure for 5 min at 35oC. Duplicate samples were prepared at each

treatment level. After HHP treatment, a portion (1 ml) of the pressured and untreated

(control) cultures in the bags was serially diluted in 0.1 % sterile peptone solution and

pour plated into Trypticase soy agar (TSA) to determine viable cell counts. Remained

cell suspensions were used for preparation of SEM and TEM samples, and for DSC

analysis.

Heat Inactivation

One of the bags containing control culture was placed in hot water bath (65oC) until no

viable cells detected (6 min). The cellular morphology and components of cells in this

heat-inactivated broth or pellet was compared with that of pressure-inactivated cells using

the electron microscopy or DSC.

Calorimetry

Cells in the pressure-treated and untreated broth were centrifuged (Beckman J2-21

centrifuge, Palo Alto, CA) at 10 000 g for 10 min at 4oC. Supernatants were discarded.

Pellets were washed with 100 ml of sterile distilled water before transferring into DSC

crucibles. A DSC run was performed with empty sample and reference crucibles to

measure the empty crucible baseline. Pellets of cells were weighed and carefully

143

Page 160: Calorimetric and Microbiological Evaluation of Bacteria

transferred into the sample crucible. The dry material content of the pellets was

determined by freeze drying (Freezone 4.5, Freeze dry system, Model 77510, Labconco,

Missouri). For each DSC run, the reference crucible was filled with distilled water equal

to the amount of water in sample. The sealed crucibles were refrigerated at 4oC until

used for DSC. The sample and reference crucibles were placed in the DSC and

equilibrated at 1oC using liquid nitrogen cooling system. Samples were heated in the

DSC instrument at 4oC/min from 1 to 150oC. DSC thermograms of samples were

corrected for differences in the empty crucibles by subtraction of an empty crucible

baseline. Peak areas (apparent enthalpies, J g-1) corresponding to the contributions of

survivors were determined from the apparent heat capacity vs. temperature graph using

software provided by the instrument manufacturer.

EM preparation

Cells in the pressure-treated, heat-inactivated, and control broth were centrifuged at

10 000 g for 10 min at 4oC to separate the cells as pellets. Pellets were washed with 150

ml of sterile distilled water. A 1 mm3 pellet was transferred to a sterile vial and

resuspended in 1 ml of 0.1 M phosphate buffer at pH 7.4. The cells were fixed on the

membrane (0.45 µm pore size) by passing with 10 ml 3% glutaraldehyde in 0.1M

phosphate buffer (pH 7.4) through the filter. Fixative was left in contact with the cells

overnight at 4oC. The cells for SEM observation were washed with buffer and post-fixed

for 1 hour in 1% osmium tetroxide in phosphate buffer. Filters were rinsed with buffer,

dehydrated with a serial concentrations of ethanol, and then dried on a critical point drier.

144

Page 161: Calorimetric and Microbiological Evaluation of Bacteria

The dried cells were coated with gold-palladium and examined using a Philips XL-30

SEM at 30 KV (FEI Inc., Oregon). Fixed cells for TEM observation were centrifuged

and the pellet embedded in 2% agar. Agar was cut into 1mm3 pieces and post-fixed for 1

hour in 1% osmium tetroxide in phosphate buffer. Samples were rinsed in distilled water

and in-block stained for 1 hour in 1% aqueous uranyl acetate. After dehydration with a

serial concentrations of ethanol, cells in agar were transferred to propylene oxide and

infiltrated and embedded in Spurr’s resin (Ted Pella Co., Redding, CA). The samples

were sliced (70 nm) with an ultramicrotome and stained with Reynold’s lead citrate

(Reynolds, 1963) prior to observing by a Philips CM-12 TEM at 60KV (FEI Inc.,

Oregon).

e

Growth of E. coli to the late exponential stag

High Hydrostatic Pressure (HHP)processing

cal a M y

Plate countingfor viability DSC for

orimetric dat

Centrifugation to obtain cell pellets

145

Heat inactivation usingwater bath

Electron icroscop

Determination of the effects of HHP on E. coli

Figure 5.1. Experimental scheme of calorimetric, EM and microbial analysis

Page 162: Calorimetric and Microbiological Evaluation of Bacteria

RESULTS

Survivability of E. coli after pressure treatment

Inactivation of E. coli K12 by HHP in TSBY broth for 5 min at 35oC was

determined by microbial counts (Fig. 5.2). The plate counts were immediately done after

the application of pressure. The viability of E. coli was not affected by pressure up to

200 MPa (Table 5.1 and Fig. 5.2). The number of survivors decreased drastically when

pressure increased above 300 MPa. The inactivation was considered complete when no

colony was observed in TSA after incubation at 37oC for 36 hours. The complete

inactivation occurred after 600 MPa and above treatments (Fig. 5.2).

-9

-8

-7

-6

-5

-4

-3

-2

-1

0

0 100 200 300 400 500 600 700

Log

(N/N

0)

)

Figure 5.2. Pressure dependence

Pressure (MPa

of fractional viability determined by plate count.

146

Page 163: Calorimetric and Microbiological Evaluation of Bacteria

0.1 MPa 200 MPa 300 MPa 400 MPa 500 MPa 600 MPa 700 MPaWater bath (65oC for 6

min)

CFU/ml 1.4 x 109 1.1 x 109 1.6 x 107 2.0 x 105 9.4 x 103 <1 x 101 <1 <1

Apparent Enthalpy(J/g)

3.83 3.64 2.99 2.60 1.93 1.76 1.75

1.69

Tm (oC) a1 a2 b c d

55.0 71.9 92.4 103.4 118.2

70.5, 74.2 91.6

102.1 118.2

55.7

67.3, 70.0 89.2 101.9 118.0

60.4, 67.2 -

100.9 117.6

61.7, 72.7 -

100.8 118.6

66.8

100.4 117.3

70.2

100.6 117.4

63.8, 71.9 85.5 102.1 118.2

147

Table 5.1. Viability, apparent enthalpy values and transition temperature (Tm) of each peak for E. coli cells after treatments.

Page 164: Calorimetric and Microbiological Evaluation of Bacteria

Comparison of DSC thermograms and viability of E. coli after pressure treatment

The shapes and temperatures of endothermic transitions observed in DSC

thermograms were different after pressure treatment for 5 min at 35oC in comparison

with thermograms of untreated cell pellets (Fig. 5.3). The fractional viability based on

apparent enthalpy data [(∆H-∆Hf)/(∆H0-∆Hf)], where ∆H is the apparent enthalpy after a

HHP treatment, ∆Hf is the residual apparent enthalpy after treatment resulting in no

viability (at 700 MPa) and ∆H0 is the apparent enthalpy of untreated cells, was plotted

versus pressure (Fig. 5.4). The slope of plots in Figure 5.4 is very steep at the pressure

range between 200 and 600 MPa where cells were exponentially inactivated in the plate

count data (Fig. 5.2). Natural logarithm of fractional viability based on the apparent

enthalpy data and plate count data (N/N0) are plotted in Figure 5.5. A linear relationship

(r2 = 0.91) between the reduced apparent enthalpy and the fraction of survivors is

observed. The decrease in apparent enthalpy values in Table 5.1 is strongly related with

the reduction of ribosomal subunit peaks (a1 and a2) in thermograms in Figure 5.3 as

pressure increases. Two endothermic transitions are observed in the temperature range of

major ribosomal subunit denaturation (70~75oC) in the thermogram of 200 MPa treated

cells (Fig. 5.3 thermogram B). Significant reductions in the apparent enthalpy value

(~22%) and the plate count value (~2 log10 unit) occurred with HHP treatment at 300

MPa (Table 5.1). The noticeable reduction of peak b, which is similar to the peak

identified by Mackey et al. (1991) as the melting of DNA, occurred in thermograms of

cells treated at 300 MPa and above pressures. Peak c became more apparent after the

HHP treatments. The area of the peak d which corresponds to the enthalpy of a

148

Page 165: Calorimetric and Microbiological Evaluation of Bacteria

component in outer membrane denaturation was decreased (~30%) after 400 MPa and

above HPP levels.

149

Page 166: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120

A

B

E

G

F

C

D

H

0.5 mW

a1

a2

b c d

Temperature (oC)

Figure 5.3. DSC thermograms of pellets of E. coli whole cell after HHP (35oC for 5 min) or heat (65oC for 6 min) treatments. Control (A), HHP treatment levels: 200 MPa (B), 300 MPa (C), 400 MPa (D), 500 MPa (E), 600 MPa (F), 700 MPa (G), and heat treatment (H). Thermograms are offset for clarity.

150

Page 167: Calorimetric and Microbiological Evaluation of Bacteria

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 100 200 300 400 500 600 700

[(∆H

-∆H

f)/(∆

H0-

∆Hf)]

Pressure (MPa)

Figure 5.4. Pressure dependence of fractional apparent enthalpy determined by DSC.

151

Page 168: Calorimetric and Microbiological Evaluation of Bacteria

0

1

2

3

4

5

6

0 5 10 15 20 25

-ln[

(∆H

-∆H

f)/(∆

H0-

∆Hf)]

)

Figure 5.5. Correlation between fractionalHHP treated E. coli.

1

-ln(N/N0

apparent enthalpy and fractional viability for

52

Page 169: Calorimetric and Microbiological Evaluation of Bacteria

SEM study on the cellular components after treatments

The shape and the surface appearance of inactivated E. coli cells were evaluated

from SEM photomicrographs obtained at 25,000X magnification after pressure treatment

at 700 MPa for 5 min at 35oC or heat treatment at 65oC for 6 min (Fig. 5.6). Before

treatment, the surface of cells in SEM micrograph was clean and smooth (Fig. 5.6a). The

appearances of cell surfaces became rough and wrinkled when the cells were inactivated

at 700 MPa (Fig. 5.6b). Many cells have protruding parts on their surfaces were also

observed in Figure 5.6b. Like pressure-inactivated cells, the surfaces of heat inactivated

cells became rough and cracked; however, no protruding formation were found on their

surfaces (Fig. 5.6c). The cell length of heat-inactivated cells appeared to be longer than

that of controls and pressure-inactivated cells.

(a)

153

Page 170: Calorimetric and Microbiological Evaluation of Bacteria

(b)

Protruding parts

(c)

Figure 5.6. SEM micrograph of control (a), pressure-inactivated (b, at 700 MPa, 35oC, 5 min), and heat-inactivated (c, at 65oC for 6 min) E. coli cells.

154

Page 171: Calorimetric and Microbiological Evaluation of Bacteria

TEM study on the cellular components after treatments

The morphology of intracellular structures and cell envelope of inactivated E. coli

cells were examined from TEM photomicrographs obtained at 70,000X magnification

after pressure treatment at 700 MPa for 5 min at 35oC or heat treatment at 65oC for 6 min

(Fig. 5.7). The internal appearance of untreated cells in TEM micrograph was

characterized by uniform density distribution of ribosomes and central electron

transparent region of DNA (Fig. 5.7a). When the cells were inactivated by 700 MPa

pressure, compacted peripheral dark regions and fibrous central area are shown in

internal bodies (Fig. 5.7b). Similar appearances are shown in internal structures of heat

inactivated (at 65oC for 6 min) cells (Fig. 5.7c). Empty spaces were observed between

internal bodies and cell envelope in the pressure inactivated cells while those were absent

in the heat inactivated cells (Fig. 5.7b,c). Most of pressure or heat inactivated cells

maintained integrity of membranes in their cell envelope.

155

Page 172: Calorimetric and Microbiological Evaluation of Bacteria

(a)

DNA

Cell envelope

Ribosomes

Figure 5.7. TEM micrographs of untreated (a), pressure-inactivated (b, at 700 MPa, 35oC, 5 min), and heat-inactivated (c, at 65oC for 6 min) E. coli cells. Bar = 1.0 µm.

156

Page 173: Calorimetric and Microbiological Evaluation of Bacteria

(b)

DNA

Ribosomes

Cell envelope

157

Page 174: Calorimetric and Microbiological Evaluation of Bacteria

DNA

Cell envelope

(c)

Ribosomes

158

Page 175: Calorimetric and Microbiological Evaluation of Bacteria

DISCUSSION

This study aims to assess the effect of HHP on viability and structure of E. coli cells

by: i. evaluating the relationship between the conformational changes in cellular

components and cell viability using plate count and DSC data; ii. comparing the

morphological and conformational changes in the components of pressure inactivated

cells with that of heat inactivated cells using electron microscopy and DSC. Preliminary

study showed that HHP treatment of 100 MPa did not affect either the viability of the

culture or transitions of cellular components in whole cell thermogram. Therefore, the

effects of 200 MPa and above treatments on the organism were evaluated in present

study.

DSC shows the changes in conformational transitions of cellular components and

apparent enthalpy of E. coli cell pellets after treated at each level of HHP (Fig. 5.3). The

ribosomal subunits are the cellular component most affected by the 200 MPa HHP

treatment (Fig. 5.3 thermogram B). Two endothermic transitions are observed in peak a2

which is proposed to be the denaturation of the main ribosomal subunit (Mackey et al.,

1991). Since a large and broad peak in the thermogram of bacterial whole cells has been

considered as the sum of several overlapped transitions, the appearance of the two peaks

possibly indicates that one of those transitions was shifted to lower temperature after the

HHP treatment while the other transition(s) was/were not affected or reversible. In the

DSC study on isolated 50S and 70S ribosomal subunits of E. coli, Bonincontro et al.

(1998) reported that the two peaks found in a ribosomal subunit is accounted as a

159

Page 176: Calorimetric and Microbiological Evaluation of Bacteria

complex unfolding process on interactions such as protein/RNA, protein/protein and

RNA/RNA within the subunit (Bonincontro et al., 1998). It has been known that

alteration of tertiary structure of proteins occurs at 200 MPa pressure level (Vidugiris et

al., 1995), while denaturation of RNA structure requires much higher level of pressure

(Smelt et al., 1998). Since ribosomal subunits are composed of RNA and protein, the

pressure-induced conformational change of ribosomal proteins may have important role

in the formation of two peaks in thermogram by altering protein/protein and protein/RNA

interactions in ribosome structures. However, there were no noticeable changes in the

apparent enthalpy and plate counts after 200 MPa in Table 5.1, indicating that the initial

conformational change in ribosomal subunits did not affect the cell viability.

The reductions in the apparent enthalpy value (~22%) and viable count (~2 log10

unit) of E. coli cells occurred after HHP at 300 MPa (Table 5.1). The temperature and

size of peak a2, which is associated with the denaturation of larger and more rigid

ribosomal subunits, were also lowered. However, peak a1 (probably 30S subunit), which

had been only partially visible due to the overlapping peak a2 in the control, became

visible (Fig. 5.3 thermogram C). It has been reported that the volume reduction of

ribosome molecule occurs during HHP treatment as results of hydration changes around

ribosome and conformational changes which are followed by denaturation (Balny and

Masson, 1993; Alpas et al., 2003). Among ribosomal subunits of bacteria, larger subunit

is believed to have greater tendency to decrease its volume in HHP treatment (Alpas et

al., 2003). Therefore, the decrease in the size of peak a2 and the appearance of peak a1 in

the thermogram of 300 MPa pressure-treated cells indicate the volume reduction of larger

160

Page 177: Calorimetric and Microbiological Evaluation of Bacteria

ribosomal subunit (probably 50S) associated in peak a2 is greater than that of smaller

ribosomal subunit (probably 30S).

The peak in the range of 90-100oC in DSC of whole bacterial cells was associated

with the melting of double strand DNA due to the thermally-induced breakage of

hydrogen bonds between base pairs of two single strand DNA structures (Miles et al.,

1986, Mackey et al., 1988). The effect of HHP on thermal stability of DNA double helix

has commonly been studied by measuring the change in the ultraviolet adsorption as a

function of temperature after pressurization of isolated DNA in salt solution. Studies

showed that the base pairing of DNA duplex are stabilized by high pressure at least up to

200~270 MPa (Héden et al., 1964; Kumar, 1995; Lin and Macgregor Jr., 1996).

However, in a recent DSC study, a decrease in thermal stability of DNA duplex in

Leuconostoc mesenteroides was observed after HHP treatment at 250 MPa and above

pressures (Kaletunç et al., 2004). Noticeable size reduction of peak b in thermograms of

cells treated at 300 MPa and above pressures in present study indicates that dissociation

of double stranded DNA, which induces irreversible change of the DNA transition,

possibly occurred in the pressure treatments. Since little information about the HHP level

which leads denaturation of the DNA double helix in bacteria cell is available in

literature, further study on the detection of change in DNA structures in vivo after HHP

treatment will be needed to confirm the result of this study. The appearance of peak c, a

peak known to be also related to denaturation of DNA with a cell wall component

(Mackey et al., 1991) became apparent due to the reduction of peak b which partially

obscured the peak c in the thermogram of untreated E. coli cells (Fig. 5.3 thermogram C).

161

Page 178: Calorimetric and Microbiological Evaluation of Bacteria

Figure 5.3 shows that the area of the peak d, which corresponds to the denaturation

enthalpy of components in outer membrane, was decreased (~30%) after 400 MPa and

above HHP treatments. In protein content analysis study of Salmonella Typhimurium

using electrophoresis SDS-PAGE, Ritz et al. (2000) observed that most of major outer

membrane proteins of the bacteria disappeared after 350 MPa and above HHP treatments.

The irreversible denaturation or forced out from the membrane has been considered as

reason for the loss of the membrane proteins by HHP (Casadei et al., 2002). However,

since thermal stabilities of the E. coli outer membrane proteins are ~70oC (Phale et al.,

1998), and these proteins are not directly interacted with thermally stable components in

outer membrane, it is doubted that pressure effect on the outer membrane proteins made

the change in peak d in the thermograms. Therefore, it is possible that the reduction of

the peak d in the thermograms is mainly related with the conformational change in other

outer membrane component such as LPS by higher levels (>400 MPa) of HHP

treatments. LPS is predominant component (~40% weight) in outer membrane of E. coli

and hypothesized to be associated with the endothermic transition (peak d) (Lee and

Kaletunç, 2002b).

The plots in Figure 5.2 and 5.4 suggest that the fraction of apparent enthalpy [(∆H-

∆Hf)/(∆H0-∆Hf)] from DSC data is closely related to the log of fractional viability [log

(N/N0)] from plate count data subsequent to pressure treatments. No slope changes up to

200 MPa treatment in both figures indicates that although it affected the peak associated

with major ribosome subunit (Fig. 5.3), HPP treatment at 200 MPa had no great influence

on either apparent enthalpy or plate count data. The steep slopes of the plots after 300

162

Page 179: Calorimetric and Microbiological Evaluation of Bacteria

MPa treatment in the Figures 5.2 and 5.4 suggest that the reductions in the apparent

enthalpy and plate count data were accompanied by a decrease in the area of the peak (a2)

corresponding to the alterations of larger ribosomal subunits (Fig. 5.3) at 300 MPa and

above pressure levels. The size reduction of one of the DNA transition (peak b) due to

possible pressure-induced damages on DNA duplex attributed the reduction of apparent

enthalpy and might cause more cell viability loss on plate count result.

A plot of the fractional apparent enthalpy versus the fractional survivors from plate

count data (Fig. 5.5) gives a linear relationship by taking the logarithm of both sides.

However, in the pressure treatment range from 300 to 500 MPa, where the cell population

was exponentially decreased in plate count method (Fig. 5.2), disparities appear between

the calorimetric data and plate count data. The result indicates that the decrease in the

fraction of plate count data against the decrease in the fraction values of apparent

enthalpy data are relatively higher in the HHP treatment levels between 300 and 500 MPa

than that in other pressure levels (200, 600 MPa).

Many adhering cells covered with protruding substances were observed in Figure

5.6b indicating components of cell envelope released during the HHP. The observed

whole cell population in our SEM and TEM studies seemed to maintain its shape without

cell rupture or breakage at 700 MPa pressure inactivation. The property of the

smoothness of Gram-negative bacterial cell surface is mainly contributed by O

polysaccharide (O antigen) in LPS structure (Todar, 2002). Therefore, we assume that

the denaturation or the detachment from the cell of LPS by either pressure or heat

inactivation might play an important role in the formation of the roughness on the surface

163

Page 180: Calorimetric and Microbiological Evaluation of Bacteria

of cells. However, it is still possible that the protruding substances also contain

intracellular components that leaked out via membranes which can lose their permeability

during high pressure treatment.

The dark compacted regions in TEM micrographs of pressure-inactivated cells have

been thought to be residual ribosomes and proteins in cytoplasm after denaturation

(Mackey et al., 1994). Because no clear evidence of cell membrane damages in TEM of

the pressure-inactivated cells (Fig. 5.7b), ribosomes and proteins might remain in the cell

and contribute the transitions in DSC thermogram. However, DSC thermogram G in

Figure 5.3 shows that peak associated with ribosome denaturation almost disappeared

after cells were inactivated by pressure while the peripheral dark region are clearly shown

in TEM of the cells. It is suggested that pressure induced denaturation and later

aggregation of ribosome result in the reduction of ribosomal peak in DSC thermogram

(Kaletunç et al., 2004). Therefore, the disappearance of ribosomal transition in the

thermogram G possibly indicates that both denaturation of residual ribosome

(endothermic event) and aggregation of denatured ribosome (exothermal event) occurred

during DSC scan to reduce ribosomal transition in the thermogram of 700 MPa treated

cells. Comparison of thermograms of pressure inactivated cells and heat inactivated cells

(Fig. 5.3. Thermogram G and H) shows that two peaks (63.5 and 68.2oC) remained in the

ribosome transition for heat inactivated cells although the apparent enthalpy for both cells

were similar. The result may suggest the effect of HHP on ribosome structure is greater

than that of heat treatment when lethal level of each treatment is applied.

164

Page 181: Calorimetric and Microbiological Evaluation of Bacteria

The condensed fibril formation in central regions of both pressure and heat

inactivated cells possibly indicates that DNA structures were affected by both treatments.

Similar formations were shown in HHP (>500 MPa) treated cells of Lactobacillus

viridescens (Park et al., 2001) and Salmonella Thompson (Mackey et al., 1994).

Thermogram G and H in Figure 5.3 show that one of the DNA transition (peak b) was

irreversible after both treatments indicating the denaturation of DNA resulted in the

formation of fibrils in the TEM of the treated E. coli cells. In the comparison of peak c,

which is also associated with DNA transition, the thermal stability (Tm) of the peak in the

thermogram of heat inactivated cells is 1.5oC higher than that of HHP inactivated cells.

The result implies that the effect on DNA is also greater in HHP inactivation.

The cause of the appearances of empty spaces during HHP (Fig. 5.7b) has still been

controversial. It was thought to be the reversible invagination of cytoplasmic membrane,

which increases its surface area due to pressure-induced phase changes in lipid bilayer,

causes the spaces in Salmonella Thompson (Mackey et al., 1994). On the other hand,

Park et al. (2001) claimed that the separation processes of deformed cell wall structures

from the internal cell bodies cause the spaces in Lactobacillus viridescens. Because the

internal bodies of the pressurized cells in Figure 5.7b seem to be physically squeezed, the

cause of the empty spaces on E. coli in this study might be similar to the idea in Mackey

et al. (1994). Therefore, it is possible that the cause of empty spaces in pressurized

bacteria cells depends on their cell wall types (Gram-positive or Gram-negative). Most

pressurized cells retained the membrane line in cell envelope region in our TEM section,

indicating that there was no apparent evidence of cell envelope rupture shown after HHP.

165

Page 182: Calorimetric and Microbiological Evaluation of Bacteria

Mackey et al. (1994) stated that the invaginated membrane return to original position

when pressure is released.

In conclusion, DSC study shows that the reversibility of transition and the change in

the thermal stability of ribosome of E. coli were affected by 200 MPa and above

pressures in HHP treatment. The enthalpy and the thermal stability of the DNA melting

transition were affected by HHP treatments above 300 MPa. The morphological changes

in structures of ribosome and DNA are also shown in TEM section. There is a close

relationship of fractional viability based on calorimetric data and plate count data. In

EM study, no rupture of cell envelope is shown in pressure- or heat- inactivated E. coli

cells. However, leakage of cell wall or outer membrane substance and empty space

between cell envelope and inside structure are exclusively observed in pressure-

inactivated cells. Comparison of the cell components by DSC combined with EM prior

and after HHP inactivation allows us to evaluate the changes resulting from HHP

processing which coincide with inactivation of microorganisms.

166

Page 183: Calorimetric and Microbiological Evaluation of Bacteria

REFERENCES

Alpas, H., Kalchayanand, N., Bozoglu, F., Sikes, A., Dunne, C.P., and Ray, B. 1999. Variation in resistance to hydrostatic pressure among strains of food-borne pathogens. Appl. Environ. Microbiol. 65(9): 4248-4251. Alpas, H., Kalchayanand, N., Bozoglu, F. and Ray, B. 2000. Interactions of high hydrostatic pressure, pressurization temperature and pH on death and injury of pressure-resistant and pressure-sensitive strains of foodborne pathogens. Int. J. Food Microbiol. 60:33-42. Alpas, H., Lee, J., Bozoglu, F., and Kaletunç, G. 2003. Differential scanning calorimetry of pressure-resistant and pressure-sensitive strains of Staphylococcus aureus and Escherichia coli O157:H7. Int. J. Food Microbiol. 87:229-237. Balny, C. and Masson, P. 1993. Effects of high pressure on proteins. Food Rev. Int. 9:611-628. Belliveau, B.H., Beaman, T.C., Pankratz, H.S. and Gerhardt, P. 1992. Heat killing of bacterial spores analyzed by differential scanning calorimeter. J. Bacteriol. 174:4463-4474. Bonincontro, A., Cinelli, S., Mengoni, M., Onori, G., Risuleo, G. and Santucci, A. 1998. Differential stability of E. coli ribosomal particles and free RNA towards thermal degradation studied by microcalorimetry. Biophys. Chem. 75:97-103. Casadei, M. A., Manas, P., Niven G., Needs, E. and Mackey, B. M. 2002. Role of Membrane Fluidity in Pressure Resistance of Escherichia coli NCTC 8164. Appl. Environ. Microbiol. 68: 5965-5972. Farkas, D.F. and Hoover, D.G. 2000. High pressure processing. J. Food Sci. 65(Supplement):47-64.

167

Page 184: Calorimetric and Microbiological Evaluation of Bacteria

Héden, C.-G., Lindahl, T. and Toplin, I. 1964. The stability of Deoxyribonucleic acid solutions under high pressure. Acta Chem. Scand. 18:1150-1156. Hite, B.H. 1899. The effects of pressure in the preservation of milk. West Virginia Agri. Exp. Station Bulletin 58:15-35. Hoover, D.G., Metrick, C., Papineau, A.M., Farkas, D.F. and Knorr, D. 1989 Biological effects of high hydrostatic pressure on food microorganisms. Food Technol. 43(3):99-107. Kalchayanand, N., Sikes, A., Dunne, C.P. and Ray. B. 1998. Factors influencing death and injury of foodborne pathogens by hydrostatic pressure-pasteurization. Food Microbiol. 15:207-214. Kaletunç, G. 2001. Thermal analysis of bacteria using differential scanning calorimetry. In Novel Process and Control Technologies in the Food Industry ed. Bozoglu, F., Deak, T. and Ray, B. pp. 227-235. Amsterdam: IOS press. Kaletunç, G., Lee, J., Alpas, H. and Bozoglu, F. 2004. Evaluation of structural changes induced by high hydrostatic pressure in Leuconostoc mesenteroides. Appl. Environ. Microbiol. 70:1116-1122. Kumar, A. 1995. Alternate view on thermal stability of the DNA duplex. Biochemistry 34:12922-12925. Lee, J. and Kaletunç, G. 2002a. Calorimetric determination of inactivation parameters of microorganisms. J. Appl. Microbiol. 93:178-189. Lee, J. and Kaletunç, G. 2002b. Evaluation by differential scanning calorimetry of the heat inactivation of Escherichia coli and Lactobacillus plantarum. Appl. Environ. Microbiol. 68:5379-5386. Lin, M-C. and Macgregor Jr, R.B. 1996. The activation volume of a DNA helix-coil transition. Biochemistry 35:11846-11851.

168

Page 185: Calorimetric and Microbiological Evaluation of Bacteria

Lopez-Caballero, M.E., Carballo, J., Solas, M.T. and Jimenez-Colmenero, F. 2002. Responses of Pseudomonas fluorescens to combined high pressure/temperature treatments. Eur. Food. Res. Technol. 214:511-515. Mackey, B.M., Forestiere, K. and Issacs, N. 1995. Factors affecting the resistance of Listeria monocytogenes to high hydrostatic pressure. Food Biotechnol. 9:1-11. Mackey, B.M., Forestiere, K., Issacs, N.S., Stenning, R., Brooker, B. 1994. The effect of high hydrostatic pressure on Salmonella Thompson and Listeria monocytogenes examined by electron microscopy. Lett. Appl. Microbiol. 19:429-432. Mackey, B.M., Miles, C.A., Parsons, S.E. and Seymour, D.A. 1991. Thermal denaturation of whole cells and cell components of Escherichia coli examined by differential scanning calorimetry. J. Gen. Microbiol. 137:2361-2374. Mertens, B. and Deplace, G. 1993. Engineering aspects of high pressure technology in the food industry. Food Technol. 47(6):164-169. Metrick, C., Hoover, D.G. and Farkas, D.F. 1989. Effects of hydrostatic pressure on heat resistant and heat sensitive strains of Salmonella. J. Food Sci. 54:1547-1549. Miles, C.A., Mackey, B.M. and Parsons, S.E. 1986. Differential scanning calorimetry of bacteria. J. Gen. Microbiol. 132: 939-952. Mohacsi-Farkas, Cs., Farkas, J., Meszaros, L., Reichart, O. and Andrassy, E. 1999. Thermal denaturation of bacterial cells examined by differential scanning calorimetry. J. Therm. Anal. Calor. 57:409-414. Mohacsi-Farkas, Cs., Farkas, J. and Simon, A. 1994. Thermal denaturation of bacterial cells examined by differential scanning calorimetry. Acta Aliment. 23:157-168. Niven, G.W., Miles, C.A. and Mackey, B.M. 1999. The effect of hydrostatic pressure on ribosome conformation in Escherichia coli: an in vivo study using differential scanning calorimetry. Microbiology 145, 419-425.

169

Page 186: Calorimetric and Microbiological Evaluation of Bacteria

Park, S.W., Sohn, K.H., Shin, J.H. and Lee, H.J. 2001. High hydrostatic pressure inactivation of Lactobacillus viridescens and its effects on ultrastructure of cells. Int. J. Food Sci. Technol. 36:775-781. Patterson, M.F., Quinn, M., Simpson, R. and Gilmore. A. 1995. Sensitivity of vegetative pathogens to high hydrostatic pressure treatment in phosphate-buffered saline and foods. J. Food Prot. 58:524-529. Phale, P.S., Philippsen, A., Kiefhaber, R., Koebnik, R., Phale, V.P., Schirmer, T. and Rosenbusch, J.P. 1998. Stability of trimeric OmpF porin: The contributions of the latching loop L2. Biochemistry 37:15663-15670. Reynolds, E.S. 1963. The use of lead citrate at high pH as an electron-opaque stain in electron microscopy. J. Cell Biol. 17:208-212. Ritz, M., Freulet, M., Orange, N. and Federighi, M. 2000. Effects of high hydrostatic pressure on membrane proteins of Salmonella typhimurium. Int. J. Food Microbiol. 55:115-119. Roberts, C.M., and Hoover, D.G. 1996. Sensitivity of Bacillus coagulans spores to combinations of high hydrostatic pressure, heat, acidity and nisin. J. Appl. Bacteriol. 81:363-368. Robey, M., Benito, A., Huston, R.H., Pascual, C., Park, S.F. and Mackey, B.M. 2001. Variation in resistant to high hydrostatic pressure and rpoS heterogeneity in natural isolates of Escherichia coli O157:H7. Appl. Environ. Microbiol. 67:4901-4907. Shigehisa, T., Ohmori, T., Saito, A., Taji, S. and R. Hayashi. 1991. Effect of high pressure on characteristics of pork slurries and inactivation of microorganisms associated with meat and meat products. Int. J. Food Microbiol. 12:207-216. Simpson, P.K. and Gilmour, A. 1997. The effect of high hydrostatic pressure on Listeria monocytogenes in phosphate-buffered saline and model food systems. J. Appl. Microbiol. 83:181-188.

170

Page 187: Calorimetric and Microbiological Evaluation of Bacteria

Smelt, J.P.P. 1998. Recent advances in the microbiology of high pressure processing. Trends Food Sci. Technol. 9:152-158. Styles, M.F., Hoover, D.G. and Farkas, D.F. 1991 Response of Listeria monocytogenes and Vibrio parahaemolyticus to high hydrostatic pressure. J. Food Sci. 56:1404-1407. Suzuki, C. and Suzuki, K. 1962. The protein denaturation by high pressure. J. Biochem. 52:67-71. Todar, Kenneth. 2002. Mechanism of bacteria pathogenecity: Endotoxins. In Textbook of bacteriology. University of Wisconsin, Madison. Vidugiris, G.J.A., Markley, J.L. and Royer, C.A. 1995. Evidence for a molten globule-like transition state in protein folding from determination of activation volumes. Biochemistry 34:4909-4912. Woo, I.S., Rhee, I.K. and Park, H.D. 2000. Differential damage in bacterial cells by microwave radiation on the basis of cell wall structure. Appl. Environ. Microbiol. 66:2243-2247. Wouters, P.C., Glaasker, E. and Smelt, J.P.P. 1998. Effects of high pressure on inactivation kinetics and events related to proton efflux in Lactobacillus plantarum. Appl. Environ. Microbiol. 64:509-514.

171

Page 188: Calorimetric and Microbiological Evaluation of Bacteria

CHAPTER 6

INACTIVATION OF SALMONELLA ENTERITIDIS STRAINS BY COMBINATION

OF HIGH HYDROSTATIC PRESSURE AND NISIN

ABSTRACT

The effects of high hydrostatic pressure (HHP) and nisin treatment alone and in

combination on cellular components and viability of two Salmonella enterica subsp.

enterica serova Enteritidis (Salmonella Enteritidis) strains were evaluated by differential

scanning calorimetry (DSC) and plate counting in order to evaluate the relative resistance

and optimize the treatment conditions. Salmonella Enteritidis FDA and OSU 799 strains

were subjected to HHP (0.1- 550 MPa for 10 min at 25oC) alone and in combination with

nisin (200 IU/ml Nisaplin®) in culture broth (TSBY). HHP (up to 200 MPa) or the nisin

alone did not affect the viability and cellular components of either strain. An 8-log

cfu/ml reduction was observed after a pressure treatment at 500 MPa for FDA strain and

450 MPa for the OSU 799 strain. When nisin was added, a similar reduction was

obtained at 400 MPa for FDA strain and 350 MPa for the OSU 799 strain. The decrease

172

Page 189: Calorimetric and Microbiological Evaluation of Bacteria

in apparent enthalpy appeared to be mainly due to reduction in the ribosome denaturation

peak for both pressure alone and pressure-nisin combination treatments. DNA might be

irreversibly damaged by the combination treatments. A linear relationship in a

logarithmic plot of fractional apparent enthalpy values [(∆H-∆Hf)/(∆H0-∆Hf)] versus the

fractional survivors from plate count data (N/N0) for treated cells indicates the apparent

enthalpy data obtained from DSC can be used to evaluate pressure levels necessary to

reduce a microbial population in the presence of nisin and provide information about

viability in shorter time with comparable accuracy to plate count.

Key Words: high hydrostatic pressure, nisin, differential scanning calorimetry,

Salmonella Enteritidis.

173

Page 190: Calorimetric and Microbiological Evaluation of Bacteria

INTRODUCTION

High hydrostatic pressure (HHP) has been shown to inactivate spoilage and

pathogenic bacteria while preserving the quality parameters of food products. HHP has

been recognized as a most likely alternative to thermal processing among emerging food

preservation technologies (Roberts and Hoover, 1996; Mertens and Deplace, 1993;

Knorr, 1993). Inactivation of pressure-resistant pathogens needs very high pressure

levels (>600 MPa) at which texture and color of many foods may be adversely altered

(Porretta et al., 1995; Mussa and Ramaswamy, 1997; Trujillo et al., 1999). Furthermore,

the initial cost and wear of equipment increase at pressure levels above 600 MPa, leading

to high maintenance costs or short equipment life (Hoover et. al., 1989; Mertens and

Deplace, 1993). If moderately high pressure treatments (up to 400 MPa) are applied,

sublethal injury and recovery during storage occur for pressure resistant pathogens

(Earnshaw, 1995; Patterson et al., 1995). Therefore, a processing protocol combining

moderately high pressures with chemical or physical preservation methods can provide

safe food without high processing cost.

The concept of “Hurdle technology” has been applied to inactivate pathogenic

bacteria by combining HHP with low pH (Alpas et al., 2000), heat (Patterson and

Kilpatrick, 1998; Benito et al., 1999; Alpas et al., 2000), lysozyme (Masschalck et al.,

2001), CO2 (Hass et al., 1989) or antimicrobial peptides (Kalchayanand et al., 1998;

Yuste et al., 1998; Garcia-Graells et al., 1999; Masschalck et al., 2000; Massachalck et

al., 2001). Most antimicrobial peptides produced by bacteria are bactericidal to Gram-

174

Page 191: Calorimetric and Microbiological Evaluation of Bacteria

positive bacteria and to sublethally stressed Gram-negative bacteria (Ray et al., 2001).

Kalchayanand et al. (1998) reported that the mixture of pediocin and nisin provided 1 to 5

log unit reductions above the additive effects of bacteriocin and HHP alone in

Staphylococcus aureus, Listeria monocytogenes, Salmonella enterica serovar

Typhimurium (S. Typhimurium) and Escherichia coli O157:H7. Among bacteriocins,

nisin, produced by some strains of Lactococcus lactis, is known to inactivate Gram-

positive bacteria by binding to the cytoplasmic membrane and forming pores which leads

to leakage of intracellular molecules and metabolites (Sahl and Bierbaum, 1998). Its

ineffectiveness against Gram-negative bacteria is attributed to the nisin-impermeable

outer membrane which prevents nisin reaching the cytoplasmic membrane (Kordel and

Sahl, 1986; Delves-Broughton, 1990). Recent studies show that HHP treated Gram-

negative bacterial cells may allow penetration of nisin through the outer membrane

(Masschalck et al., 2000; Masschalck et al., 2001; Ray et al., 2001). Garcia-Graells et al.

(1999) described that pressurization of E. coli in the presence of nisin (400 IU/ml)

decreased the survivability by an additional 3 logs in skim milk at 550 MPa. An

additional 1 to 2 log unit reductions in Pseudomonas fluorescens, E. coli O157:H7, and a

Salmonella sp. were achieved by nisin (100 IU/ml) under moderate (<300 MPa) HHP

treatment (Masschalck et al., 2001). Although the primary target for nisin and HHP in

bacterial cell is proposed to be the cytoplasmic membrane (Kalchayanand et al., 1998;

Masschalck et al., 2001), the exact mechanism of the inactivation of bacteria by HHP or

nisin is still not clearly known.

175

Page 192: Calorimetric and Microbiological Evaluation of Bacteria

DSC thermograms of whole bacterial cells display endothermic transitions

associated with the cellular components of bacteria to heat under linearly increasing

temperature condition. The peak temperature corresponding to each transition represents

the thermal stability of a cellular component of bacteria (Miles et al., 1986; Anderson et

al., 1991; Mackey et al., 1991; Belliveau et al., 1992; Kaletunç, 2001; Lee and Kaletunç

2002b). In addition, DSC measurement provides amount of heat energy (apparent

enthalpy, ∆H) associated with the transition. The apparent enthalpy can be used to

calculate fractional survivors following heat treatment (Lee and Kaletunç, 2002a,b).

DSC has also been utilized for characterization of changes in cellular components of

bacteria by recording scans before and after exposure to a high hydrostatic pressure, a

physical stress (Niven et al, 1999; Alpas et al., 2003; Kaletunç et al., 2004). Niven et al.

(1999) reported reductions in ribosome associated transitions for Escherichia coli NCTC

8164 cells as a function of pressure between 50-250 MPa. Alpas et al. (2003) used

apparent enthalpy data to evaluate the relative high pressure resistance of two bacterial

strains from E. coli O157:H7 and S. aureus.

Development of an understanding for inactivation of Gram-negative bacteria by

pressure-nisin treatment is necessary for optimization of HHP processing conditions to

inactivate Gram-negative bacteria using nisin. The objective of this study was to evaluate

the effect nisin and HHP alone and in combination on cellular components of Salmonella

Enteritidis (Gram-negative food-borne pathogen) using DSC. The fraction of survivors

as a function of treatment was determined using plate count and calorimetric data.

176

Page 193: Calorimetric and Microbiological Evaluation of Bacteria

MATERIALS AND METHODS

Bacterial strains

Salmonella Enteritidis FDA was obtained from the Food Microbiology Laboratory,

University of Wyoming. Salmonella Enteritidis OSU 799 was obtained from the Culture

Collection Center at the Ohio State University. Previous study has shown FDA strain to

be strongly resistant to high pressure among Salmonella Enteritidis strains (Alpas et al.,

1999). An isolated pure colony on of each organism grown on XLD agar plate was

suspended in 10 ml Trypticase soy broth supplemented with 0.6 % (w/w) yeast extract

(TSBY) and incubated at 37oC for 18 hours. Culture was stored frozen (-80oC) in 30 %

(v/v) sterile glycerol. A loopful of stock culture was transferred to 10 ml Trypticase soy

broth and incubated 10 hrs at 37oC before use.

Preparation of organisms for pressurization

Each culture was inoculated (1 % v/v) into a broth containing TSBY and incubated

at 37oC until reaching early stationary growth phase. The growth phase was determined

using viable counts. The growth was used in the subsequent studies to obtain cells at

early stationary phase for pressure treatment. After reaching to early stationary phase,

culture was removed from the incubator and placed (150 ml) in 3-mil-thick sterile

stomacher bags (Fisher scientific, Ottawa, Canada). Nisaplin® (Aplin and Barrett,

Milwaukee, USA) contains 2.5% nisin, was added to the bags containing grown culture

to obtain concentration of Nisaplin (0 and 200 IU/ml). After heat sealed, the bag was

177

Page 194: Calorimetric and Microbiological Evaluation of Bacteria

then placed inside a second bag filled with chlorine solution (20% v/v) and heat sealed to

prevent contamination of the high-pressure unit if the primary package failure is to occur.

Air was removed from all of the bags as much as possible. Prepared bags were vacuum

sealed in sterile plastic bags prior to pressurization.

High Hydrostatic Pressure Processing

Pressurization of bacteria culture was carried out using a High Pressure Processing

Unit (ABB Quintus Food Processor QFP-6 Cold Isostatic Press, Columbus, Ohio). The

hydrostatic pressurization unit is capable of operating at 900 MPa. The pressure chamber

was filled with 50% (v/v) aqueous glycol solution. The temperature of the liquid was

controlled by an electric heating system around the chamber. The rate of pressure

increase was about 400 MPa per min. The pressure level, time, and temperature of

pressurization were controlled. All the processing parameters were maintained during

the pressurization cycle.

Various pressures (150, 200, 250, 300, 350, 400, 450, 500 and 550 MPa) were

applied for 10 min at 25oC. After HHP treatment, a portion (1 ml) of the pressured and

untreated (control) cultures in the bags were serially diluted in 0.1 % sterile peptone

solution and pour plated into Trypticase soy agar to determine viable cell counts. Cells in

the pressure-treated and untreated broth were centrifuged at 10,000 g for 10 min at 4oC

(Beckman J2-21 centrifuge). Supernatants were discarded. Pellets were washed with

100 ml of sterile distilled water before transferring into DSC crucibles for calorimetry.

178

Page 195: Calorimetric and Microbiological Evaluation of Bacteria

Calorimetry of whole cell

A differential scanning calorimeter (DSC 111, Setaram, Lyon, France) was used to

record the thermograms of the cells heated at a 4oC min-1 after HHP treatment. All DSC

measurements were conducted using fluid-tight, stainless steel crucibles. A DSC run was

performed with empty sample and reference crucibles to assure a reproducible empty

crucible baseline. Pellets of whole cells were transferred into the empty sample crucible

and were weighed (70 ± 0.5 mg wet weight). The dry material content of the pellets was

determined by freeze drying (Freezone 4.5, Freeze dry system, Model 77510, Labconco,

Missouri) as 18 ± 0.2 % on a wet basis. The standard deviation was calculated based on

four freeze dried pellets for Salmonella Enteritidis FDA. For each DSC run, reference

crucible was filled with water (~57 mg) equal to the amount of moisture in sample.

Crucibles were sealed using aluminum o-rings and were refrigerated at 4oC prior to DSC

runs. The sample and reference crucibles were placed in the DSC and equilibrated at 1oC

using liquid nitrogen cooling system. Samples were reweighed after DSC measurements

to check for the loss of mass during heating. Thermograms of samples showing signs of

leakage were discarded. DSC thermograms of samples were corrected for differences in

the empty crucibles by subtraction of an empty crucible baseline. Peak areas (apparent

enthalpies, J g-1) corresponding to the contributions of survivors were determined from

the apparent heat capacity vs. temperature graph using software provided by the

instrument manufacturer. A curved baseline using three-temperature points was

developed between the segment of the thermogram prior to the first thermally-induced

179

Page 196: Calorimetric and Microbiological Evaluation of Bacteria

transition (~30oC) and the segment of the thermogram at a temperature below the onset of

the final peak (~110oC).

Statistical Analysis

Viability after the experiments was analyzed using MINITAB statistical program

(Minitab Inc., State College, PA). The means from each variables (pressure level and

concentration of nisin) were determined by two-way analysis to compare the effect of

variables on treatment. Differences in inactivation between pressure only treated cells

and pressure-nisin combination treated cells were compared using t-test. The level of

significance was set for P<0.05.

180

Page 197: Calorimetric and Microbiological Evaluation of Bacteria

Growth of the Salmonella strains to the end of exponential stagephase

Plate counting

a

DSC for calorimetric dat

t

Analysis of the effect of HHP

or Nisin alone treatment

181

Combination treatment by HHPand nisin

Plate counting

Analysis of the effect of the combination treatmen

Treatment by HHP or nisin

Centrifugation to obtain cell pellets

Figure 6.1. Experimental scheme of calorimetric and microbial analysis
Page 198: Calorimetric and Microbiological Evaluation of Bacteria

RESULTS

Thermograms of Salmonella Enteritidis whole cells

Figure 6.2 shows the DSC thermogram for untreated pellet of Salmonella

Enteritidis FDA (A) and OSU 799 (B) cells. Several overlapping endothermic peaks (a,

b, c and d) appear in the thermograms. The transition temperature (Tm) of first major

peak, a, which was identified as denaturation of ribosome in bacteria (Mackey et al.,

1991; Stephens and Jones, 1993; Lee and Kaletunç, 2002b), appears at a higher

temperature in the FDA strain thermogram (71.5oC) in comparison to OSU 799 strain

thermogram (69.5oC). Other peaks, which have been proposed to be thermally induced

transitions of DNA (b, Tm 93oC) and DNA together with cell wall components (c, Tm

100oC), are similar for the two strains. Figure 6.2 also shows that the transition of the

denaturation of outer membrane components of Gram-negative bacteria (Mackey et al.,

1991) appears to be at least two overlapping peaks (d, Tm 125 and 132.3oC ) in the FDA

strain thermogram, while it appears as one broad peak (Tm 119oC) in the OSU 799

thermogram. The total apparent enthalpy values for both untreated cells were 4.0 ± 0.4

J/g wet cell.

182

Page 199: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120

a

bc d

Heat Flow0.5 mW

Temperature (oC)

Figure 6.2. Thermograms of whole cells of Salmonella Enteritidis OSU Salmonella Enteritidis FDA (B) obtained by DSC (1 to 150oC with 4oC mrate).

183

A

B

140

799 (A) and in-1 heating

Page 200: Calorimetric and Microbiological Evaluation of Bacteria

Evaluation of inactivation Salmonella Enteritidis from plate count data after HHP treatment in combination with nisin Pressures at 200, 250, 275, 300 or 350 MPa and Nisaplin at 200, 400 or 600 IU/ml

were used to inactivate Salmonella Enteritidis FDA culture (Appendix Figs. 1,2 and 3).

Nisin alone did not affect the viability of the culture at Nisaplin concentration levels of

200-600 IU/ml. The reduction of viability based on plate counting and apparent enthalpy

data was greater for nisin-pressure combination than for pressure alone (Appendix Figs. 4

and 5). Viability of cells were not significantly different (P>0.05) with respect to nisin

concentration at each pressure treatment level as assessed by both calorimetric and plate

count (Appendix Table 1). Therefore, 200 IU/ml Nisaplin was applied to cultures of two

Salmonella Enteritidis strains to use minimum amount of antimicrobial agent.

Table 6.1 gives the viability of both strains treated by nisin and pressure alone and

in combination. Nisin alone did not affect the viability of either strain. Table 6.1 also

shows that the combination treatment was more effective in comparison with pressure

alone treatment for both strains. At 200 MPa, the combined treatment of pressure and

nisin caused ~ 1 log unit decrease in cell viabilities of both strains, while the pressure

alone did not affect viabilities. At 300 MPa and above pressure, addition of nisin caused

between 2 and 3 log units additional inactivation for both strains.

Fractional viability values from plate count data (N/N0) was plotted against pressure

for two strains, which were pressurized with or without nisin (Fig. 6.3). The OSU 799

strain was more sensitive than FDA strain for both pressure alone and nisin-pressure

combination treatment (P<0.05). An 8 log unit reduction was observed after a pressure

treatment at 500 MPa for FDA strain and 450 MPa for the OSU strain. When nisin was

184

Page 201: Calorimetric and Microbiological Evaluation of Bacteria

added, 8 log unit reduction was obtained at 400 MPa for FDA strain and 350 MPa for the

OSU strain, where the difference in log survivals between pressure alone and nisin-

combination treatments was greatest (Fig. 6.3). The difference of log survivals between

two strains in pressure alone treatment was greatest (3 log) at 450 MPa, while the

difference was greatest (~2.5 log) at 350 MPa in pressure-nisin combination treatment.

The difference of log survivals between the cells of two strains treated by a given

pressure level (except for 300 MPa level) was greater for pressure-nisin combination

treated cells than for pressure-alone treated cells (Fig. 6.3).

185

Page 202: Calorimetric and Microbiological Evaluation of Bacteria

Viable counts (cfu/ml) Apparent enthalpy (J/g)

S. Enteritidis OSU S. Enteritidis FDA S. Enteritidis OSU S. Enteritidis FDAPressure

(MPa) 0 Nisaplin 200 IU/ml

Nisaplin 0 Nisaplin 200 IU/ml

Nisaplin 0 Nisaplin 200 IU/ml

Nisaplin 0 Nisaplin 200 IU/ml

Nisaplin 0.1 1.7 x 109 1.4 x 109 1.3 x 109 1.2 x 109 3.99 3.91 3.96 3.91

150 1.7 x 109 5.6 x 108

3.80 3.40

200 1.3 x 109 8.2 x 107 1.2 x 109 3.0 x 108 3.35 2.88 3.65 2.92

250 1.4 x 108 1.4 x 106 4.8 x 108 1.4 x 107 3.06 2.60 3.62 2.81

300 3.6 x 105 1.8 x 103 1.3 x 107 2.1 x 104 2.55 2.27 3.00 2.31

350 4.9 x 103 < 1.0 x 101 6.5 x 104 1.2 x 102 2.11 1.89 2.82 2.12

400 5.0 x 101 1.5 x 104 < 1.0 x 101 1.76 1.75 2.46 1.75

450 < 1.0 x 101 1.9 x 103 1.74 2.18

500 < 1.0 x 101 1.74 1.65

550 1.73

186

Table 6.1. Viability and apparent enthalpy values for cells of Salmonella Enteritidis strains after HHP treatments.

Page 203: Calorimetric and Microbiological Evaluation of Bacteria

0.000000001

0.00000001

0.0000001

0.000001

0.00001

0.0001

0.001

0.01

0.1

1

0 100 200 300 400 500

N/N

0

Figure 6.3. Pressure dependence of fracdetermined by plate count. The cells trstrain ♦). The cells treated with 200 IU/m

Pressure (MPa)

tional viability of Salmonella Enteritidis strains eated without nisin (FDA strain ▲, OSU 799 l nisin (FDA strain ∆, OSU 799 strain ◊).

187

Page 204: Calorimetric and Microbiological Evaluation of Bacteria

Evaluation of inactivation Salmonella Enteritidis from apparent enthalpy data after HHP treatment in combination with nisin The effect of the pressure-nisin combination treatment on two strains was evaluated

by comparing apparent enthalpies of pressure-nisin treated cells with those of pressure

alone treated cells (Table. 6.1). No apparent differences found between thermograms of

control and the nisin alone treated cells for both strains (Figs 6.4a,b Thermogram A and

Figs 6.5a,b thermogram A). Similar to results of plate count method, the reduction of

apparent enthalpy was greater for combined nisin-pressure treatment for both strains.

The decrease in apparent enthalpy (∆H) appeared to be mainly due to reduction in the

ribosome denaturation peak, a, for both pressure alone and nisin-combination treatments

(Figs. 6.4a and 6.5a). As pressure increased, the greater decrease in the apparent

enthalpy of the pressure-nisin treated cells was observed due to the greater size reduction

of peak c as well as the greater decrease in the ribosome peak for both strains (Figs. 6.4b

and 6.5b).

Initial thermograms reveal an overlapping b (Tm 93oC) and c (Tm 100oC) peaks.

Peak, c, disappeared from thermograms revealing the endotherm b at 200 MPa and above

pressure treated cells on both strains (Figs. 6.4a and 6.5a). However, the peak c remained

(Tm 103oC) in the thermograms up to 300 MPa when nisin was added. Substantial

reduction of the size of the peak b was observed after a pressure treatment at 400 MPa for

FDA strain and 350 MPa for the OSU strain. When nisin was added, the peak b of FDA

strain seemed to be disappeared after 350 MPa treatment for FDA strain while an

endothermic transition still remained in the peak b area after 400 MPa for OSU strain.

188

Page 205: Calorimetric and Microbiological Evaluation of Bacteria

The fractional viability based on apparent enthalpy data [(∆H-∆Hf)/(∆H0-∆Hf)] was

plotted against pressure for two strains, which were pressurized with or without nisin

(Fig. 6.6). The apparent enthalpy (∆H) and the residual apparent enthalpy (∆Hf) are

obtained after treatment resulting in no viability and ∆H0 is the apparent enthalpy of

untreated cells. A residual apparent enthalpy (∆Hf) was reached at 500 MPa for FDA

strain and 450 MPa for OSU strain in the pressure treatment. The ∆Hf was obtained at

lower pressure levels when nisin was added, the value was reached at 400 MPa for FDA

strain and 350 MPa for OSU strain. The slope of plots in Figure 6.6 became much

steeper after 250 MPa alone and nisin combination treatments for FDA strain and after

150 MPa for OSU strain. About 80% reduction of the enthalpy fraction was observed

after a pressure treatment at 450 MPa for FDA strain and 350 MPa for the OSU strain.

When nisin was added, the similar reduction was obtained at 350 MPa for FDA strain and

300 MPa for the OSU strain.

Natural logarithm of fractional viability values calculated from calorimetric data

[(∆H-∆Hf)/(∆H0-∆Hf)] and plate count data (N/N0) are plotted in Figure 6.7 and 6.8. The

correlation coefficients of all four curves were greater than 0.93. The difference of the

slope of the lines between pressure only and pressure-nisin combination treated cells are

shown in the graph of FDA strain, while very similar slope of those are observed in the

graph of OSU strain.

189

Page 206: Calorimetric and Microbiological Evaluation of Bacteria

(a) (b)

20 40 60 80 100 120 140

0.4 mWIa

dcb

A

G

F

E

D

C

B

I

H

20 40 60 80 100 120 140

a

dcb

A

G

F

E

D

C

B

0.4 mWI

Temperature (oC) Temperature (oC)

Figure 6.4. DSC thermograms of Salmonella Enteritidis FDA pellets after pressure alone treatment (a) or pressure-nisin combination treatment (b). Control (A), 200 MPa (B), 250 MPa (C), 275 MPa (D), 300 MPa (E), 350 MPa (F), 400 MPa (G), 450 MPa (H) and 500 MPa (I). Thermograms are offset for clarity.

190

Page 207: Calorimetric and Microbiological Evaluation of Bacteria

(a) (b)

20 40 60 80 100 120 140

0.4 mWIa

dcb

A

G

F

E

D

C

B

I

H

20 40 60 80 100 120 140

0.4 mWIa

dcb

A

F

E

D

C

B

G

Temperature (oC) Temperature (oC)

Figure 6.5. DSC thermograms of Salmonella Enteritidis OSU 799 pellets after pressure alone treatment (a) or pressure-nisin combination treatment (b). Control (A), 150 MPa (B), 200 MPa (C), 250 MPa (D), 300 MPa (E), 350 MPa (F), 400 MPa (G), 450 MPa (H) and 500 MPa (I). Thermograms are offset for clarity.

191

Page 208: Calorimetric and Microbiological Evaluation of Bacteria

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 100 200 300 400 500

[(∆H

-∆H

f)/(∆

H0-∆

Hf)]

Figure 6.6. Pressure dependence of fradetermined by calorimetric data. The cestrain ♦). The cells treated with 200 IU

Pressure (MPa)

ctional viability of Salmonella Enteritidis strains lls treated without nisin (FDA strain ▲, OSU 799

/ml nisin (FDA strain ∆, OSU 799 strain ◊).

192

Page 209: Calorimetric and Microbiological Evaluation of Bacteria

y = 0.0888x + 0.0584R2 = 0.9323

y = 0.1014x + 0.236R2 = 0.9398

0

0.5

1

1.5

2

2.5

0 5 10 15 20

-ln[

(∆H

-∆H

f)/(∆

H0-∆

Hf)]

-ln(N/N0)

Figure 6.7. Correlation between fractional apparent enthalpy and fractional viability for Salmonella Enteritidis FDA after HHP treatment. The cells treated without nisin (♦). The cells treated with 200 IU/ml nisin (◊).

193

Page 210: Calorimetric and Microbiological Evaluation of Bacteria

y = 0.1225x + 0.133R2 = 0.9644

y = 0.1151x + 0.1865R2 = 0.9505

0

0.5

1

1.5

2

2.5

0 5 10 15 20

-ln[

(∆H

-∆H

f)/(∆

H0-∆

Hf)]

-ln(N/N0)

Figure 6.8. Correlation between fractional apparent enthalpy and fractional viability for Salmonella Enteritidis OSU 799 after HHP treatment. The cells treated without nisin (♦). The cells treated with 200 IU/ml nisin (◊).

194

Page 211: Calorimetric and Microbiological Evaluation of Bacteria

DISCUSSION

For successful commercialization of HHP, a pressure processing protocol which will

employ the combination of low to intermediate pressure with antimicrobial agents and will

achieve a high level of destruction of foodborne pathogens must be developed

(Kalchayanand et al., 1998; Masschalck et al., 2001). Nisin has been approved as a food

preservative world wide. Therefore, commercially available nisin, Nisaplin, was used in

present study to evaluate the effect of pressure-nisin combination treatment on

inactivation of foodborne bacteria. Although the extent of bacterial death is directly

proportional to the pressure level, resistance to high pressure varies among the species and

among strains of the same species (Alpas et al., 1999; Robey et al., 2001; Alpas et al., 2003).

Therefore, the differences in pressure sensitivities of two different strains of Salmonella

Enteritidis were evaluated after HHP treatment with or without nisin.

The denaturation of ribosome has been considered as an important factor in death of

bacteria by both heat (Mackey et al., 1993; Mohacsi-Farkas et al., 1999; Lee and

Kaletunç, 2002b) and HHP treatments (Niven et al., 1999; Alpas et al., 2003). This study

confirms the findings of previous DSC studies that the change in the peak areas (apparent

enthalpy, ∆H) associated with denaturation of ribosome are correlated to loss of viability

in microorganisms as a function of pressure (Niven et al., 1999; Alpas et al., 2003). The

greater survivability of Salmonella Enteritidis FDA strain observed from plate counting

data (Fig. 6.3) and apparent enthalpy data (Fig. 6.6) subsequent to HHP treatment

indicates that the FDA strain have a relatively higher resistance to the pressure treatment

195

Page 212: Calorimetric and Microbiological Evaluation of Bacteria

in comparison to Salmonella Enteritidis OSU 799 strain. Comparison of thermal

stabilities of ribosomal subunits for two control strains showed higher (~2oC) thermal

stability for the FDA strain. Thermal stability of ribosome has been shown to be related

to thermal resistance of bacterial cells (Mackey et al., 1993; Lee and Kaletunç, 2002b).

Thermal inactivation study on E. coli and Lactobacillus plantarum showed that the

greater thermal stability of ribosome in E. coli coincided with the greater resistance of E.

coli to heat treatment (Chapter 2). However, a recent study (Alpas et al., 2003) on

comparison of pressure sensitivities of two strains of Staphylococcus aureus

demonstrated that there is no clear relationship between the thermal stability of ribosome

and pressure resistance of bacteria. Alpas et al. (2003) observed that while S. aureus 765

has a higher thermal stability of ribosome than S. aureus 485, S. aureus 485 is more

resistant to pressure than S. aureus 765. Therefore, the relationship between the thermal

stability of ribosome and pressure resistance of bacterial strains may vary with the

different organisms.

DSC thermograms of two Salmonella Enteritidis strains also showed two

overlapping endotherms as peak d for FDA strain while a single endotherm occurred as

peak for OSU 799 strain. The endotherms had a temperature range of 115 to 130oC for

both strains are described as the denaturations of outer membrane components of Gram-

negative bacteria (Mackey et al., 1991; Lee and Kaletunç, 2002b).

The plate count results (Table 6.1) indicate that nisin itself is ineffective against

Salmonella Enteritidis strains. The finding agrees with the theory that nisin alone

treatment is ineffective against Gram-negative bacteria due to nisin-impermeable barrier

196

Page 213: Calorimetric and Microbiological Evaluation of Bacteria

(outer membrane) in their cell wall (Kordel and Sahl, 1986; Delves-Broughton, 1990).

However, the susceptibility of the bacterial cells to nisin increased with pressure level

increased. It has been suggested that HHP causes alterations of structures in the outer

membrane of the cells thereby facilitating penetration of small water soluble proteins

such as nisin and lysozyme into the Gram-negative cells via permeabilized outer

membrane (Hauben et al., 1996; Kalchayanand et al. 1998). The pressure-induced

alteration of outer membrane components was confirmed by the leakage of a periplasmic

enzyme at 110 MPa and above treatments (Hauben et al., 1996) and the dissociation of

metal ions such as Ca2+ and Mg2+, which are involved in stabilization of outer membrane,

at 220 MPa and above treatments on E. coli (Hauben et al., 1997). However, there were

no apparent changes in the transitions of the outer membrane components (peak d)

detected in the thermograms of both strains after the pressure-nisin treatments (Figs. 6.4a

and 6.5b) indicating pressure-induced conformational change of the outer membrane

might be reversible due to immediate recovery after the release of pressure or nisin into

the cells. Hauben et al. (1996) reported that pressure-mediated change in permeability of

the outer membrane was not permanent at least up to 320 MPa pressure levels on E. coli

cells, and the cells regained their resistance to nisin immediately upon relief of pressure.

Another possible explanation for no change in the outer membrane transition in the

thermograms is that the energy requirements for the thermal events in outer membrane

components were not great enough to change the size of the transition.

The comparison of DSC data for pressurized (≥ 200 MPa) cells shows that the lack

of endotherm b, which is proposed to be transition of DNA denaturation, in the

197

Page 214: Calorimetric and Microbiological Evaluation of Bacteria

thermograms of pressurized cells in the presence of nisin indicates that DNA might be

irreversibly damaged by the penetrated nisin in combination with the moderate HHP.

The absence of peak b and the reduction of ribosomal peaks (Figs. 6.4 and 6.5)

significantly (P<0.05) decreased apparent enthalpy (∆H) for pressure-nisin combination

treated cells than that for pressure alone treated cells. The difference of ∆H was greatest

at 250 MPa for FDA strain and 200 MPa for OSU 799 strain where a decrease in ∆H of

the pressure-nisin treated cells occurred due to the disappearance of the DNA peak (Fig

6.6). On the other hand, the difference of N/N0 increased as pressure increased and was

greatest at 400 MPa for FDA strain and 350 MPa for OUS 799 strain where tails on the

survival curves started (Fig. 6.3). Above results suggest the effectiveness of nisin

addition on the reduction of apparent enthalpy is higher at the low pressure treatment

levels while that on viability loss from plate count method is higher at the high pressure

treatment levels. These findings should be considered when apparent enthalpy data is

used to determine viability of bacteria after the pressure-nisin combination treatment.

The pressure levels less than 200 MPa enhance the stability of synthetic double

stranded DNA when the thermal stability of the DNA is higher than 50oC (Dubins et al.,

2001). Since the thermal stability (Tm) of bacterial DNA is generally 95oC, we may

expect that there is no damage on DNA structures during moderate HHP treatment.

Because little information available in literature, it is difficult to explain about the

mechanism of pressure-nisin combination treatment which irreversibly changes transition

of DNA (peak b) in DSC thermograms after 200 MPa for FDA strain and 150 MPa for

OSU strain (Figs. 6.4b and 6.5b). Possible hypothesis can be the additional dropping of

198

Page 215: Calorimetric and Microbiological Evaluation of Bacteria

intracellular pH (pHi) by addition of nisin in the HHP treatment. Wouters et al. (1998)

reported that pHi of L. plantarum dropped from 7 to 5.5 after the cells pressurized at 250

MPa for 10 min while external pH was maintained at pH 5.5 during the pressurization. It

is known that HHP treatment drops pHi by damaging membrane bound ATPase which

regulates the efflux of protons (Cheftel, 1995). There is an evidence that pHi can also be

lowered by nisin treatment. In the study of the interaction between Listeria

monocytogenes and nisin, Budde and Jakobsen (2000) observed that pHi the cell

decreased from 7.9 to 5.5 in the nisin treatment (500 IU/ml, Nisaplin) for 12 min while

external pH was maintained at pH 5.5. It has been reported that the proton motive force

can be collapsed due to the dissipation of the membrane potential and the pH gradient

after nisin-induced pore formation in cytoplasmic membrane (Bruno et al., 1992; Chung

et al., 2000). DNA structure tends to be denatured in acidic conditions due to elimination

of hydrogen bonding between single strands (Cooper, 1997). Therefore, it is possible that

the damage of cytoplasmic membrane, which leads to affecting DNA by lowering pHi, is

greater when nisin is applied in HHP treatment. Further study on the measurement of pHi

of cells treated with pressure-nisin combination might be helpful to confirm above

hypothesis. In the comparison of DNA transitions in the thermograms of two strains,

there were no noticeable transition temperature and size differences between the

transitions of both DNA before and after treatment (peak b, Figs. 6.2, 6.4, and 6.5). The

result suggests that the thermal stabilities and reversibility following to pressure or/and

nisin treatments of the DNA for both strains are similar.

199

Page 216: Calorimetric and Microbiological Evaluation of Bacteria

The plots in Figure 6.3 and 6.6 suggest that the fraction of apparent enthalpy [(∆H-

∆Hf)/(∆H0-∆Hf)] from DSC data is correlated to the log of fractional viability (N/N0) from

plate count data subsequent to pressure-nisin treatments for both strains. Previous

research showed that the apparent enthalpy is related to the fractional survivors after a

thermal treatment (Chapters 2 and 3). A linear relationship in the plots of [(∆H-

∆Hf)/(∆H0-∆Hf)] versus (N/N0) (Fig. 6.7 and 6.8) is observed after taking natural

logarithm. The slope of the lines from the fractions derived from the two data is lower

for pressure only treated cells than for pressure-nisin combination treated cells on FDA

strain (Fig. 6.7). The result indicates that the fraction values of apparent enthalpy data

against the fraction of plate count data for pressure only treated cells are relatively higher

than those for pressure-nisin combination treated cells. The plots of pressure only treated

cells in Figure 6.7 shows that the disparity between the apparent enthalpy data and plate

count data is greatest at 350 MPa treatment. At 350 MPa treatment, the fraction of

viability greatly dropped from 300 MPa treatment (Fig. 6.3) while the fraction of

apparent enthalpy was not much decreased from 300 MPa treatment (Fig. 6.6) for

pressure only treated cells. Thermograms in Figure 6.4 show that the main reason for

less change in the fraction of apparent enthalpy for pressure only treated cells is due to

reversibility of DNA transition (peak b). Unlike thermograms from three other treated

cells, in which the size of DNA transition is gradually decreased as the treatment pressure

increased, the DNA transition in the thermogram from pressure only treated cells of FDA

strain did not change up to 350 MPa treatment.

200

Page 217: Calorimetric and Microbiological Evaluation of Bacteria

Moderate pressures (200 to 400 MPa) increased the susceptibility of both relatively

pressure-resistant and -sensitive Salmonella Enteritidis strains to nisin, although onset of

the susceptibility occurred in lower pressure level for the pressure-sensitive strain. The

pressure might cause alterations in the outer membrane of the cells thereby facilitating

penetration of nisin into the cell. Comparison of various final states achieved under

different treatment conditions starting from same initial state using DSC data allowed us

to investigate the effectiveness of combination treatment of HHP with the antimicrobial

agent for inactivation of bacteria. The apparent enthalpy data obtained from DSC can be

used to evaluate pressure levels necessary to reduce a microbial population in the

presence of nisin and provide information about viability in significantly shorter time

with comparable accuracy to plate count. Furthermore, DSC helps to identify changes in

cellular components as a function of treatment conditions. DNA might be irreversibly

damaged by the combination treatments. The finding of this combination treatment can

allow the rational design of HHP processing protocols for manufacture of

microbiologically safe and minimally processed food products.

201

Page 218: Calorimetric and Microbiological Evaluation of Bacteria

REFERENCES

Alpas, H., Kalchayanand, N., Bozoglu, F., Sikes, A., Dunne, C.P. and Ray, B. 1999. Variation in resistance to hydrostatic pressure among strains of food-borne pathogens. Appl. Environ. Microbiol. 65:4248-4251. Alpas, H., Kalchayanand, N., Bozoglu, F. and Ray, B. 2000. Interactions of high hydrostatic pressure, pressurization temperature and pH on death and injury of pressure-resistant and pressure-sensitive strains of foodborne pathogens. Int. J. Food Microbiol. 60:33-42. Alpas, H., Lee, J., Bozoglu, F. and Kaletunç, G. 2003. Differential scanning calorimetry of pressure-resistant and pressure-sensitive strains of Staphylococcus aureus and Escherichia coli O157:H7. Int. J. Food Microbiol. 87:229-237. Anderson, W.A., Hedges, N.D., Jones, M.V. and Cole, M.B. 1991. Thermal inactivation of Listeria monocytogenes studied in differential scanning calorimetry. J. Gen. Microbiol. 137:1419-1424. Belliveau, B.H., Beaman, T.C., Pankratz, H.S. and Gerhardt, P. 1992. Heat killing of bacterial spores analyzed by differential scanning calorimeter. J. Bacteriol. 174: 4463-4474. Benito, A., Ventoura, G., Casadei, M., Robinson, T. and Mackey, B. 1999. Variation in resistance of natural isolates of Escherichia coli O157 to high hydrostatic pressure, mild heat, and other stresses. Appl. Environ. Microbiol. 65:1564-1569. Bruno, M.E., Kaiser, A. and Montville, T.J. 1992. Depletion of proton motive force by nisin in Listeria monocytogenes cells. Appl. Environ. Microbiol. 58:2255-2259. Budde, B.B. and Jakobsen, M. 2000. Real-time measurements of the interaction between single cells of Listeria monocytogenes and nisin on a solid surface. Appl. Environ. Microbiol. 66:3586-3591.

202

Page 219: Calorimetric and Microbiological Evaluation of Bacteria

Cheftel, J.C. 1995. High pressure, microbial inactivation and food preservation. Food Sci. Technol. 1:75-90. Chung, H.-J., Montville and Chikindas, M.L. 2000. Nisin depletes ATP and proton motive force in mycobacteria. Lett. Appl. Microbiol. 31:416-420. Delves-Broughton, J. 1990. Nisin and it application as a food preservative. J. Soc. Dairy Technol. 43:73-76. Dubins, D.N., Lee, Adrian., Macgregor, R.B. and Chalikian, T.V. 2001. On the stability of double stranded nucleic acids. J. Am. Chem. Soc. 123:9254-9259. Earnshaw, R.G., Appleyard, J. and Hurst, R.M. 1995. Understanding physical inactivation processes: combined preservation opportunities using heat, ultrasound and pressure. Int. J. Food Microbiol. 28:197-219. Farkas, D.F. and Hoover, D.G. 2000. High pressure processing. J. Food Sci. 65(Supplement):47-64. Garcia-Graells, C., Masschalck, B. and Michiels, C.W. 1999. Inactivation of Escherichia coli in milk by high-hydrostatic pressure treatment in combination with antimicrobial peptides. J. Food Prot. 62:1248-1254. Gross, M. and Jaenicke, R. 1994. Proteins under pressure. The influence of high hydrostatic pressure on structure, function and assembly of proteins and protein complexes. Eur. J. Biochem. 221:617-630. Hancock, R.E.W. 1984. Alteration in outer membrane permeability. Annu. Rev. Microbiol. 38:237-264. Hass, G.J., Prescott, H.E., Dudley, E., Dik, R., Hintlan, C. and Keane, L. 1989. Inactivation of microorganisms by carbon dioxide under pressure. J. Food Safety. 9:253-265.

203

Page 220: Calorimetric and Microbiological Evaluation of Bacteria

Hauben, K., Bartlett, D.H., Soontjens, C., Cornelis, K., Wuytack, E. and Michiels, C. 1997. Escherichia coli mutants resistant to inactivation by high hydrostatic pressure. Appl. Environ. Microbiol. 63:945-950. Hauben, K., Wuytack, E., Soontjens, C. and Michiels, C. 1996. High pressure transient sensitization of Escherichia coli to lysozyme and nisin by disruption of outer-membrane permeability. J. Food Prot. 59:350-355. Hoover, D.G., Metrick, C., Papineau, A.M., Farkas, D.F. and Knorr, D. 1989 Biological effects of high hydrostatic pressure on food microorganisms. Food Technol. 43:99-107. Kalchayanand, N., Sikes, A., Dunne, C.P. and Ray. B. 1998. Factors influencing death and injury of foodborne pathogens by hydrostatic pressure-pasteurization. Food Microbiol. 15:207-214. Kaletunç, G. 2001. Thermal analysis of bacteria using differential scanning calorimetry. In Novel Process and Control Technologies in the Food Industry ed. Bozoglu, F., Deak, T. and Ray, B. pp. 227-235. Amsterdam: IOS press. Kaletunç, G., Lee, J., Alpas, H. and Bozoglu, F. 2004. Evaluation of structural changes induced by high hydrostatic pressure in Leuconostoc mesenteroides. Appl. Environ. Microbiol. 70:1116-1122. Knorr, D. 1993. Effect of high hydrostatic pressure process on food safety and quality. Food Technol. 47(6):156-161. Kordel, M., and Sahl, H.G. 1986. Susceptibility of bacterial, eukaryotic and artificial membranes to the disruptive action of cationic peptide pep5 and Nisin. FEMS Microbiol. Lett. 34:139. Lee, J. and Kaletunç, G. 2002a. Calorimetric determination of inactivation parameters of microorganisms. J. Appl. Microbiol. 93:178-189.

204

Page 221: Calorimetric and Microbiological Evaluation of Bacteria

Lee, J. and Kaletunç, G. 2002b. Evaluation by differential scanning calorimetry of the heat inactivation of Escherichia coli and Lactobacillus plantarum. Appl. Environ. Microbiol. 68:5379-5386. Linton, M., McClements, J.M.J. and Patterson, M.F. 2001. Inactivation of pathogenic Escherichia coli in skimmed milk using high hydrostatic pressure. Inno. Food. Sci. Emer. Technol. 2:99-104. Mackey, B.M., Miles, C.A., Parsons, S.E. and Seymour, D.A. 1991. Thermal denaturation of whole cells and cell components of Escherichia coli examined by differential scanning calorimetry. J. Gen. Microbiol. 137:2361-2374. Mackey, B.M., Miles, C.A., Seymour, D.A. and Parsons, S.E. 1993. Thermal denaturation and loss of viability in Escherichia coli and Bacillus stearothermophilus. Lett. Appl. Microbiol. 16: 56-58. Masschalck, B., Garcia-Graells, C., Haver, E.V. and Michiels, C.W. 2000. Inactivation of high pressure resistant Escherichia coli by lysozyme and nisin under high pressure. Inno. Food Sci. Emer. Technol. 1:39-47. Massachalck, B., Van Houdt, R. and Michiels, C.W. 2001. High pressure increases bacterial activity and spectrum of lactoferrin, lactoferricin and nisin. Int. J. Food Microbiol. 64:325-332. Mertens, B., and G. Deplace. 1993. Engineering aspects of high pressure technology in the food industry. Food Technol. 47:164-169. Miles, C.A., Mackey, B.M. and Parsons, S.E. 1986. Differential Scanning Calorimetry of Bacteria. J. Gen. Microbiol. 132: 939-952. Mohacsi-Farkas, Cs., Farkas, J., Meszaros, L., Reichart, O. and Andrassy. E. 1999. Thermal denaturation of bacterial cells examined by differential scanning calorimetry. J. Therm. Anal. Calor. 57:409-414.

205

Page 222: Calorimetric and Microbiological Evaluation of Bacteria

Mussa, D.M. and Ramaswamy, H. 1997. Ultra high pressure pasteurization of milk: kinetics of microbial destruction and changes in physico-chemical characteristics. Leben. Wisssen. Technol. 30:551-557. Niven, G.W., Miles, C.A. and Mackey, B.M. 1999. The effect of hydrostatic pressure on ribosome conformation in Escherichia coli: an in vivo study using differential scanning calorimetry. Microbiology 145:419-425. Patterson, M.F. and Kilpatrick, D.J. 1998. The combined effect of high hydrostatic pressure and mild heat on inactivation of pathogens in milk and poultry. J. Food Prot. 61:432-436. Patterson, M.F., Quinn, M., Simpson, R. and Gilmore. A. 1995. Sensitivity of vegetative pathogens to high hydrostatic pressure treatment in phosphate-buffered saline and foods. J. Food Prot. 58:524-529. Porretta, S., Birzi, A., Ghizzoni, C. and Vicini, E. 1995. Effects of ultra-high hydrostatic pressure treatments on the quality of tomato juice. Food Chem. 52:35-41. Ray, B., Kalchayanand, N., Dunne, P. and Sikes, A. 2001. Microbial destruction during hydrostatic pressure processing of food. In Novel Process and Control Technologies in the Food Industry ed. Bozoglu, F., Deak, T. and Ray, B. pp. 95-122. IOS press: Amsterdam. Roberts, C.M. and Hoover, D.G. 1996. Sensitivity of Bacillus coagulans spores to combinations of high hydrostatic pressure, heat, acidity and nisin. J. Appl. Bacteriol. 81:363-368. Robey, M., Benito, A., Huston, R.H., Pascual, C., Park, S.F. and Mackey, B.M. 2001. Variation in resistant to high hydrostatic pressure and rpoS heterogeneity in natural isolates of Escherichia coli O157:H7. Appl. Environ. Microbiol. 67:4901-4907. Rodriguez-Torres, A., Ramos-Sanchez, M. C., Orduna-Domingo, A., Martin-Gil, F. J. and Martin-Gil, J. 1993. Differential scanning calorimetry investigations on LPS and free lipids A of the bacterial cell wall. Res. Microbiol. 144:729-740.

206

Page 223: Calorimetric and Microbiological Evaluation of Bacteria

Sahl, H-G. and Bierbaum, G. 1998. Lantibiotics: biosynthesis and biological activities of uniquely modified peptides from Gram-positive bacteria. Annu. Rev. Microbiol. 52:41-79. Trujillo, A.J., Royo, C., Ferragut, V. and Guamis, B. 1999. Ripening profiles of goat cheese produced from milk treated with high pressure. J. Food Sci. 64(5):833-837. Wouters, P.C., Glaasker, E. and Smelt, J.P.P. 1998. Effects of high pressure on inactivation kinetics and events related to proton efflux in Lactobacillus plantarum. Appl. Environ. Microbiol. 64:509-514. Yuste, J., Mor-Mur, M., Capellas, M., Guamis, B. and Pla, R. 1998. Microbiological quality of mechanically-recovered poultry meat treated with high hydrostatic pressure and nisin. Food Microbiol. 15:407-414.

207

Page 224: Calorimetric and Microbiological Evaluation of Bacteria

GENERAL CONCLUSIONS

The main goal of this study was to evaluate the effectiveness of physical and

chemical food preservation treatments for inactivation of bacteria using DSC. The

treatments studied included heat (linearly increasing temperature), chemical agents (HCl,

acetic acid, ethanol, NaCl, and nisin), and HHP. DSC of the whole bacterial cells

allowed us the detection of treatment induced changes in their cellular components

including ribosomes, nucleic acids, and cell wall. The changes in transition of ribosomal

subunits and the loss of cell viability appear to be related.

Since thermal processing has been the main choice of food preservation for

inactivation of bacteria in order to produce a safe product, the evaluation of thermal

sensitivity is important. The lower onset and peak temperatures of ribosomal transitions

of the L. plantarum thermogram indicate that the thermal stabilities of L. plantarum

ribosomes are lower than those of E. coli ribosomes. The viability loss was related to the

apparent enthalpy change for both organisms and L. plantarum cells were more sensitive

to heat treatment. The findings suggest heat resistance of bacteria is related to the onset

temperature and the thermal stability of the ribosomal subunit transition.

The amount of thermal energy (apparent enthalpy, ∆H) associated with denaturation

of cellular components due to heat treatment applied in DSC under linearly increasing

208

Page 225: Calorimetric and Microbiological Evaluation of Bacteria

temperature conditions was found to be related to the number of viable cells of E. coli.

Similar D and z values were calculated from plate count data and calorimetric data,

indicating that apparent enthalpy obtained from calorimetric data can be used to

determine the kinetic inactivation parameters. Use of DSC data to determine kinetic

parameters can be accomplished in shorter time in comparison with plate count method

because this approach eliminates incubation time of plates thereby saving more than two

days to produce results.

Mild chemical treatment on E. coli cells resulted in a decrease in the onset and peak

temperature of ribosome transition. The transition of DNA was irreversibly affected after

HCl (pH≤4) and acetic acid (≥0.5%) treatments. Reduced survivability of chemically

treated cells in subsequent heat treatment in comparison with untreated cells indicates

mild chemical treatments affect the thermal stability of ribosomal subunits in the cell,

thereby increasing the sensitivity of bacteria to heat treatment. The thermal sensitivity

was greater for acid-treated cells due to both ribosome and DNA structures were

irreversibly affected after the treatments. The result can support the “hurdle technology”

concept, in which mild heating in conjunction with chemical agents have been utilized to

reduce processing requirements.

The effect of high hydrostatic pressure treatment, a potential non-thermal food

treatment, on E. coli was evaluated using DSC. The decrease in apparent enthalpy with

increasing pressure was mainly due to the reduction of ribosome and DNA peak size.

Comparison of the transmission electron micrographs indicated that the structures of

ribosome and DNA were damaged during HHP treatment. A close relationship of

209

Page 226: Calorimetric and Microbiological Evaluation of Bacteria

fractional viability based on calorimetric data and plate count data indicates apparent

enthalpy data can be used to predict the viability of target bacteria during HHP

inactivation.

Salmonella Enteritidis is a Gram-negative bacterium, which has nisin-impermeable

barrier (outer membrane). Therefore, nisin treatment alone is not expected to inactivate

Salmonella Enteritidis. Recent HHP studies on Gram-negative bacteria showed that high

pressure combined with nisin increases the inactivation of bacteria in comparison with

HHP alone. Present study was performed to evaluate the changes in cellular components

using DSC and viability of relatively pressure-resistant and -sensitive Salmonella

Enteritidis strains after HHP treatment with and without nisin. Addition of nisin at

pressures of 200 to 400 MPa caused additional viability loss and change in DNA

transition for both strains, indicating the pressures might cause alterations in the outer

membrane of the cells thereby facilitating penetration of nisin into the cell membrane.

Addition of nisin also reduced the difference of pressure sensitivity between two strains.

There were close relationships of fractional viability based on calorimetric data and plate

count data for both strains suggest the apparent enthalpy data obtained from DSC can be

used to evaluate pressure levels necessary to reduce a microbial population in the

presence of nisin and provide information about viability.

Overall, DSC helps to identify changes in cellular components of bacteria as a

function of treatment conditions. The effects of thermal and non-thermal treatments can

be evaluated by comparing the corresponding thermograms of DSC before and after

treatment. The apparent enthalpy data obtained from DSC can be used to determine

210

Page 227: Calorimetric and Microbiological Evaluation of Bacteria

viability and the kinetic inactivation parameters for bacteria. The findings of this study

help to develop the design of food processing protocols for manufacture of

microbiologically safe and minimally processed food products.

211

Page 228: Calorimetric and Microbiological Evaluation of Bacteria

BIBLIOGRAPHY

Abee, T. and Wouters, J.A. 1999. Microbial stress response in minimal processing. Int. J. Food Microbiol. 50:65-91. Allwood, M.C. and Russell. A.D. 1967. Mechanism of the thermal injury in Staphylococcus aureus. I. Relationship between viability and leakage. Appl. Microbiol. 15:1266-1269. Alpas, H., Kalchayanand, N., Bozoglu, F., Sikes, A., Dunne, C.P. and Ray, B. 1999. Variation in resistance to hydrostatic pressure among strains of food-borne pathogens. Appl. Environ. Microbiol. 65(9):4248-4251. Alpas, H., Kalchayanand, N., Bozoglu, F. and Ray, B. 2000. Interactions of high hydrostatic pressure, pressurization temperature and pH on death and injury of pressure-resistant and pressure-sensitive strains of foodborne pathogens. Int. J. Food Microbiol. 60:33-42. Anderson, W.A., Hedges, N.D., Jones, M.V. and Cole, M.B. 1991. Thermal inactivation of Listeria monocytogenes studied in differential scanning calorimetry. J. Gen. Microbiol. 137:1419-1424. Andrieu, J., Laurent, M., Puaux, J.P. and Oshita, S. 1989. Thermal properties of unfrozen and frozen food gels determined by an automatic flash method. In Proceedings of the Fifth International Congress on Engineering and Food. ed. Spiess, W.E.L., Schubert, H. pp. 447-455. Elsevier Applied Science. Bach, D. and Chapman, D. 1980. Calorimetric studies of biomembranes and their molecular components. In Biological microcalorimetry ed. Beezer, A.E. pp. 275-309. Academic Press: London.

212

Page 229: Calorimetric and Microbiological Evaluation of Bacteria

Balny, C. and Masson, P. 1993. Effects of high pressure on proteins. Food Rev. Int. 9:611-628. Bearson, B. L., Wilson, L. and Foster, J.W. 1998. A low pH-inducible, PhoPQ-dependent acid tolerance response protects Salmonella typhimurium against inorganic acid stress. J. Bacteriol. 180:2409-2417. Belliveau, B.H., Beaman, T.C., Pankratz, H.S. and Gerhardt, P. 1992. Heat killing of bacterial spores analyzed by differential scanning calorimeter. J. Bacteriol. 174:4463-4474. Benito, A., Ventoura, G., Casadei, M., Robinson, T. and Mackey, B., 1999. Variation in resistance of natural isolates of Escherichia coli O157 to high hydrostatic pressure, mild heat, and other stresses. Appl. Environ. Microbiol. 65:1564-1569. Boer, H. and Koivula, A. 2003. The relationship between thermal stability and pH optimum studied with wild-type and mutant Trichoderma reesei cellobiohydrolase Cel7A. Eur. J. Biochem. 270:1-8. Bonincontro, A., Cinelli, S., Mengoni, M., Onori, G., Risuleo, G. and Santucci, A. 1998. Differential stability of E. coli ribosomal particles and free RNA towards thermal degradation studied by microcalorimetry. Biophys. Chem. 75:97-103. Bowler, K., Duncan, C.J., Gladwell, R.T. and Davison, T.F. 1973. Cellular heat injury. Comp. Biochem. Physiol. 45:441-450. Bowler, K. and Manning, R. 1994. Membrane as the critical targets in cellular heat injury and resistance adaptation. In Temperature adaptation of biological membranes ed. Cossins, A.R. pp. 185-203. Portland Press: London. Brul, S. and Coote, P. 1999. Preservative agents in foods – Mode of action and microbial resistance mechanisms. Int. J. Food Microbiol. 50:1-17.

213

Page 230: Calorimetric and Microbiological Evaluation of Bacteria

Cheftel, J.C. 1995. High pressure, microbial inactivation and food preservation. Food Sci. Technol. 1:75-90. Cheftel, J.C. and Culioli, J. 1997. Effects of high pressure on meat. Meat Sci. 46:211-236. Chowdhry, B.Z. and Cole, S.C. 1989. Differential scanning calorimetry: applications in biotechnology. Trends Biotechnol. 7:11-18. Condon, S., Garcia-Lopez, M.L., Otero, A. and Sala, F.J. 1992. Effect of culture agar, preincubation at low temperature and pH on the thermal resistance of Aeromonas hydrophila. J. Appl. Bacteriol. 72(4):322-326. Crawford, Y.J., Murano, E.A., Olson, D.G. and Shenoy, K. 1996. Use of high hydrostatic pressure and irradiation to eliminate Clostridium sporogenes in chicken breast. J. Food Prot. 59:711-715. Csonka, L.N. 1989. Physiological and genetic response of bacteria to osmotic stress. Microbiol. Rev. 53:121-147. Daniel, R.M. and Cowan, D.A. 2000. Biomolecular stability and life at high temperatures. Cell. Mol. Life Sci. 57:250-264. Datta, A.K. and Noyogi, S.K. 1976. Biochemistry and physiology of bacterial nucleases. Pro. Nucleic Acid Res. Mol. Biol. 17:271-308. Davis, B.D. 1990. Growth and death of bacteria. In Microbiology. ed. Davis, B.D., Dulbecco, R., Eisen, H.N. and Ginsberg, H.S. pp. 51-64. J. B. Lippincott Co.: Philadelphia, Pa. De Ley, J. 1970. Re-examination of the association between melting point, buoyant density and chemical base composition of deoxyribonucleic acid. J. Bacteriol. 101:738-754.

214

Page 231: Calorimetric and Microbiological Evaluation of Bacteria

Delves-Broughton, J. 1990. Nisin and it application as a food preservative. J. Soc. Dairy Technol. 43:73-76. Dove, W.F. and Davidson, N. 1962. Cation effects on the denaturation of DNA. J. Mol. Biol. 5:467-478. Ellar, D.J. 1978. Spore specific structures and their function. Symp. Soc. Gen. Microbiol. 28:295-334. Earnshaw, R.G., Appleyard, J. and Hurst, R.M. 1995. Understanding physical inactivation processes: combined preservation opportunities using heat, ultrasound and pressure. Int. J. Food Microbiol. 28:197-219. Farkas, D.F. and Hoover, D.G. 2000. High pressure processing. J. Food Sci. 65(Supplement):47-64. Flink, I and Pettijohn, D.E. 1975. Polyamines stabilize DNA folds. Nature London 253:62-63. Garcia-Graells, C., Masschalck, B. and Michiels, C.W. 1999. Inactivation of Escherichia coli in milk by high-hydrostatic pressure treatment in combination with antimicrobial peptides. J. Food Prot. 62:1248-1254. Gervilla, R., Capellas, M., Ferragut, V. and Guamis, B. 1997. Effect of high hydrostatic pressure in Listeria innocua 910 CECT inoculated into ewe’s milk. J. Food Prot. 60:33-37. Gesteland, R.F. 1993. The RNA world : The nature of modern RNA suggests a prebiotic RNA world, ed. Gesteland, R.F. and Atkins, J.F. Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY. Gutierrez, C., Abee, T. and Booth, I.R. 1995. Physiology of the osmotic stress response in microorganisms. Int. J. Food Microbiol. 28:233-244.

215

Page 232: Calorimetric and Microbiological Evaluation of Bacteria

Hammond, S.M., Lambert, P.A. and Rycroft, A.N. 1984. Walls of Gram-positive bacteria. In The bacterial cell surface. pp. 29-56. Kapitan Szabo Publishers: Washington, DC. Hass, G.J., Prescott, H.E., Dudley, E., Dik, R., Hintlan, C. and Keane, L. 1989. Inactivation of microorganisms by carbon dioxide under pressure. J. Food Safety. 9:253-265. Hauben, K.J.A., Bartlett, D.H., Soontjens, C.C.F., Cornelis, K., Wuytack, E.Y. and Michiels, C.W. 1997. Escherichia coli mutants resistant to inactivation by high hydrostatic pressure. Appl. Environ. Microbiol. 63:945-950. Heremans, K. 1982. High pressure effects on proteins and other biomolecules. Annu. Rev. Biophys. Bioeng. 11:1-21. Hill, C., O’Driscoll, B. and Booth, E. 1995. Acid adaptation and food poisoning microorganisms. Int. Food Microbiol. 28:245-254. Hisao, C. and Siebert, K.J. 1999. Modeling the inhibitory effects of organic acids on bacteria. Int. J. Food Microbiol. 47:189-201. Hite, B.H. 1899. The effects of pressure in the preservation of milk. West Virginia Agri. Exp. station bulletin 58:15-35. Hohne, G.W.H., Hemminger, W. and Flammersheim, H.-J. 1996. Differential Scanning Calorimetry: An Introduction for Practitioners. Berlin; New York: Springer-Verlag. Hoover, D.G., Metrick, C., Papineau, A.M., Farkas, D.F. and Knorr, D. 1989. Biological effects of high hydrostatic pressure on food microorganisms. Food Technol. 43:99-107. Hurst, A. 1984. Reversible heat damage. In Repairable Lesions in Microorganisms ed. Hurst, A. and Nasim, A. pp. 303-318. Academic Press: London.

216

Page 233: Calorimetric and Microbiological Evaluation of Bacteria

Hurst, A. and Hughes, A. 1978. Stability of ribosomes of Staphylococcus aureus S6 sublethally heated in different buffers. J. Bacteriol. 133:564-568. Ingram, L.O. 1986. Microbial tolerance to alcohols: Role of the cell membrane. Trends Biotechnol. 4:40-44. Ingram, L.O. 1990. Ethanol tolerance in bacteria. Crit. Rev. Biotechnol. 9:305-319. Ingram, L.O. and Buttke, T. 1984. Effects of alcohols on microorganisms. Adv. Microbial Physiol. 25:253-300. Isaacs, N.S., Chilton, P. and Mackey, B. 1995. Studies on the inactivation of microorganisms by high pressure. In High pressure food processing ed. Ledward, D.A., Johnston, D.E., Earnshaw, R.G. and Hasting, A, Nottingham University Press. Jay, J.M. 1996. Intrinsic and extrinsic parameters of foods that affect microbial growth. In Modern Food Microbiology, 4th ed. pp. 38-68. Chapman and Hall, New York, NY. Jay, J.M. 1996. High temperature preservation and characteristics of thermophilic microorganisms. In Modern Food Microbiology, 4th ed. pp. 347-369. Chapman and Hall, New York, NY. Jung, H. 1986. A generalized concept for cell killing by heat. Radiat. Res. 106:56-72. Kaletunç, G. 2001. Thermal analysis of bacteria using differential scanning calorimetry. In Novel Process and Control Technologies in the Food Industry ed. Bozoglu, F., Deak, T. and Ray, B. pp. 227-235. IOS press: Amsterdam. Kalchayanand, N., Hanlin, M.B. and Ray, B. 1992. Sublethal injury makes Gram-negative and resistant Gram-positive bacteria sensitive to the bacteriocins, pediocin AcH and nisin. Lett. Appl. Microbiol. 15:239-243.

217

Page 234: Calorimetric and Microbiological Evaluation of Bacteria

Kalchayanand, N., Sikes, A., Dunne, C.P. and Ray. B. 1998. Factors influencing death and injury of foodborne pathogens by hydrostatic pressure-pasteurization. Food Microbiol. 15:207-214. Karatzas, A.K., Bennik, M.H.J., Smid, E.J. and Kets, E.P.W. 2000. Combined action of S-carbone and mild heat treatment on Listeria monocytogenes Scott A. J. Appl. Microbiol. 89:296-301. Knorr, D. 1993. Effects of high-hydrostatic-pressure processes on food safety and quality. Food Technol. 6:156-161. Kordel, M., and Sahl, H.G. 1986. Susceptibility of bacterial, eukaryotic and artificial membranes to the disruptive action of cationic peptide pep5 and nisin. FEMS Microbiol. Lett. 34:139. Kumar, A. 1995. Alternate view on thermal stability of the DNA duplex. Biochemistry 34:12922-12925. Kurganov, B.I., Lyubarev, A.E., Sanchez-Ruiz, J.M. and Shnyrov, V.L. 1997. Analysis of differential scanning calorimetry data for proteins criteria of validity of one-step mechanism of irreversible protein denaturation. 1997. Biophys. Chem. 69:125-135. Leistner, L. 2000. Basic aspects of food preservation by hurdle technology. Int. J. Food. Microbiol. 55:181-186. Leistner, L. and Gorris, L.G.M. 1995. Food preservation and hurdle technology. Trends Food Sci. Technol. 6:41-46. Lepock, J.R., Frey, H.E. and Inniss, W.E. 1990. Thermal analysis of bacteria by differential scanning calorimetry: relationship to protein denaturation in situ to maximum growth temperature. Biochim. Biophys. Acta 1055:19-26. Ludwig, H. and Schreck, C. 1997. The inactivation of vegetative bacteria by pressure. In High Pressure Research in the Bioscience and Biotechnology ed. Heremans, K. pp. 221-224. Leuven University Press, Leuven, Belgium.

218

Page 235: Calorimetric and Microbiological Evaluation of Bacteria

Mackey, B.M., Forestiere, K. and Issacs, N. 1995. Factors affecting the resistance of Listeria monocytogenes to high hydrostatic pressure. Food Biotechnol. 9:1-11. Mackey, B.M., Forestiere, N.S., Isaaca, R., Stenning, R. and Brooker, B. 1994. The effect of high hydrostatic pressure on Salmonella Thompson and Listeria monocytogenes examined by electron microscopy. Lett. Appl. Microbiol. 19:429-432. Mackey, B.M., Miles, C.A., Parsons, S.E. and Seymour, D.A. 1991. Thermal denaturation of whole cells and cell components of Escherichia coli examined by differential scanning calorimetry. J. Gen. Microbiol. 137:2361-2374. Mackey, B.M., Miles, C.A., Seymour, D.A. and Parsons, S.E. 1993. Thermal denaturation and loss of viability in Escherichia coli and Bacillus stearothermophilus. Lett. Appl. Microbiol. 16: 56-58. Mackey, B.M., Parsons, S.E., Miles, C.A. and Owen, R.J. 1988. The relationship between base composition of bacterial DNA and its intracellular melting temperature as determined by differential scanning calorimetry. J. Gen. Microbiol. 134:1185-1195. Maeda, Y., Noguchi, S. and Koga, S. 1974. Differential scanning calorimetric study of spontaneous germination of Bacillus megaterium spore by water vapor. J. Gen. Microbiol. 20:11-19. Masschalck, B., Garcia-Graells, C., Haver, E.V. and Michiels, C.W. 2000. Inactivation of high pressure resistant Escherichia coli by lysozyme and nisin under high pressure. Inno. Food Sci. Emer. Technol. 1:39-47. Masschalck, B., Van Houdt, R. and Michiels, C.W. 2001. High pressure increases bacterial activity and spectrum of lactoferrin, lactoferricin and nisin. Int. J. Food Microbiol. 64:325-332. Mertens, B., and Deplace, G. 1993. Engineering aspects of high pressure technology in the food industry. Food Technol. 47:164-169.

219

Page 236: Calorimetric and Microbiological Evaluation of Bacteria

Metrick, C., Hoover, D.G. and Farkas, D. 1989. Effects of high hydrostatic pressure on heat-resistant and heat-sensitive strains of Salmonella. J. Food. Sci. 54:1547-1549, 1564. Miles, C.A., Mackey, B.M. and Parsons, S.E. 1986. Differential scanning calorimetry of Bacteria. J. Gen. Microbiol. 132: 939-952. Mohacsi-Farkas, Cs., Farkas, J., Meszaros, L., Reichart, O. and Andrassy, E. 1999. Thermal denaturation of bacterial cells examined by differential scanning calorimetry. J. Therm. Anal. Calor. 57:409-414. Mohacsi-Farkas, Cs., Farkas, J. and Simon, A. 1994. Thermal denaturation of bacterial cells examined by differential scanning calorimetry. Acta Aliment. 23:157-168. Munns, R., Greenway, H., Setter, T.L. and Kuo, J. 1983. Tugor pressure, volumetric elastic modulus, osmotic volume and ultrastructure of Chlorella emersonii grown at high and low external NaCl. J. Exp. Bot. 34:144-155. Murray, R. G. E., Steed, R. and Elson, H. E. 1965. The location of the mucopeptide in sections of the cell wall of Escherichia coli and other gram-negative bacteria. Can. J. Microbiol. 11:547-560. Mussa, D.M. and Ramaswamy, H. 1997. Ultra high pressure pasteurization of milk: kinetics of microbial destruction and changes in physico-chemical characteristics. Leben. Wisssen. Technol. 30:551-557. Novak, J.S. and Juneja, V.K. 2001. Detection of heat injury in Listeria monocytogenes Scott A. J. Food Prot. 64:1739-1743. Ohshima, T., Ushio, H. and Koizumi, C. 1993. High-pressure processing of fish and fish products. Trends Food Sci. Technol. 4:370-375. Pace, B. and Campbell, L.L. 1967. Correlation of maximal growth temperature and ribosome heat stability. Proc. Natl. Acad. Sci. USA 57:1110-1116.

220

Page 237: Calorimetric and Microbiological Evaluation of Bacteria

Patterson, M.F. and Kilpatrick, D.J. 1998. The combined effect of high hydrostatic pressure and mild heat on inactivation of pathogens in milk and poultry. J. Food Prot. 61:432-436. Patterson, M.F., Quinn, M., Simpson, R. and Gilmour, A. 1995. Effect of high pressure on vegetative pathogens. In High pressure food processing ed. Ledward, D.A., Johnston, D.E., Earnshaw, R.G. and Hasting, A. Nottingham University Press. Paul, P., Chawala, S.P., Thomas, P. and Kesavan, P.C. 1997. Effect of high hydrostatic pressure, gamma-irradiation and combined treatments on the microbiological quality of lamb meat during chilled storage. J. Food Safety 16:263-271. Phale, P.S., Philippsen, A., Kiefhaber, R., Koebnik, R., Phale, V.P., Schirmer, T. and Rosenbusch, J.P. 1998. Stability of trimeric OmpF porin: The contributions of the latching loop L2. Biochemistry 37:15663-15670. Piper, P. 1997. The yeast heat shock response. In Yeast stress responses ed. Hohmann, S. and Mager, W. pp. 75-100. Molecular Biology Intelligence Unit; Springer. Poirier, I., Marechal, P.-A., Evrard. C. and Gervais, P. 1998. Escherichia coli and Lactobacillus plantarum responses to osmotic stress. Appl. Microbiol. Biotechnol. 50:704-709. Ponce, E., Pla, R., Mor-Mur, M., Gervilla, R. and Guamis, B. 1998. Inactivation of Listeria innocua inoculated in liquid whole egg by high hydrostatic pressure. J. Food Prot. 61:119-122. Popham, D.L., Helin, J., Costello, C.E. and Setlow, P. 1996. Analysis of the peptidoglycan structure of Bacillus subtilis endospores. J. Bacteriol. 178:6451-6458. Popper, L. and Knorr, D. 1990. Application of high-pressure homogenization for food preservation. Food Technol. 44:84-89.

221

Page 238: Calorimetric and Microbiological Evaluation of Bacteria

Porretta, S., Birzi, A., Ghizzoni, C. and Vicini, E. 1995. Effects of ultra-high hydrostatic pressure treatments on the quality of tomato juice. Food Chem. 52:35-41. Roberts, C.M. and Hoover, D.G. 1996. Sensitivity of Bacillus coagulans spores to combinations of high hydrostatic pressure, heat, acidity and nisin. J. Appl. Bacteriol. 81:363-368. Robey, M., Benito, A., Huston, R.H., Pascual, C., Park, S.F., and Mackey, B.M. 2001. Variation in resistant to high hydrostatic pressure and rpoS heterogeneity in natural isolates of Escherichia coli O157:H7. Appl. Environ. Microbiol. 67:4901-4907. Rodriguez-Torres, A., Ramos-Sanchez, M. C., Orduna-Domingo, A., Martin-Gil, F. J. and Martin-Gil, J. 1993. Differential scanning calorimetry investigations on LPS and free lipids A of the bacterial cell wall. Res. Microbiol. 144:729-740. Rosenthal, L.J. and Iandolo, J.J. 1970. Thermally induced intracellular alteration of ribosomal ribonucleic acid. J. Bacteriol. 103:333-335. Russell, A.D. 2003. Lethal effects of heat on bacterial physiology and structure. Sci. Prog. 86:115-137. Russell, A.D. and Harries, D. 1967. Some aspects of thermal injury in Escherichia coli. Appl. Microbiol. 15:407-410. Russell, A.D. and Harries, D. 1968. Damage to Escherichia coli on exposure to moist heat. Appl. Microbiol. 16:1394-1399. Saenger, W. 1984. Principles of the nucleic acid structure. Springer-Verlag: New York, NY. Salton, M.R.J. 1963. the relationship between the nature of the cell wall and the Gram strain. J. Gen. Microbiol. 30:233-235.

222

Page 239: Calorimetric and Microbiological Evaluation of Bacteria

Shadbolt, C.T., Ross, T. and McMeekin, T.A. 1999. Nonthermal death of Escherichia coli. Int. J. Food. Microbiol. 49:129-138. Shigehisa, T., Ohmori, T., Saito, A., Taji, S. and R. Hayashi. 1991. Effect of high pressure on characteristics of pork slurries and inactivation of microorganisms associated with meat and meat products. Int. J. Food Microbiol. 12:207-216. Simpson, R.K. and Gilmour, A. 1997. The effect of high hydrostatic pressure on Listeria monocytogenes in phosphate-buffered saline and model food systems. J. Appl. Microbiol. 83:181-188. Steim, J.M., Tourtellote, M.E., Reinert, J.C., McElhaney, R.N. and Rader, R.L. 1969. Calorimetric evidence for the liquid-crystalline state of lipids in a biomembrane. Pro. Nat. Acad. Sci. 63:104-109. Styles, M.F., Hoover, D.G. and Farkas, D.F. 1991 Response of Listeria monocytogenes and Vibrio parahaemolyticus to high hydrostatic pressure. J. Food Sci. 56:1404-1407. Suzuki, C. and Suzuki, K. 1962. The protein denaturation by high pressure. J. Biochem. 52:67-71. Szybalski, W. 1967. In Thermobiology ed. Rose, A.H. p. 73. Academic Press: London. Teixeira, P., Castro, H., Mohacsi-Farkas, C. and Kirby, R. 1997. Identification of sites of injury in Lactobacillus bulgaricus during heat stress. J. Appl. Microbiol. 83:219-226. Trujillo, A.J., Royo, C., Ferragut, V. and Guamis, B. 1999. Ripening profiles of goat cheese produced from milk treated with high pressure. J. Food Sci. 64:833-837. Tsuchido, T., Katsui, N., Takeuchi, A., Takano, M. and Shibasaki, I. 1985. Destruction of the outer membrane permeability barrier of Escherichia coli by heat treatment. Appl. Environ. Microbiol. 50:298-303.

223

Page 240: Calorimetric and Microbiological Evaluation of Bacteria

Verrips, C.T. and Kwast, R.H. 1977. Heat resistance of Citrobacter freundii in media with various water activities. Eur. J. Appl. Microbiol. 4: 225-231. Warth, A.D. 1985. Mechanisms of heat resistance. In Fundamental and Applied Aspects of Bacterial Spore ed. Dring, G.J., Ellar, D.J. and Gould, G.W. pp. 209-225. Academic Press: London. Weitzel, G., Pilatus, U. and Rensing, L. 1987. The cytoplasmic pH, ATP content and total protein synthesis rate during heat-shock protein inducing treatments in yeast. Exp. Cell Res. 170:64-79. Worcel, A. and Burgi, E. 1972. On the structure of the folded chromosome of E. coli. J. Mol. Biol. 71:127-147. Wright, A and Tipper, D.J. 1979. The outer membrane of Gram-negative bacteria. In The Bacteria. Vol. 7. ed. Sokatch, J.R. and Ornston, L.N. pp. 175-256. Academic Press: San Diego. Yuste, J., Mor-Mur, M., Capellas, M., Guamis, B. and Pla, R. 1998. Microbiological quality of mechanically-recovered poultry meat treated with high hydrostatic pressure and nisin. Food Microbiol. 15:407-414.

224

Page 241: Calorimetric and Microbiological Evaluation of Bacteria

APPENDIX

Figures and Table of the evaluation of Salmonella Enteritidis inactivation after HHP

treatment with different concentrations of nisin

225

Page 242: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120 140

0.2 mWI

350 MPa

300 MPa

275 MPa

250MPa

200 MPa

Control

Temperature (oC)

Figure Appendix.1. DSC thermograms of Salmonella Enteritidis FDA pellets after combinations of pressure and nisin (200 IU/ml Nisaplin) treatments.

226

Page 243: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120 140

0.2 mWI

350 MPa

300 MPa

275 MPa

250MPa

200 MPa

Control

Temperature (oC)

Figure Appendix.2. DSC thermograms of Salmonella Enteritidis FDA pellets after combinations of pressure and nisin (400 IU/ml Nisaplin) treatments.

227

Page 244: Calorimetric and Microbiological Evaluation of Bacteria

20 40 60 80 100 120 140

0.2 mWI

300 MPa

275 MPa

250MPa

200 MPa

Control

Temperature (oC)

Figure Appendix.3. DSC thermograms of Salmonella Enteritidis FDA pellets after combinations of pressure and nisin (600 IU/ml Nisaplin) treatments.

228

Page 245: Calorimetric and Microbiological Evaluation of Bacteria

0.000000001

0.00000001

0.0000001

0.000001

0.00001

0.0001

0.001

0.01

0.1

1

0 100 200 300 400 500

0 Nisin

200 IU/mlNisaplin

400 IU/mlNisaplin

600 IU/mlNisaplin

N/N

0

Figure Appendix.4. Pressure dependeFDA determined by plate count.

Pressure (MPa)

nce of fractional viability of Salmonella Enteritidis

229

Page 246: Calorimetric and Microbiological Evaluation of Bacteria

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 100 200 300 400 500

0 Nisin

200 IU/mlNisaplin

400 IU/mlNisaplin

600 IU/mlNisaplin

[(∆H

-∆H

f)/(∆

H0-∆

Hf)]

Figure Appendix.5. Pressure deFDA determined by calorimetric

Pressure (MPa)

pendence of fractional viability of Salmonella Enteritidis data.

230

Page 247: Calorimetric and Microbiological Evaluation of Bacteria

Nisaplin (IU/ml)

0 200 400 600

Pressure (MPa) N1 ∆H2 N ∆H N ∆H N ∆H

0 1.3 x 109 3.96 1.2 x 109 3.91 1.2 x 109 3.88 1.1 x 109 3.83

200 1.2 x 109 3.65 3.0 x 108 2.92 1.4 x 108 2.79 1.3 x 108 2.64

250 4.8 x 108 3.62 1.4 x 107 2.81 5.3 x 106 2.76 5.7 x 106 2.63

275 8.4 x 107 3.15 1.2 x 106 2.61 1.0 x 106 2.58 1.0 x 106 2.58

300 1.3 x 107 3.00 2.1 x 104 2.22 1.5 x 104 2.16 5.6 x 103 2.10

350 6.5 x 104 2.82 1.2 x 102 2.12 <1 x 101 1.75 1.3 x 109 4.65

400 1.5 x 104 2.46 <1 x 101 1.75 1.3 x 109 4.65 1.3 x 109 4.65

450 1.9 x 103 2.18 <1 x 101 4.65 1.3 x 109 4.65 1.3 x 109 4.65

500 <1 x 101 1.65 <1 x 101 4.65 1.3 x 109 4.65 1.3 x 109 4.65 1 Viable counts (cfu/ml) in culture 2 Apparent enthalpy of cell pellet DSC

Table Appendix.1. Viability and apparent enthalpy values for cells of Salmonella Enteritidis FDA after HHP treatments in combination with nisin.

231