‘ ulna 115?. l? 4*;“fi1mtfifi " “ " "’ ”it‘fififl'éif r ' I: ‘ ‘ . i' - . 5' [4325' .i v _ u :‘Wfi 1' :f ; .' " 3 .. m!" {3131:51. E“. _‘ N; _ . _ ..§.\.'~‘.-¢nl';..~:’”" ' ”1" " 't.‘ -:-- -.1. mm . w «w ’ w 4 .' ' '. . ‘A L5 J 4 . , a"? .1 . w w: . .‘. . . I. , . - ....__. .. up.-. V “'<~..:""~ '6‘ .f harm-u: '"P'n _.- - . 43...") 'v.v\~4 ,,.: . , ._ 44?.“ _ w ‘5‘:' I .qJJIY: r. 'A .. I “-7 4—. ' mac?” ‘ J .z.. m .1. u .35" ( ~ ~97 A~v-vo 1:"; . #9 4m. . ’1 L17": ‘ r ,_ ‘ o' ? iv ‘ L-‘u‘ffi !, . -‘——.——d—-. ; ‘ 77:29.4 3‘55: #:935- : 119 L "I; ‘L ‘73.; 74.11'1113-17- T4, .‘Hfifinflfi ‘ THESTS Z) IllllllllllllllllllllltlllllllIlllllllllllllllllllllll 3 1293 016826 This is to certify that the dissertation entitled GENERAL PRINCIPLES OF MEMBRANE FUNCTIONS IN ADAPTATION AND ADAPTABILITY OF BACTERIA TO EXTREME ENVIRONMENTS presented by Carol Ann Mindock has been accepted towards fulfillment of the requirements for PI”! . D . degree in mm \A l/i 2W Melli- profest r Date O‘ILU fir MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY M‘Chlgan State UnIversIty PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE 1193 WpGS-QM GENERAL PRINCIPLES OF MEMBRANE FUNCTIONS IN ADAPTATION AND ADAPTABILITY OF BACTERIA TO EXTREME ENVIRONMENTS By Carol Ann Mindock A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Biochemistry 1998 ABSTRACT GENERAL PRINCIPLES OF MEMBRANE FUNCTIONS IN ADAPTATION AND ADAPT ABILITY OF BACTERIA TO EXTREME ENVIRONMENTS By Carol Ann Mindock Based on the hypothesis that membranes function to propagate the available thermal energy spectrum to the cell interior, using an Arctic permafrost bacterium we further confirm that there is a regular (mathematical) relationship between the available thermal energy spectrum and the membrane acyl chain length distributions. These acyl chain length distribution curves are Gaussian, in which the mean and variance are determined by temperature, indicating that at least for Gram-positive organisms the acyl chain length selection process is random. The relationship between the variances of the distribution curves and their respective growth temperatures are shown to be linear over the broad permissible growth range of the organism with deviation from linearity at the extremes of its growth temperature limit. We show that this relationship is general for Gram-positive organisms using data from the literature. This suggests that knowing the temperature, one can predict the fatty acid composition over a temperature range for a particular bacterium. We also show that effects of chemical perturbants, such as solvents and antibiotics, alter the Gaussian acyl chain distribution curves in the same manner as increasing or decreasing temperatures depending on whether they have a fluidizing or rigidifying effect on the membrane, indicating that a temperature could be used to allow the bacterium to grow in the presence of normally unfavorable solvent concentrations. Since heat is the driving force for all cellular chemical processes, it was of interest to determine whether temperature affects the balance of metabolic processes in the cell and how this relates to membrane structure. NMR and l4C-labeling experiments show that the Arctic permafrost bacterium utilizes glucose much more slowly at 4°C than at 24°C. This bacterium makes primarily branched chain fatty acids, however, an increase in straight chains (in particular n-Cm) is seen as the growth temperature increases from 4°C to 30°C. This is consistent with the increased utilization of glucose at higher temperatures. Here we show that Arctic permafrost strain 45-3 does not respond to near freezing conditions by synthesizing so called compatible metabolites. Rather the adaptations seen are changes in cell surface compositions, namely alterations in membrane acyl chains, the change to nonpolar headgroups, and a reduction in the amount of polysaccharide capsule. At the other extreme of temperature, we further characterize the ability of Sarcina ventriculi to form transmembrane fatty acids in response to any fluidizing effect on its membrane. We show that exogenously added fatty acids can be joined tail—to-tail at their (o-l positions to fatty acids in S. ventriculi membrane preparations in the presence of a low molecular weight ketone. This coupling is independent of the acyl chain headgroup as shown by the coupling of exogenously added n-heptane and hexadecylpyridinium chloride. The process is suppressed if cells are cultured in deuterium oxide indicating that the trigger is a kinetic factor related to motional dynamics. These results and results from NMR spectroscopy indicate that the process is regulated by motional dynamics and occurs via a radical mechanism in which a hydrogen atom is extracted from each acyl chain and the two acyl radicals join to form a new carbon-carbon bond. Further work is being done to isolate possible enzymes involved in this mechanism. Dedicated to my parents for all of their love and support. iv ACKNOWLEDGIVIENTS In the many years it has taken me to accomplish this goal, I have had a great deal of support and encouragement from a lot of wonderful people whom I wish to thank. First and foremost, I wish to thank my mentor and friend Dr. Rawle I. Hollingsworth. I feel privileged to have learned science and research under his guidance. He has taught me to think beyond the current views of science by applying fundamental concepts of math, physics, and chemistry to living systems. His continual support and encouragement as well as his friendship have been instrumental in helping me reach this goal. I also want to thank the members of my committee: Dr. William Wells, Dr. Estelle McGroarty, Dr. J. Gregory Zeikus, and Dr. Melvin Yokoyama for their helpful advice and suggestions regarding this dissertation project. I want to thank past and present members of the Hollingsworth group for all of their help, advice, support, and for making the lab such a fun place to work. I have alsoenjoyed both the scientific and cultural discussions we have had. I would like to thank Beverly Chamberlin in the Mass Spectrometry Facility for her help in running the 505 GC/MS. I would also like to thank Dr. Long Le and Dr. Kermit Johnson in the NMR Facility for their assistance in setting up numerous NMR experiments. In my years here I have been fortunate to have met some very great friends. I would especially like to thank Carla, Brian, and Casey Swan; Scott Herr; Kim and Greg Katch; and V Upal and Tracy Sengupta. Thank you so much for all of the fun times. Because of you I have many special memories of my years here. Most of all I wish to express my sincerest love, thanks, and appreciation to my family: my parents, Jean and Jim; and Jim, Brenda, Joe, Katurah, Daniel, Stephanie, and Jennifer; and my grandparents, J can and Del Hahn. My accomplishments are a result of your continual unconditional love, support, and encouragement. vi TABLE OF CONTENTS LIST OF TABLES ..................................................... ix LIST OF FIGURES ..................................................... x LIST OF ABBREVIATIONS ........................................... xiv CHAPTER I OVERVIEW OF GENERAL BACTERIAL ADAPT ATIVE MECHANISMS . . . . 1 Membrane Structure and Organization ................................. 3 General Theories on the Roles of Membranes in Adaption .................. 9 Molecular mechanisms of adaptation ................................. 15 Membrane alterations ........................................ 15 Protein stability and function .................................. 24 Osmotic balance ............................................ 28 Special issues in subzero temperature adaptation ........................ 3O Salt/metabolite concentrations and oxygen availability .............. 31 Prevention of freezing of intracellular water ...................... 32 Maintenance energy ......................................... 32 Carbon source and heat generation ............................. 34 The universality of adaptive responses to various environmental perturbations .................................................... 35 References ...................................................... 37 CHAPTER II DEVELOPMENT OF A THEORY ON THE SIGNIFICANCE OF FATTY ACID HETEROGENEITY ................................... 45 Abstract ........................................................ 46 Introduction ..................................................... 47 Materials and Methods ............................................. 54 Results and Discussion ............................................ 57 References ...................................................... 86 vii CHAPTER III THE INTERRELATION BETWEEN TEMPERATURE-INDUCED SHIFTS IN CARBON METABOLISM AND MEMBRANE LIPID CHEMISTRY .......... 89 Abstract ........................................................ 90 Introduction ..................................................... 9 1 Materials and Methods ............................................. 94 Results and Discussion ............................................ 95 References ..................................................... 106 CHAPTER IV THE ADAPTATION OF ARCTIC BACTERIA TO LOW PRE-FREEZING TEMPERATURES INVOLVES MEMBRANE ALTERATIONS BUT NOT INCREASED PRODUCTION OF OSMOLYTES .......................... 108 Abstract ....................................................... 109 Introduction .................................................... 109 Materials and Methods ............................................ 111 Results ........................................................ 1 17 Discussion ..................................................... 142 References ..................................................... 149 CHAPTER V CHARACTERIZATION OF A MECHANISM OF ADAPTATION OF A pH AND SOLVENT RESISTANT BACTERIUM Sarcina ventriculi .................. 150 Abstract ....................................................... 151 Introduction .................................................... 152 Materials and Methods ............................................ 156 Results and Discussion ........................................... 159 References ..................................................... 1 80 CHAPTER VI SUMMARY AND PERSPECTIVES ..................................... 182 References ..................................................... 188 APPENDIX A PREPARATION OF THE RADIOLABELED PHOTOAFFINITY PROBE FOR IDENTIFYING PUTATIVE CROSS-LIN KIN G ACTIVITY ................ 189 Preliminary Results .............................................. 193 References ..................................................... 195 viii LIST OF TABLES Table 3.1: Percent of fatty acid methyl ester derivatives in the total methanolysate of the cells grown at 4°C, 14°C, 24°C, and 30°C in order of elution. . . . . 96 Table 4.1: Percent of fatty acid methyl ester derivatives in the total methanolysate of the cells grown at 4°C and 24°C in order of elution. .............. 141 ix Figure 1.1: Figure 1.2: Figure 1.3: Figure 1.4: Figure 1.5: Figure 2.1: Figure 2.2: Figure 2.3: Figure 2.4: Figure 2.5: Figure 2.6: LIST OF FIGURES Representation of the cell envelopes of (A) Gram-positive bacteria and (B) Gram-negative bacteria. ....................................... 5 Structures of some common classes of membrane lipids. ............. 6 Lipid polymorphism in the liquid crystalline state. ................. 13 Structures of novel lipids from bacteria. ......................... 21 Simplified pictures of the motional spectrum of lipids ............... 23 Diagram depicting the transfer of the thermal energy spectrum from the environment via longitudinal accordian modes of membrane acyl chains to the cell interior. ............................................ 49 Gas chromatographs of fatty acid methyl esters obtained from Arctic strain 45-3 cultured at (A) 4°C and (B) 24°C. ......................... 58 Plot of probability distributions at 4°C, 14°C, 24°C, and 30°C for Arctic strain 45-3 ................................................ 60 Plots of probability distributions at several growth temperatures for (A) Listeria monocytogenes (B) Bacillus amyloliquefaciens and (C) Bacillus subtilis. ................................................... 62 Plot of probability distributions at 60°C, 70°C, and 80°C for Bacillus caldolyticus. ............................................... 65 Plots of the variances of the acyl chain distribution curves versus growth temperature (in degrees Kelvin) and the equations for the curves for (A) Arctic strain 45-3, (B) L. monocytogenes, (C) B. amyloliquefaciens, and (D) B. subtilis .............................................. 66 Figure 2.7: Figure 2.8: Figure 2.9: Figure 2.10: Figure 2.11: Figure 2.12: Figure 2.13: Figure 3.1: Figure 3.2: Figure 3.3: Figure 3.4: Figure 3.5: Figure 3.6: Gas chromatograms of fatty acid methyl esters obtained from Arctic strain 45-3 cultured in the presence of two solvents. (A) 24°C control, (B) 0.1 M cyclohexane, and (C) 0.1M 1,4-cyclohexadiene ................. 69 Computer generated pictures showing the packing interactions between a straight chain hydrocarbon and cyclohexane and an w-cyclohexyl acyl chain from two different views (A) and (B). ...................... 72 Structures of the chemical perturbants used studies with Bacillus subtilis PY79. .................................................... 75 Gas chromatogram of fatty acid methyl esters of Bacillus subtilis PY79 from a 37°C culture. ........................................ 76 The effects of cis- and trans-decalins on the plot of the probability distributions of Bacillus subtilis PY79. .......................... 77 Effects of three antibiotics on the plot of the probability distributions of Bacillus subtilis PY79 ........................................ 79 A three dimensional graph showing hypothetical growth rate curves for conditions of high and low temperatures and solvent (fluidizing type) concentrations ............................................. 8 1 Hypothetical energy profiles for two different metabolic pathways are shown, a higher energy requiring profile in (A) and a lower energy requiring profile in (B). ...................................... 92 1H NMR of TSB media before and after growth of Arctic strain 45-3. . . 98 TLC separation of l4C-glucose, l4C-valine, and l4C-leucine for one time point and its corresponding analog profile indicating the relative intensities of the three spots. .......................................... 100 The extent of utilization of 14C—glucose by Arctic strain 45-3 at 4°C and 24°C. ................................................... 101 Total protein profiles for Arctic strain 45-3 ...................... 103 Schematic representation of the relationship of the balance between two types of cellular metabolism, the potential thermal or enzymatic control of the metabolic processes, and the pools of available acyl-CoA primers for fatty acid biosynthesis. ...................................... 104 xi Figure 4.1: Figure 4.2: Figure 4.3: Figure 4.4: Figure 4.5: Figure 4.6: Figure 4.7: Figure 4.8: Figure 4.9: Figure 4.10: Figure 4.11: Figure 4.12: Figure 4.13: Figure 4.14: Figure 5.1: Siberian permafrost bacterial sampling region in the tundra zone of north- eastern Russia, Kolyma-Indigirka Lowland (152—162°E and 68-72°N). ................................................ 112 Electron micrographs (at the same magnification) of Arctic strain 45-3. .................................................... 1 18 Fluorescent micrographs using RIT C optics (at the same magnification) of Artic strain 45-3 stained with 0.1% acradine orange attached to silanized coverslips. ............................................... l 19 1H N MR spectra of the cytoplasmic metabolites. ................. 121 Two dimensional DQF-COSY spectrum of the cytoplasmic metabolites from cells cultured at 4°C. ................................... 123 1H N MR spectra of some common osmolytes. (A) glucose (B) glycerol (C) mannitol (D) sorbitol. ................................... 125 Gas chromatograph profiles of the capsular polysaccharides from the bacteria cultured at (A) 4°C and (B) 24°C. ...................... 127 1H N MR spectra of the total lipids extracted from the bacteria grown at (A) 4°C and (B) 24°C ....................................... 131 1H NMR spectra of the lipid headgroups from the bacteria grown at (A) 4°C and (B) 24°C. ......................................... 133 1H NMR spectrum of the methanol insoluble component from cells cultured at 4°C. ........................................... 136 Positive ion electrospray mass spectrum of the triacylglycerol from cells grown at 4°C. ............................................. 137 Gas chromatographs of fatty acid methyl esters obtained from bacteria cultured at (A) 4°C and (B) 24°C .............................. 139 SDS-PAGE gel of total proteins of lysates from cells cultured at 4°C and 24°C. ................................................... 143 Model showing the solvation of the region between two bilayer membranes. .............................................. 148 Conceptual model for the coupling of alkyl chains of the lipid components by an integral membrane catalytic system. ...................... 154 xii Figure 5.2: Figure 5.3: Figure 5.4: Figure 5.5: Figure 5.6: Figure 5.7: Figure 5.8: Figure 5.9: Figure A.1: Figure A2: The electron impact mass spectrum of one of the chimeric bifunctional fatty acid dimethyl esters formed by tail-to—tail coupling between an exogenously added fatty acid (heptadecanoic acid) and a hexadecanoic acid residue from an S. ventriculi lipid component. ............... 161 The electron impact mass spectrum of a 23 carbon fatty acid methyl ester resulting from coupling of n-heptane to a hexadecanoic acid residue from a S. ventriculi lipid component at the w-l position ................. 163 1H NMR spectra of (A) total fatty acid methyl esters from S. ventriculi cultured with hexadecylpyridinium chloride, and (B) isolated fatty acid methyl esters containing hexadecylpyridinim coupled to native S. ventriculi membrane lipids. .................................. 165 Bacterial (proposed) and synthetic mechanisms of the reduction and subsequent oxidation of the pyridine ring of hexadecylpyridinium chloride. ................................................. 168 Gas chromatogram of fatty acid methyl esters derivatized from total lipid extracts of S. ventriculi in culture at pH 9.7 ..................... 169 Gas chromatographs of fatty acid methyl esters from S. ventriculi cells cultured in (A) H20 and (B) D20. ............................. 172 Results from radical trap/scavenger experiments indicating the percent of tail-to-tail coupling of S. ventriculi membrane lipids normalized to the 50°C controls. ............................................ 175 1H NMR spectra showing the effect of cyclohexane and 1,4- cyclohexadiene on the lipid packing arrangement of dipalmitoyl phosphatidylcholine. ....................................... 177 Synthesis of the l4C-labeled photoaffinity probe. ................. 194 Phosphoimage of the nylon membrane containing proteins from several G- 75 column fractions ........................................ 194 xiii ACP DGDG DPG DPPC DQF—COSY DSC DTA ESR FAME GC GC-MS 1H NMR LAM MeOH MGDG PA PC PE LIST OF ABBREVIATIONS Acyl carrier protein Diglucosyldiglyceride Diphosphatidylglycerol (cardiolipin) DL—a-phosphatidylcholine, dipalmitoyl Double quantum filtered-correlated spectroscopy Differential scanning calorimetry Differential thermal analysis Electron spin resonance spectroscopy Fatty acid methyl esters Gas chromatography Gas chromatography-mass spectrometry Proton nuclear magnetic resonance spectroscopy Longitudinal accordian modes Methanol Monoglucosyldiglyceride Phosphatidic acid Phosphatidylcholine Phosphatidylethanolamine xiv PG Phosphatidyl glycerol PI Phosphatidylinositol PS Phophatidylserine RITC Rhodamine iso-thiocyanate SDS-PAGE Sodium dodecylsulfate-polyacrylamide gel electrophoresis TAG Triacylglycerol TLC Thin layer chromatography TOCSY Total correlated spectroscopy TSB Trypticase soy broth XV CHAPTER I OVERVIEW OF GENERAL BACTERIAL ADAPTATIVE MECHANISMS Bacteria, as a class of organisms, occupy almost all the environmental niches known on earth including many that are man-made, from extremes of temperature, pH, pressure, salinity, and solvent concentrations. A key to adaptation in the face of such vastly different circumstances is the maintenance of a functioning biochemical machinery and structural integrity. What are the most important physical properties bacteria must maintain to survive? Are there common adaptative mechanisms the various species have developed to survive more than one type of environmental condition? Or are there different mechanisms used even among species adapted to the same types of environments? One characteristic bacteria must maintain is membrane integrity. However, it is not enough just to keep the membrane intact to ensure survival to extreme environments or to variations in those niches. It must also function under the new conditions. As single-celled organisms, the membrane in bacteria is in constant contact with its environment, and as such may serve as the primary sensor to environmental fluctuations. The membrane may also serve to transmit the available thermal energy (determined by temperature) into the cell via certain vibrations of the acyl chains [1]. Membrane fluidity is another important consideration. It is commonly known that bacteria alter their lipid and fatty acid composition as a result of various environmental perturbations. One theory for the purpose of these alterations is to maintain the same degree of membrane fluidity, termed "homeoviscous adaptation" [2]. Others argue that maintaining a certain amount of the necessary bilayer type phase is more important, termed "homeophasic adaptation" [3]. Other characteristics are also important to cell survival such as maintaining protein/enzyme stability and function; and altering the balance of metabolic pathways to utilize available nutrients and generate needed energy. These factors also play a role in determining membrane structure and composition. This literature review will focus on theories of membrane functions in adaptation and adaptability of bacteria to extreme environments as well as common molecular mechanisms of adaptation used by various types of extremophiles. Since much of the work in this dissertation was done using a Siberian permafrost bacterium, a section of this review focuses on special issues in subzero temperature adaptation. Membrane Structure and Organization All cells require at least one membrane to contain the cytoplasm and all cellular functions. Procaryotes, the simplest of all organisms, contain membranes only in their cell envelope. The peripheral membrane(s) is not only a boundary for the cell, but also serves as a semipermeable barrier to the outer environment as well as a location for many cell processes to take place. While the composition of membranes is complex and varies greatly between organisms, the basic structure is still considered to be the fluid mosaic system described by Singer and Nicolson [4]. This model depicts membranes as containing integral and peripheral proteins embedded in a fluid-like matrix consisting of a bilayer of lipids. In addition to the cell envelope membrane(s), procaryotes usually also contain a rigid cell wall composed of the macromolecule peptidoglycan. In Gram-positive bacteria this cell wall lies outside the cytoplasmic membrane. In Gram-negative bacteria the cell wall lies between the inner (cytoplasmic) and outer membrane (Figure 1.1). Peptidoglycan is a polymer composed of glycan strands cross—linked by peptides. The glycan strands are usually made of alternating units of N-acetylglucosamine and N-acetylmuramic acid which are B- l ,4- linked. The short peptide chain is linked by its amino terminus to the carboxyl group of muramic acid and contains L and D amino acids. These peptide chains are cross-linked by interpeptide bridges enabling the formation of a rigid macromolecule that surrounds the cell. While there is a high degree of variability in the peptidoglycan structures among various procaryotes especially in the peptide structures, there are relatively few cases of phenotypic alterations when a bacterium is cultured under different growth conditions. When structural alterations do occur, it is usually when certain nutrients are limiting factors for growth [5]. The lipid composition of membranes, however, is highly variable not only among cell types but also due to numerous environmental perturbations. This diversity is due to variations in both the polar head groups and the fatty acyl chains. The most common membrane lipids are shown in Figure 1.2. The membranes of eucaryotes primarily consist of glycerol-based phospholipids, including phosphatidylcholine (PC), phosphatidylethanolamine (PE), phosphatidylinositol (PI), phosphatidylserine (PS), and cardiolipin. Another major class of eucaryotic membrane lipids are the sphingosine-based lipids, including sphingomyelin and glycosphingolipids. Glycolipids, both glycerol- and sphingosine-based, often function as cell-surface associated antigens and recognition factors in eucaryotes. Eucaryotic membranes, especially mammalian plasma membranes, contain oo‘ncfi-‘—-- / b/ 5 j E } cauwm Figure 1.1: Representation of the cell envelopes of (A) Gram-positive bacteria and (B) Gram-negative bacteria. (adapted from ref. 94) O I 0" ,0! l W0 CH,—O-:-—OH o 0 ‘0-Gs-Cth-NOM. ‘0—CNn-M“NH- -O—CH.—fH—NH. 00? OH —o-cn.-i:u—cu.ou -o‘ p 0 OH m. ’ OH W\ ,3? + o a4.—o—r--o-a1.-ai.-m o“ w-oa-as-m. O “\leY x?“ 0— We ‘9‘!“0‘7‘0 o O 0‘ 7H. 0’ o\m_°_z_o Ono’gag—o—E—Oxfix 0 Wm OH 3,- M1 ‘ag-O-P-Ox+/ W I N-\- O 0 phosphatidic acid phosphatidylcholine phosphatldylethanolamlne phoephatldylserlne phoopmudytgiyoeml phoephatldyflnooltol Figure 1.2: Structures of some common classes of membrane lipids. cholesterol as well. While the eubacteria, cyanobacteria, and the archaea are typically distinguishable by their lipid composition, a common feature among them (with few exceptions) is that they all lack sterols. Among the eubacterial class phosphatidylglycerol (PG), PE, and cardiolipin are the major membrane lipids. In addition, Gram-positive bacteria also contain mono- and diglycosyldiacylglycerols. Cyanobacteria (which are photosynthetic) typically contain PG and glycolipids. The archaea, which seem to lie phylogenetically between procaryotes and eucaryotes, contain a special class of lipids in which the usual ester- linked fatty acids of the above mentioned lipids are replaced with ether linked phytanyl alkyl chains. The most common types are analogues of PG, phosphatidylglycerol phosphate (PGP), and phosphatidylglycerol sulphate (PGS). Another degree of membrane lipid diversity lies in the fatty acyl and alkyl structures. These structures vary in length, unsaturation, and branching. Some have carbocyclic rings. The possible combinations between lipid headgroups and hydrocarbon tails easily allow for hundreds of different lipid molecules to make up the membrane lipid matrix. The lipid species mentioned above are just the most common lipids found in membranes. Many organisms make other unusual lipids, especially those organisms that live in extreme environments. Another feature of biological membranes that may serve an important function in adaptation and adaptability is the high degree of lateral organization. Experiments using several techniques show that membrane lipids (and proteins) partition into many phases or domains. These domains are either gel or liquid—crystalline in nature [6]. The driving forces for phase formation are not well understood, though some possibilities are the minimization of free energy and geometric packing constraints. Free energy minimization would be achieved by organizational features that minimize the void spaces, the interaction of acyl chains with the aqueous phase, and the gaucheztrans ratio of the carbon-carbon bonds in the acyl chains for the lipid molecules in a domain. Free energy minimization would also be achieved by maximizing the interchain interactions and by specific molecular associations that give rise to phases of different symmetry [7]. The molecular geometric packing of lipids may also be a driving force for phase separation. The size of the head group and degree of unsaturation of the acyl chains leads to three dimensional lipid structures that are cylindrical, conical, or inverted conical shaped [8]. In addition the 2-fatty acid chain of phospholipids initially extends perpendicularly to the glycerol backbone, but it bends back at the second carbon atom to become parallel to the first acyl chain. This nonequivalence of the acyl chains leads to tilting of some phospholipids [9]. As a result phospholipids with different tilts and/or different acyl chain lengths are more prone to separate into phases such that packing defects are plated out to the domain boundaries [7]. The enormous membrane lipid diversity and the organization of membrane lipids into different domains are two characteristics of membranes that enable procaryotes (and simple eucaryotes) to survive in constantly changing environments. As will be discussed further, bacteria alter their membrane lipid composition in response to numerous perturbations. The sensory mechanism of these perturbations, though, may be triggered by disruptions in the membrane phases. General Theories on the Roles of Membranes in Adaption Biological membranes exist in a phase transition between a gel state, in which the lipid hydrocarbon tails are primarily in an all-trans configuration, and a liquid—crystalline state, in which the hydrocarbon tails contain a number of gauche conformers. The motional dynamics (fluidity) of these lipids is critical to the numerous functions carried out by the membrane. A large proportion of membrane lipids needs to be in the liquid-crystalline phase to maintain transport and certain enzymatic functions. Studies with Escherichia coli and Acholeplasma laidlawii indicate that growth of these organisms is severely impaired when more than 50% of their membrane lipids are in the gel phase [10]. Perturbations that alter membrane fluidity often result in a rapid alteration of lipid acyl chains and/or the modification of the lipid head groups. Typically shorter and more unsaturated acyl chains are incorporated when perturbations cause the membrane to become more gel-like; while longer and more saturated acyl chains are made in response to fluidizing effects. Other modifications include acyl chain branching and cyclization. All of these modifications affect membrane fluidity. One theory for the alteration of membrane lipids is to maintain the same degree of fluidity regardless of growth temperature (and possibly due to other variations in the environment). This theory, termed homeoviscous adaptation, was first described for E. coli using electron spin resonance spectroscopy (ESR) [2]. ESR was also used to demonstrate 10 homeoviscous adaption in two other organisms, A. laidlawii [1 l] and Bacillus stearothermophilus [12]. All three organisms are able to alter their membrane lipid composition in response to temperature changes to maintain a virtually constant membrane fluidity. Other organisms also conserve membrane fluidity, though it does not always seem to be a complete compensation over the growth temperature range for particular organisms. Cossins and Sinensky proposed a coefficient of homeoviscous efficacy to quantify the compensatory response. The efficacy was calculated as the ratio of the change in membrane order to the difference in growth conditions (temperature, pH, etc.). An efficacy value of 1.0 indicates complete compensation of membrane fluidity, while values between 0 and 1.0 indicate partial fluidity compensation for the different conditions. The efficacy for procaryote membranes varies between 0.2 and 1.0, while the efficacy for all eucaryotes is less than 0.5 [13]. The large variation in efficacy values may be due in part to the techniques and/or probes used. Membrane fluidity is related to the properties of its composite lipids. The lipid bilayer, however, is not a typical liquid and as such does not behave like a bulk fluid. For typical liquids, like water, fluidity is defined as the reciprocal of viscosity. Membranes have different characteristics which also must be taken into consideration when trying to determine fluidity. For instance membranes are microscopic fluids with stringent boundary conditions and are essentially two-dimensional. Also, while typical fluids have distinct phases with regards to temperature (solid, liquid, or gas), membranes have numerous phases that can exist simultaneously (solid and liquid-crystalline). 11 Due to these special characteristics of bilayers, "membrane fluidity" is not very well defined and thus is hard to quantify. Typically membrane fluidity refers to the disorder of the acyl chains (structure) and/or their ease of movement (dynamics). Therefore, quantization of membrane fluidity encompasses several parameters such as diffusion (translational and rotational), order, packing, and permeability. These parameters are related, though how is not clearly understood [14]. As a result different techniques and different spectroscopic probes may measure different membrane parameters--yielding widely varied results in the measurement of membrane fluidity. Differential scanning calorimetry (DSC) and differential thermal analysis (DTA) are nonperturbing methods used to measure biomembrane phase transitions, but are not capable of direct measurement of fluidity within a given phase state. ESR and fluorescence spectroscopy are often used to directly measure lipid phase states and fluidity, but the probes used may partition into certain nonrepresentative domains as well as perturb the membranes. Nuclear magnetic resonance (NMR) techniques, such as Tl relaxation measurements, are capable of accurately monitoring phase transitions and providing orientations and rates of motion of the lipid molecules [10]. NMR techniques are likely to provide the most accurate measurements of membrane lipid fluidity, though currently they are not widely used. Despite the different techniques used, it is clear that most organisms do alter their membrane compositions to maintain a constant degree of membrane fluidity. Another theory has been proposed for membrane lipid composition alterations based on the highly variable homeoviscous efficacy coefficients. This theory, termed homeophasic l2 adaption [15], suggests that it is more important for organisms to maintain a particular ratio of bilayer to non-bilayer forming lipids. A large fraction of non-bilayer forming phases would decrease the stability of the bilayer. The three dimensional structure of lipid molecules depends on the size and shape of its headgroups and acyl chains. Lipids that have a cylindrical shape (like PC, PS, and PG) would favor the typical bilayer phase. PE containing unsaturated acyl chains, and cardiolipin or PA complexed with calcium form conical molecular structures, which favor inverted hexagonal or cubic phases (Figure 1.3). Lysophospholipids form inverted cones and would favor micelle type structures [8, 16]. Non-bilayer structures have been implicated in such physiological functions as adhesion, joining, and fusion [16, 17]. In support of this theory Wieslander and coworkers proposed that the regulation of lipid composition in A. laidlawii is governed by molecular shape. The dominate lipids in this organism are monoglucosyldiglyceride (MGDG) and diglucosyldiglyceride (DGDG), which have wedge and rodlike molecular shapes, respectively. The organism regulates the level of these two lipids to maintain a balance of lamellar and hexagonal phases [18-20]. While this theory is interesting, evidence supporting the formation of non-bilayer phases primarily comes from experiments with single lipids, or a mixture of only a few types of lipids. Membranes are extremely complex mixtures, and the fact that simple (few components) lipid mixtures assume non-bilayer phases in water does not mean they will form the same structures in a complex mixture. 13 Figure 1.3: Lipid polymorphism in the liquid crystalline state. (A) Lamellar phase (L,); (B) Hexagonal inverted phase (Hn); (C) Hexagonal phase (HI); and (D) Cubic phase. (adapted from ref. 95) l4 Membrane acyl chain heterogeneity also serves another function in the adaptability of organisms to their niches. As membranes completely surround cells, they are capable of transferring the available thermal energy (which are propagated by collisions) to the interior of cells. The heat or infrared quanta from the environment (of the magnitude "kT") is transferred by molecular vibrations. These quanta are passed via water in biological systems by collisions with other molecules. The acyl chains of membrane lipids serve as oscillators to focus these quanta to the cell interior. The quanta are transferred by specific acyl chain vibrations called longitudinal accordian modes (LAMS). By changing the lengths of the acyl chains, the membrane is tailored to propagate the available thermal energy spectrum. Evidence that membranes transform the available energy is shown in experiments in which the probability distributions of the main acyl chain length series are plotted from organisms at different growth temperatures and in the presence of D20. Probabilities of the energies per chain are plotted. The graphs indicate a Gaussian distribution of the acyl chain lengths and their energies, and there is a direct correlation between the variance (02) of the Gaussian distributions and kT. As temperature decreases, kT decreases as well. The distribution of available energies becomes more narrow as indicated by an increase in the height of the Gaussian curve. Therefore, low temperature effects (including D20 effects) indicate cells have a more limited set of available energies which can be utilized. This study indicates that organisms are able to "tune" their membranes to the value of kT by adjusting the acyl chain distributions of the membrane lipids [1]. 15 Molecular mechanisms of adaptation Procaryotes and the archaea, unlike most eucaryotes, have the ability to adapt to many environmental extremes. Procaryotes and the archaea colonize most of the harshest environments on earth, from extremes in temperature (hot and cold), pH (acidic and alkaline), salinity, osmotic pressure, and they survive in the presence of many detrimental chemical substances. To survive in all of these environmental conditions organisms must maintain certain functions, including membrane structural and functional integrity, osmotic balance, protein stability and function, and the ability to carry out numerous metabolic processes. In adapting to the environment (and changes within) these organisms have certain molecular mechanisms which they have altered (or can alter) to maintain critical cellular functions. These alterations include changes in membrane and other cell surface structures, osmolyte changes, protein stability and expression, and the ability to balance metabolic processes. Membrane alterations Bacteria can alter their membrane structure in many ways. One way is to alter the lipid composition, though it is not always affected by environmental changes--the most commonly studied of which is temperature. In one study psycrophiles, and not psycrotrophs, were found to alter their lipid composition in response to temperature [21]. In another study, four psychrophilic Vibrio species were found to have thermosensitive conversion of PS to PE or of PG to DPG, though the direction of the temperaure change that induced these conversions was not the same in the four organisms [22]. On the other hand, several 16 psychrotropic organisms like Pseudomonas spp. [23] and Micrococcus cryophilus [24] do not alter there phospholipid composition in response to temperature changes. It has been observed for several organisms, including Tetrahymena [25], Mycobacterium smegmatis [26], and several temperature-sensitive mutants of E. coli [27, 28], that the ratio of cardiolipin to phosphatidylglycerol increases with temperature. The increase in cardiolipin effectively cross-links the lipid headgroups to counteract increased membrane fluidity with the increase in temperature. Similarly, Sarcina ventriculi also cross-links its lipid headgroups in response to numerous perturbants that increase membrane fluidity [29]. One strain of Bacillus stearothermophilus increases its phosphoglycolipid content when the growth temperature is raised from 45 to 71°C [30] This result is in agreement with the trend towards a general increase in the amount of glycolipids with increasing temperature observed in organisms ranging from psychrophiles to moderate and extreme thermophiles [3]. In Tetrahymena the alterations in lipid headgroups is much slower, and is suggested to be a secondary adaptive mechanism to fine tune membrane dynamics after altering acyl chain composition [31]. While bacteria may or may not alter their lipid headgroup composition, most eubacteria and cyanobacteria alter their fatty acid composition in response to perturbations in their environments. Eubacteria can alter the membrane fatty acid chain length, unsaturation, branching, and cyclization. Cyanobacteria can only alter the chain length and unsaturation of their acyl chains [3]. Typically changes to the membrane acyl chains occur rapidly in response to perturbations. Shorter, more unsaturated or branched acyl chains occur 17 when bacteria are exposed to cold, high pressures, and substances that reduce the membrane fluidity. Fatty acid cyclization may also decrease on lowering growth temperatures in Gram- negative bacteria. This is seen in such organisms as E. coli [32, 33], P. fluorescens [34, 35], P. halosaccharolytica [36, 37], Proteus mirabilis [38], and Serratia marcescens [39]. In response to temperature changes the amount of branched chain fatty acids and/or the relative proportion of anteiso versus iso-branched fatty acids is altered in Gram-positive bacteria. Low temperatures result in an increase in the lower melting anteiso-fatty acids and decreasing amounts of higher melting, odd-carbon, iso-fatty acids [40]. The reduction of cyclopropyl fatty acids seen in Gram-negative bacteria as temperatures drops is probably due to the synthetic mechanism for forming these fatty acids. The formation of these fatty acid involves methylation by S-adenosylmethionine (SAM). The synthesis of SAM requires a lot of energy (ATP). As will be discussed in more detail, the available thermal energy which drives cellular processes is much less at low temperatures causing the formation of SAM to be thermodynamically unfavorable, such that less SAM would be available at lower temperatures. The increase in branched fatty acids in Gram-positive organisms also is most likely due to thermodynamic alterations in the balance of cellular metabolic pathways. A shift in metabolism at low temperatures to favor amino acid metabolism over carbohydrate metabolism would increase the branched acyl-CoA primers available for fatty acid synthesis causing an increase in the amount of branched lipid acyl chains. In aquatic environments organisms must contend with hydrostatic pressure. On average hydrostatic pressures below the ocean continental shelf are 200 atmospheres and 18 reach as high as 1100 atmospheres at the deepest part of the ocean. In addition temperatures below 1000 m are usually less than 4°C. Hydrostatic pressure compresses lipid bilayer causing an ordering effect on the acyl chains increasing the gel state of membranes. Hence pressure acts similarly to low temperatures in causing a decrease in membrane fluidity and as a result similar alterations in membrane acyl chains are observed [41]. The barophilic marine bacterium CNPT3, most likely in the Vibrio genus, increases it's proportion of unsaturated acyl chains particulary Cm and to a lesser extent C ,8, I at the expense of C 14,0, Cm, and Cl 4. 1- While the amount of Cm decreases with increasing pressure this bacterium still contains a larger percent of CW than most marine vibrios [42]. In addition to temperature and pressure, the chemical composition of the aqueous environment surrounding cells influences cellular structures and functions. In regards to cell membranes, a major portion of the lipids are anionic resulting in a negative surface potential which is stabilized by cations in the aqueous phase. As a result the salinity and pH of the surrounding aqueous phase effects the membrane surface potential and ionization of the anionic phospholipids [41]. It has been observed that an increase in salinity results in similar acyl chain changes cause by low temperatures and increased pressures. Namely in Gram- negative bacteria cyclopropyl and unsaturated fatty acids increase while in Gram-positive bacteria an increase in branched chain fatty acids is observed. The increase in cyclopropyl fatty acids of Gram-negative bacteria must be interpreted with caution since in some strains the major change is from unsaturated to cyclopropyl acyl chains, which would not alter the fluidity greatly [43]. In the halotolerant (meaning no specific salt requirements for growth) 19 bacterium Staphylococcus epidermidis a large increase in anteiso-15:0 fatty acids occurs in response to salt concentrations of 25%, which is again a similar response to low temperature [44]. The increase in salinity would affect the polar headgroups of lipids, and cellular membranes in general have an overall negative charge. In fact, an increase in the anionic membrane lipid content relative to zwitterionic lipid has been observed with increased salinity. In Gram-negative bacteria the major change is an increase in PG over PE, and sometimes an increase in DPG as well. In Gram-positive bacteria, which usually have a more complex lipid composition, an increase in phosphoglycolipids, PG, and/or DPG is seen with increased salinity [43]. These changes in headgroups, however, are similar to the fluidizing effects caused by increased temperatures. Some aspects of these trends in acyl chain versus headgroup alterations are confusing and may arise from differences between the systems studied and the ranges of salt concentrations used in the studies. For instance the cations of the salts at low to medium concentrations may stabilize the negative charges of the lipid headgroups through electrostatic interactions causing an ordering effect on the acyl chains and rigidifying the membrane. As a result more unsaturated, cyclopropyl, and/or branched acyl chains would offset this rigidifying effect. However, at high salt concentrations the abundance of ions would have a chaotropic effect on the membrane. To counteract this fluidizing effect increases in DPG (cardiolipin) would cross-link the lipid headgroups and stabilize the membrane. Longer, saturated and straight acyl chains occur in membranes when bacteria are exposed to high temperatures, extremes of pH, and substances that increase membrane 20 fluidity. In addition to these general acyl chain structural changes, many bacteria have special types of membrane lipids that allow these organisms to thrive in harsh environments or adapt to habitats that undergo many extremes. These include, in particular, thermophilic bacteria or bacteria that can adapt to high temperatures. Some of these specialized lipids are shown in Figure 1.4. These lipids include the tetraether and diether lipids of the archaea. The alkyl ether linkage is resistant to cleavage over the wide pH range encountered by these organisms. All three classes of the archaea (extreme halophiles, methanogens, and thermoacidophiles) have membranes composed of lipids derived from one basic core structure, diphytanylglycerol ether. In addition, some archaea lipids contain one to four diphytanylglycerol ether cyclopentane rings [45]. The membranes of these organisms are very tolerant to extreme conditions. Archaea inhabit climates with temperatures above 100°C as well as niches with 36% salt content, pHs above 11 and less than 1, and pressures as high as 1100 atmospheres in the deepest part of the ocean. [46]. The extreme halophilic archaea contain only diether derivatives. This may, however, be a reflection of their typical growth temperature rather than their tolerance for salt, meaning they may not be adapted for extremely high growth temperatures which might require tetraether lipids. But as temperature increases the percent of tetraether lipids increases to the point where the therrnoacidophiles contain only tetraether derivatives. Similarly, Sarcina ventriculi forms onto-fatty acids and cross-links its headgroups forming a bipolar monolayer, which is reminiscent of the tetraether lipids in archaea. S. ventriculi synthesizes these lipids in the presence of high temperatures, acidic and alkaline pHs, and in the presence of certain fluidizing solvents [29, 47]. The anaerobic thermophilic bacterium T hermoanaerobacter 21 C3308 A : o H 33 ' o Err . r H B EY-OH c.a.c0— n. wen-carcmucn. Cm H ‘OCWCHDEE‘EETCHflu—COO-TH n H ‘03 HgC(CHq)n—CH-CH H: H [Eco—9 ' c OH OH OH OH Figure 1.4: Structures of novel lipids from bacteria. (A) diether lipid and (B) tetraether lipid from Archaea; (C) a diabolic acid-containing phospholipid from Butyrivibrio S2; and (D) a tetrahydroxybacteriohopane lipid from Acetobacter. 22 ethanolicus 39E [48] and the hypertherrnophile Thermatoga maritima [49] also synthesizes a,w-dicarboxylic fatty acids in response to high temperatures. Other specialized lipids include hopanes, and co-cyclohexyl fatty acids. Hopanes are the bacterial equivalent of cholesterol [50]. These structures have been found in diverse bacteria including Acetobacter [51], cyanobacteria, obligate methylotrophs, and in many Gram-negative and Gram—positive chemoheterotrophs [52]. w-Cyclohexyl fatty acids are the predominant fatty acids in the thermoacidophiles isolated from hot springs all over the world. Most have been initially classified as strains of Bacillus acidocaldarius. But (o-cyclohexyl fatty acids are also major components in some mesophilic bacteria such as Curtobacterium pusillum [53, 54]. Several strains of therrnoacidophilic bacteria isolated from Japanese hot springs included anywhere from 74 to 93% of these fatty acids, and biosynthesis of these lipids increases with increasing glucose in the medium [53]. This is necessary because acids and solvents are made from glucose as a primary fermentation activity. To protect its membrane, the levels of these (o—cyclohexyl fatty acids must increase as glucose concentration increases and fermentation activity increases as a consequence. The region of greatest hydrocarbon chain mobility in a biomembrane is the area between the two leaflets (Figure 1.5) [55]. By incorportating an (o-cyclohexyl group in this zone, the organism can reduce the extent of chain terminus rotation and wagging. This affords a very high degree of motional restriction and hence control of dynamics. The flat (all trans-fused) hopane molecule does the same, but along the entire length of the hydrocarbon chain. 23 Figure 1.5 : Simplified pictures of the motional spectrum of lipids. (Pm indicates the maximum angle subtended by the two acyl chains at the headgroup. It is a measure of the extent to which the acyl chains are sprayed. (A) Monolayer form, and (B) Bilayer form. 24 Protein stability and function Cells must also maintain protein stability and function. The studies to date primarily focus on protein stability at high temperatures, mainly from thermophilic organisms. While the exact mechanisms that confer high temperature tolerance are not known, some suggestions are fewer amino acids that are vulnerable to heat and the formation of "salt- bridges," which consists of ionic bonds through the interior preventing denaturation. Enzymes from thermophilic barophiles, isolated from deep sea vents, often consist of a more dense hydrophobic interior as well as containing salt bridges. Increased pressure stabilizes van der Waals interactions and is thought to destabilize salt bridges. In general enzymes isolated from bacteria adapted to high temperatures are also extremely resistant to high pressure [56]. Proteins from extreme halophilic and even some moderate halophiles have been shown to contain a high content of acidic amino acids and are also relatively deficient in nonpolar amino acids compared to nonhalophiles [43]. The tendency to increase the amount of acidic arrrino acids relative to basic amino acids may not be general. Increased proportions of polar amino acids will, in general, help in keeping the proteins hydrated in this highly polar environment. Increases in N aCl and KCl increases the extent of interactions of hydrophobic groups again by a "salting out effect." Proteins with less nonpolar amino acids will maintain a more flexible conformation and a higher degree of hydration with increased salinities. In addition to protein stability, certain enzyme activities and metabolic processes must be maintained. Since membrane alterations, especially changes in acyl chain structures, 25 occur in response to environmental perturbations, some aspect(s) of lipid biosynthesis must be regulated by the effects of the environment on the cell. Temperature regulated changes in membrane fatty acid composition have been studied the most. Two mechanisms for making unsaturated lipids are known. These pathways are 'aerobic' or 'anaerobic' and can usually be distinguished by the position of the double bond in the acyl chain. The aerobic pathway requires molecular oxygen and results in palmitoleate (16:1A9cis) and oleate (18: 1A9cis) as the primary unsaturated products. The anaerobic pathway yields palmitoleate and cis-vaccenate (18:1A11cis) as the major unsaturated products [57]. All anaerobic and facultative anaerobic bacteria use the anaerobic pathway of fatty acid biosynthesis. In E. coli, which uses the anaerobic pathway, the ratio of saturated to unsaturated acyl chains is thought to be determined by the thermal regulation of B-ketoacyl-ACP synthase. This enzyme exists in two forms in E. coli, type I and H, and catalyzes the two carbon elongation reaction of fatty acid biosynthesis. Type H is more efficient in elongating palmitoleate than palmitate (16:0) and is more temperature labile. Thus at low temperatures palmitoleate is elongated to cis-vaccenate and less palmitate is made. In this case the effect of longer chains is outweighed by the effect of increased unsaturation. In this organism the thermal regulation of fatty acid composition does not require de novo protein synthesis, only changes in activity of existing enzymes occurs [58]. The aerobic pathway for producing unsaturated acyl chains involves desaturation of saturated fatty acids. In this case the fatty acid synthetase yields saturated fatty acids only. The desaturases, which are membrane-bound, are believed to work on membrane 26 phospholipids directly. Temperature then could regulate these enzymes by inducing membrane phase changes to alter the enzymatic activity, or by induction of new protein synthesis [58]. In Bacillus strains the desaturase enzyme is temperature labile resulting in faster turnover rates with increased temperatures. Drops in temperature then cause an increase in the amount of desaturase present in the membrane (slower turnover rate). Lower temperatures also are known to induce desaturase activity in these organisms [59]. As noted earlier, changes in membrane acyl chain length occur with alteration of temperature. It is proposed that an elongase enzyme that is thermally regulated is responsible for this. Hence, Micrococcus cryophilus alters the ratio of oleic and palmitoleic acids, the only fatty acids in its membrane, in response to temperature supposedly through an elongase enzyme that interconverts Cl6 and C18 fatty acids by way of a C14 intermediate [58]. Russell and coworkers suspect an essential component of this process such as acyl carrier protein (ACP) turns over rapidly because inhibitors of protein synthesis only partially inhibit this interconversion [58]. An alternative idea is that the acyl chain length selection process is random, and that the most viable cells at a given temperature have a particular acyl chain distribution. Dramatic shifts in temperature select for the bacterial populations with the best suited acyl chain length distributions or the ones that can alter their distributions to function at the new temperature. Other alterations of fatty acid structures are seen in response to temperature including changes in the cyclization, branching, and cross-linking of fatty acids, though the 27 mechanisms of thermal regulation of these syntheses has not yet been shown. The synthesis of iso- and anteiso-branched fatty acids requires de novo fatty acid synthesis using primers derived from amino acids. Temperature regulation of the transaminases that convert amino acids to their a-keto acids has not been demonstrated as of 1984 [58], and a subsequent literature search to the present has found no further evidence that this has been demonstrated. As will be discussed later, the thermal regulation of the synthesis of the branched fatty acids relates to the available pools of acyl-CoA primers as well as to the available thermal energy. It has been shown that the tail-to—tail cross-linking of S. ventriculi membrane acyl chains is temperature dependent as well. The regulation occurs presumably through activation of a constitutively expressed enzyme, since inhibitors of protein synthesis do not affect this activity [29]. Similarly, the Archaeon Methanococcusjannashcii has been shown to increase the proportions of tetraether lipids (C40) over the level of diether lipids (C20) with increasing temperature [60], indicating some aspect of the formation of tetraethers is also thermally regulated. In addition to regulating enzymes involved in controlling fatty acid composition many organisms produce 'heat shock' proteins in response to detrimental stress factors. Though they are called heat shock proteins, they are synthesized in response to extreme changes in temperature (hot or cold), acid, ethanol, and other physical and chemical perturbations. These proteins are constitutively expressed at very low levels, but with a sudden large temperature increase (or extreme changes due to other perturbants) the levels of these proteins rapidly increase. As the cells acclimate to the new environmental condition, the heat 28 shock protein levels decrease slowly, usually over a couple of hours, back to basal levels. If cells, however, are slowly acclimated to a higher temperature (or other condition), the levels of heat shock proteins do not increase [61]. The major heat shock proteins (Hsp) are believed to function as molecular chaperones to aid in protein folding and prevent protein aggregation of unfolded proteins [62]. Some of these same proteins induced by heat are also induced by other stresses such as ethanol and inhibitors of oxidative phosphorylation [63]. E. coli, Bacillus cereus WSBC 10201, and Pseudomonasfragi all produce a new series of proteins in response to cold shock, one in particular is CspA [64-66]. Some of these cold shock proteins are thought to function in control of transcription and translation, namely to prevent initiation of protein synthesis [67]. E. coli and Salmonella typhimurium are two examples of bacteria that produce acid shock proteins [68, 69]. While the exact functions of most shock proteins are not known, they may simply reflect the shifts in metabolic pathways that must occur in response to changing environmental conditions. Osmotic balance Changes in concentrations of solutes in the aqueous environment surrounding organisms determines the flow of water into and out of the cell. Cell membranes are readily permeable to water but not to many solutes. As a result, these osmotic changes alter cell volume and/or pressure, and changes in cell volume lead to fluctuations in intracellular solute concentrations as well. While the cellular signals for sensing osmotic changes are not really known at this time [70], several physicochemical changes that might be used by Gram- negative bacteria have been suggested. These theories include baroreceptors for detecting 29 hydrostatic pressure changes in the interior bulk liquid phase of the cell, which have been viewed as unlikely [71]; turgor pressure sensors in the inner membrane which respond to compression against the peptidoglycan, and wall stretch receptors which respond to tangential stress within the peptidoglycan layer [72]; stretch receptors for detecting changes in membrane surface area [73]; and chemoreceptors for detecting internal and external solute concentration or water activity [74]. It is thought that a common adaptive response for many organisms is to alter intracellular concentrations of solutes that do not produce adverse effects on cellular macromolecules [71, 75, 76]. These solutes are referred to as osmolytes or compatible solutes. Compatible solutes include K+ ions; amino acids such as glutamate, glutamine, proline, and glycine betaine (an N-methyl substituted amino acid); trehalose and mannosucrose, which are disaccharides; and others such as monosaccharides and glycerol. Osmolytes in high concentrations are seemingly not too toxic, but they do affect some biochemical processes. It has been proposed that increased intracellular concentrations of osmolytes has overcome temperature—sensitive mutations [76] as well as increased the upper temperature growth limit [77, 78], and conferred cryotolerance [79] on many bacteria. Results from these studies should be taken with caution, though, because the media were altered and analysis on the membrane chemistry and the levels of protein expression were not canied out. Bacteria that are thought to produce compatible solutes include Escherichia coli [80-82], several marine methanogens [83], Rhizobium meliloti [84], and many others [70]. It should be noted, however, that an increase in compatible solutes may not indicate 30 osmotic adaptation. The increase (or decrease) of certain metabolites may be a consequence of the new conditions to which the cell has adapted. Alterations in metabolic pathways are to be expected as environmental conditions change. So an increase in metabolites may not be the cause of adaptation to changes in osmotic pressure. For instance, if the increase in a particular osmolyte is still low compared to the external osmolality there is little significance in this event with regards to adapting to changes in osmotic pressure. For example, the glutamine pool in E. coli remains proportional to the external osmolality, but is 10% less than the intracellular K+ ion concentration [80]. In halophilic archaea the glycine betaine concentration is about 100 fold less than the major compatible solutes (K+ and Cl’ ions) [85]. So in these instances the presence of the osmolytes may serve a different function and/or be a consequence of shifts in metabolism [70]. Again, it is thought that bacteria respond to the reverse situation, hypoosmotic shock (water influx), by decreasing osmolyte concentrations via excretion, dilution by growth, catabolism, and polymerization. As will be discussed in more detail in chapter 4, osmolytes are also thought to aide in cryoprotection of microorganisms, but this is not the only means by which bacteria can survive subzero temperatures. In the next section special issues relevant to survival in subzero climates is reviewed. Special issues in subzero temperature adaptation A great deal of research has focused on the adaptation of organisms to extreme heat. One reason for this is the potential industrial use of enzymes and organisms that can withstand such high temperatures. However from the aspect of adaptation of organisms to 31 extreme environments, the majority of the earth's biosphere is subjected to subzero temperatures permanently or seasonally. Water can be supercooled to the extent that it can still be a liquid down to -40°C and below. As such, organisms deal with stresses from supercooled water as well from freezing. In supercooled water the degree of ionization is greatly decreased, affecting its ability to serve as conjugate bases or acids in reactions. Also the viscosity of the water increases and thus diffusion rates decrease [86]. In light of these and other changes that occur as water is chilled or frozen, the organisms that inhabit these climates have special circumstances which must be dealt with in order acclimate to subzero temperatures. These issues include responding to changes in salt/metabolite concentrations and oxygen availability; prevention of freezing of intracellular water; and provision of cellular maintenance energy, carbon sources and heat generation. Salt/metabolite concentrations and oxygen availability As water freezes it generally separates out as pure water (ice) causing an increase in salt and metabolite concentrations in the remaining liquid phase. Bacteria in the liquid phase would need to adjust to these increases. However, bacteria that may be trapped within the ice see the other extreme: a sharp decrease in available salts both organic and inorganic. Diffusion of these salts and metabolites for use by the cells is now extremely limited. Oxygen on this planet is generated by photosynthesis, which obviously requires light. In most subzero climates, oxygen is primarily generated by algae but only on the surface of snow. For microorganisms buried underneath these frozen habitats, oxygen becomes 32 limiting very quickly because it cannot diffuse well through the ice. The bacteria must then develop efficient anaerobic mechanisms for generating energy using other inorganic terminal electron acceptors that are available. These electron acceptors may be nitrate, carbon dioxide, sulphate, thiosulphate, or sulfite [87]. Prevention of freezing of intracellular water Freezing of intracellular water causes several problems, the major one is cell rupture due to the expansion of frozen water. Another problem is an imbalance in electrolyte concentration. As ice forms it forces out the solutes creating areas with high electrolyte concentrations causing osmotic changes that result in the efflux of water. It is also possible for freezing to result in eutectic mixtures in which some of the solutes remain within the ice, preventing their availability for use by the cell. Cells must also contend with decreased diffusion of metabolites as water becomes supercooled and/or freezes. Maintenance energy All cells expend a certain amount of energy to maintain organization at the molecular level. This includes the spatial orientation of membranes, proteins, nucleic acids, etc. As is well known, there exists a natural tendency towards randomness and disorder (increase in entropy) in the universe. Therefore maintaining the necessary levels of organization as well as carrying out the biosynthesis of macromolecules, requires energy. More than two decades ago calculations were made by several researchers to estimate 33 the amount of ATP needed for maintenance energy by bacteria. John Pirt determined a maintenance coefficient for aerobic and anaerobic continuous cultures of Aerobacter cloacae [88]. This is an index that tells the amount of glucose needed to maintain a gram of cells. This coefficient was defined as the slope of the line of plots of the reciprocal of the yield value versus the reciprocal of growth rate. The values for the aerobic and anaerobic cultures were 0.09 and 0.47 grams of glucose per gram dry weight bacterium, respectively. The higher coefficient for anaerobic cultures is expected due to less efficient energy yields from anaerobic metabolism. Using the coefficient for the anaerobic culture, the amount of ATP needed to maintain one gram dry weight bacterium was one-twentieth (~0.75 mmole) the ATP necessary to produce one gram dry weight of this bacterium. It should be noted, though, that this value far exceeds the free ATP pool (~10 pmoles per gram dry weight bacterium) usually available in a given microorganism [89]. About ten years later Tempest and colleagues [90], by modifying Pirt's derivation, determined another linear relationship. The slope of their plot of energy source/cells/hour versus growth rate is also defined as a maintenance coefficient, but the error in this slope is primarily in the intercept. In the Pirt plots, the error is in the slope. While these two derivations for determining maintenance energy are often used, other calculations for determining maintenance yield have been done [91]. While there is no one way to calculate maintenance energy, the concept is still very important. The idea of maintenance energy was first suggested when it was observed that low levels of energy sources were not sufficient for subculturing bacteria, indicating that 34 cells may need some energy just to maintain viability [92]. At low temperatures cells require much more energy to produce one gram of cells (also partly due to decreased oxygen availability), but the total amount of energy put into the system (as measured by carbon source used) is even greater than that calculated to produce one gram of cells [89]. So at low temperatures, the amount of energy needed to maintain viable cells is also much greater. This can be explained by the driving force of cellular chemical processes--heat (thermal energy). Since cell maintenance requires turnover of macromolecules, to drive the chemical processes involved in turnover requires a certain amount of heat, which makes these processes more difficult at low temperatures. Heat expenditure is also necessary to sustain membrane dynamics at low temperatures. Low temperatures cause lipids to solidify, preventing the diffusion and lateral motion of substrates and enzymes used in the many chemical reactions that take place within the membrane. In addition maintaining the electromotive force across the membrane is entropically unfavorable. So a certain amount of energy is used to preserve membrane asymmetry. Enzymes, the biological macromolecular catalysts, also require thermal energy to sustain the optimal vibrational modes at which they function. Thus overall, with the decrease in available thermal energy at colder temperatures, cells must burn vast amounts of chemical fuels just to maintain an equivalent cell mass, much less allow for cell growth. Carbon source and heat generation This increased need for heat requires cells to be more efficient in generating heat and in using energy resources. Typically cells use three classes of substances for deriving 35 chemical energy: carbohydrates, amino acids, and fatty acids. Carbohydrates are the most oxidized of these compounds and hence generate the least amount of heat per gram than amino acids and fatty acids. A substantial loss in the net generation of energy from burning carbohydrates comes from the priming of them for metabolic degradation. This process requires a large investment of energy through the generation of phosphorylated species. Amino acids are an oxidation state lower than carbohydrates and fatty acids are the among the most reduced substrates for microorganisms. Both would provide more energy than carbohydrates and would be used with greater efficiency at subzero temperatures, as is indicated by their respective heats of combustion. The heat of combustion for glucose is 669.94 kg-cal/g. While for leucine (855.6 kg-cal/g) and two fatty acids, palmitate (2,384.76 kg-cal/g) and oleate (2,657.4 kg-cal/g), the heats of combustion are higher, especially for the fatty acids [93]. Further evidence for the preferred use of fatty acids as energy at subzero temperatures is the observation of lipid droplets in many Arctic and other psycrophilic organisms (Danielle Prévost, Agriculture and Agri-Food, Canada; personal communication). These lipid droplets consist of triacylglycerides, which are the storage form of fatty acids. The universality of adaptive responses to various environmental perturbations From the evidence documented above one can see that numerous environmental perturbations cause the same types of molecular mechanisms of adaptation. Therefore these stresses may induce the same sets of sensory mechanisms in a given organism. The peripheral membrane of procaryotes is the ideal candidate for sensing all of these variations within an environment. In surrounding the cell it has constant contact with its surroundings, 36 it transmits the available thermal energy to the cell interior, and its organization and adaptability allow for rapid alterations in its structure to compensate for any physical and chemical perturbations it senses. The primary sensory mechanism could be simply perturbations in the membrane dynamics sensed by disturbances of the gel and liquid- crystalline phases. As discussed above both physical and chemical stresses (including temperature, pressure, pH, salinity, and chemical substances) all result in membrane alterations which are thought to maintain a certain degree of fluidity. Cold temperatures, increased pressure, increased salinity (depending on the type and concentration of cation), and certain solvents all rigidify the membrane by inducing closer packing of the lipid acyl chains and/or by stabilizing electrostatic lipid headgroup interactions. Thus all these perturbations result in membranes with greater unsaturated acyl chains and shorter chain lengths, which disrupts this induced tighter packing. Other factors that fluidize the membrane (high temperature, pH and salinity extremes, and some solvents) by disrupting acyl chain packing and/or destabilizing electrostatic interactions of the lipid headgroups, result in membranes with longer, more saturated acyl chains which increases the acyl chain interactions causing a reduction in the acyl chain motion. The membrane, by virtue of its component lipid acyl chains, is tuned to the available thermal energy and transmits it to the cell interior. This available energy is what determines which metabolic reactions can be carried out, and the cell's metabolic processes are adjusted accordingly. The products of the metabolic process, however, determine the available pool of acyl-CoA primers which can be used for fatty acid synthesis. So there is a circular 37 relationship between available thermal energy, metabolism, and membrane structure. A change in any one of these parameters by any type of perturbation affects the other two. Psychrophilic and psychrotropic bacteria offer a unique opportunity to establish the relationship between energy, metabolism, and fatty acid structures. As discussed above, the available thermal energy is much less as temperature drops. Thus the balance between carbohydrate and amino acid metabolism is affected, which is also reflected in the membrane lipid acyl chain structures. In this work, principles of adaptation and adaptability of organisms to extreme environments are developed using an Arctic permafrost bacterium. These principles are then applied to data regarding other bacteria to test their generality. REFERENCES l. Hollingsworth, R. I. (1995) Membrane lipid alkyl chain heterogeneity is a molecular transform of the available thermal energy spectrum. J. Phys. Chem, 99: 3406-3410. 2. Sinensky, M. (1974) Homeoviscous adaptation--a homeostatic process that regulates the viscosity of membrane lipids in Escherichia coli. Proc. Natl. Acad. Sci. USA, 71: 522-525. 3. Russell, N. J., and N. Fukunaga. (1990) A comparison of thermal adaptation of membrane lipids in psychrophilic and thermophilic bacteria. FEMS Microbiol. Revs, 75: 171-182. 4. Singer, S. J. and G. L. N icolson. (1972) The fluid mosaic model of the structure of cell membranes. Science, 175: 720-731. 5. Schleifer, K. H., and O. Kandler. (1972) Peptidoglycan types of bacterial cell walls and their taxonomic implications. Bacteriol. Rev., 36: 407-477. 6. Karnovsky, M. J ., A. M. Kleinfeld, R. L. Hoover, E. A. Dawidowicz, D. E. McIntyre, E. A. Salzman, and R. D. Klausner. (1982) Lipid domains in membranes. Ann. N. Y. Acad. 38 Sci., 401: 61-75. 7. Jain, M. K. (1983) Nonrandom lateral organization in bilayers and biomembranes. In Membrane Fluidity in Biology, ed. R. C. Aloia, 1: 1-37. 8. Cullis, P. R., and B. DeKruijff. ( 1979) Lipid polymorphism and the functional roles of lipids in biological membranes. Biochim. Biophys. Acta, 559: 399-420. 9. Lee, A. G. (1983) Lipid phase transitions and mixtures. In Membrane Fluidity in Biology, ed. R. C. Aloia, 2: 43-88. 10. McElhaney, R. N. (1984) The relationship between membrane lipid fluidity and phase state and the ability of bacteria and myc0plasmas to grow and survive at various temperatures. Biomembranes, 12: 249-278. 11. Huang, L., S. K. Lorch, G. G. Smith, and A. Haug. (1974) Control of membranes lipid fluidity in Acholeplasma laidlawii. FEBS Letters, 43: 1-5. 12. Esser, A. F., and K. A. Souza. (1974) Correlation between thermal death and membrane fluidity in Bacillus stearothermophilus. Proc. Natl. Acad. Sci. USA, 71: 4111-4115. 13. Cossins, A. R., and M. Sinensky. (1984) Adaptation of membranes to temperature, pressure, and exogenous lipids. In CRC Physiology of Membrane F luidity Vol. [1, ed. M. Shinitzky, CRC Press, Boca Raton, FL. pp. 1-20. 14. van der Meer, W. (1984) Physical aspects of membrane fluidity. In CRC Physiology of Membrane F luidity Vol. 1, ed. M. Shinitzky, CRC Press, Boca Raton, FL. pp. 53-71. 15. McElhaney, R. N. (1984) The structure and function of the Acholeplasma laidlawii plasma membrane. Biochim. Biophys. Acta, 779: 1-42. 16. Rilfors, L., G. Lindblom, A. Wieslander, and A. Christiansson. (1984) Lipid bilayer stability in biological membranes. Biomembranes, 12: 205-245. 17. Cullis, P. R., M. J. Hope, and C. P. S. Tilcock. (1986) Lipid polymorphism and the roles of lipids in membranes. Chemistry and Physics of Lipids, 40: 127- 144. 18. Wieslander, A., A. Christiansson, L. Rilfors, and G. Lindblom. (1980) Lipid bilayer stability in membranes. Regulation of lipid composition in Acholeplasma laidlawii as governed by molecular shape. Biochemistry, 19: 3650-3655. 19. Wieslander, A., L. Rilfors, L. B.-A. Johansson, and G. Lindblom. (1981) Reversed cubic phase with membrane glucolipids from Acholeplasma laidlawii. 'H, 2H, and diffusion nulear magnetic resonance measurements. Biochemistry, 20: 730-735. 39 20. Wieslander, A., L. Rilfors, and G. Lindblom. (1986) Metabolic changes of membrane lipid composition in Acholeplasma laidlawii by hydrocarbons, alcohols, and detergents: Arguments for effects on lipid packing. Biochemistry, 25: 7511-7517. 21. Herbert, R. A., and M. Bhakoo. (1979) Microbial growth at low temperatures. In Cold Tolerant Microbes in Spoilages and the Environment, eds. A. D. Russell and R. Fuller. Academic Press, London. pp. 1-16. 22. Bhakoo, M., and R. A. Herbert. (1979) The effects of temperature on the fatty acid and phopholipid composition of four obligately psychrophilic Vibrio spp. Arch. Microbiol., 121: 121-127. 23. Bhakoo, M, and R. A. Herbert. (1980) Fatty acid and phospholipid composition of five psychrotrophic Pseudomonas spp. grown at different temperatures. Arch. Microbiol., 126: 51-55. 24. Russell, N. J. (1974) The lipid composition of the psychrophilic bacterium Micrococcus cryophilus. J. Gen. Microbiol., 80: 217-225. 25. Yamauchi, T., K. Ohki, H. Maruyama, and Y. Nozawa. (1981) Thermal adaptation of Tetrahymena membranes with special reference to mitochondria. Role of cardiolipin in fluidity of mitochondrial membranes. Biochim. Biophys. Acta, 649: 385-392. 26. Taneja, R., and G. K. Khuller. (1981) Phospholipid metabolism in Mycobacterium smegmatis ATCC 607 grown at 37°C and 27°C. Arch. Microbiol., 129: 81-84. 27. Hawrot, E., and E. P. Kennedy. (1978) Phospholipid composition and membrane function in phosphatidylserine decarboxylase mutants of Escherichia coli. J. Biol. Chem, 253: 8213-8220. 28. Raetz, C. R. (1976) Phosphatidylserine synthetase mutants of Escherichia coli. Genetic mapping and membrane phospholipid composition. J. Biol. Chem, 251: 3242-3249. 29. Lee, J ., S. Jung, S. Lowe, J. G. Zeikus, and R. I. Hollingsworth. (1995) A dynamically regulated transformation of the bilayer membrane of bacteria to a cross-linked 2-dimensional sheet during adaptation to unfavorable environmental pressures. Submitted to J. Amer. Chem. Soc. 30. J urado, A. S., M. S. Costa, and V. M. C. Madeira. (1988) Therrnotropic behavior of the polar lipids of B. stearothermophilus. Abstr. FEMS Int. Symp. on The Microbiology of Extreme Environments and its Biotechnological Potential. Troia, Portugal. 31. Thompson, G. A., and Y. Nozawa. (1984) The regulation of membrane fluidity in Tetrahymena. Biomembranes, 12: 397-432. 40 32. Marr, A. G., and J. L. Ingraham. (1962) Effect of temperature on the composition of fatty acids in Escherichia coli. J. Bacteriol., 84: 1260. 33. Kito, M., S. Aibara, M. Kato, and T. Hata. (1972) Differences in fatty acid composition among phosphatidylethanolamine, phosphatidyl glycerol, and cardiolipin of Escherichia coli. Biochim. Biophys. Acta, 260: 475-478. 34. Cullen, J ., M. C. Phillips, and G. G. Shipley. ( 1971) The effects of temperature on the composition and physical properties of the lipids of Pseudomonasfluorescens. Biochem. J., 125: 733-742. 35. Gill, C. O. (1975) Effect of growth temperature in the lipids of Pseudomonas fluorescens. J. Gen. Microbiol., 89: 293-298. 36. Hyono, A., S. Kuriyama, H. Hara, 1. Yano, and M. Masui. (1980) ESR studies on the membrane properties of moderately halophilic bacterium. 1. Physical properties of lipid bilayers in whole cells. J. Biochem., 88: 1267-1274. 37. Hara, H., A. Hyono, S. Kuriyama, I. Yano, and M. Masui. (1980) ESR studies on the membrane properties of a moderately halophilic bacterium. II. Effect of extreme growth conditions on liposome properties. J. Biochem., 88: 1275-1282. 38. Rottem, S., O. Markowitz, and S. Razin. (1978) Thermal regulation of the fatty acid composition of lipopolysaccharides and phopholipids of Proteus mirabilis. Eur. J. Biochem. , 85: 445-450. 39. Bishop, D. G., and J. L. Still. (1963) Fatty acid metabolism in Serratia marcescens: the effect of temperature on fatty acid composition. J. Lipid Res., 4: 87. 40. Suutari, M., and S. Laakso. (1994) Microbial fatty acids and thermal adaptation. Crit. Rev. Microbiol., 20: 285-328. 41. Hazel, J. R., and E. E. Williams. (1990) The role of alterations in membrane lipid composition in enabling physiological adaptation of organisms to their physical environment. Prog. Lipid Res, 29: 167-227. 42. DeLong, E. F., and A. A. Yayanos. (1985) Adaptation of the membrane lipids of a deep- sea bacterium to changes in hydrostatic pressure. Science, 228: 1101-1103. 43. Russell, N. J. (1989) Adaptive modifications in membranes of halotolerant and halophilic microorganisms. J. Bioenerg. Biomem, 21: 93-1 13. 44. Komaratat, P., and M. Kates. (1975) The lipid composition of a halotolerant species of Staphylococcus epidermidis. Biochim. Biophys. Acta, 398: 464-484. 41 45. Kates, M. (1992) Archaebacterial lipids: structure, biosynthesis and function. Biochem. Soc. Symp., 58: 51-72. 46. Pool, R. (1990) Pushing the envelope of life. Science, 247: 158-160. 47. Jung, S., S. E. Lowe, R. I. Hollingsworth, and J. G. Zeikus. (1993) Sarcina ventriculi synthesizes very long chain dicarboxylic acids in response to different forms of environmental stress. J. Biol. Chem, 268: 2828-2835. 48. Jung, 8., J. G. Zeikus, and R. I. Hollingsworth. (1994) A new family of very long chain moo-dicarboxylic acids is a major structural fatty acyl component of the membrane lipids of Thermoanaerobacter ethanolicus 39E. J. Lipid Res., 35: 1057-1065. 49. Carballeira, N. M., M. Reyes, A. Sostre, H. Huang, M. F. J. M. Verhagen, and M. W. W. Adams. (1997) Unusual fatty acid compositions of the hyperthermophilic Archaeon Pyrococcus furiosus and the bacterium Thermotoga maritima. J. Bacteriol., 179: 2766- 2768. 50. Ourisson, G. (1987) Bigger and better hopanoids. Nature, 326: 126-127. 51. Forster, H., K. Biemann, W. G. Haigh, N. H. Tattrie, and J. R. Colvin. (1973) The structure of novel C35 pentacyclic terpenes from Acetobacter xylinum. Biochem. J., 135: 133-143. 52. Rohmer, M., P. Bouvier—Nave, and G. Ourisson. (1984) Distribution of hopanoid triterpenes in prokaryotes. J. Gen. Microbiol., 130: 1137-1150. 53. Oshima, M., and T. Ariga. (1975) w-Cyclohexyl fatty acids in acidophilic thermophilic bacteria. J. Biol. Chem, 250: 6963-6968. 54. Mantsch, H. H., C. Madec, R. N. A. H. Lewis, and R. N. McElhaney. (1989) An infrared spectroscopic study of the therrnotropic phase behavior of phosphatidylcholines containing w-cyclohexyl fatty acid chains. Biochim. Biophys. Acta, 980: 42-49. 55. Yeagle, P. (1992) The dynamics of membrane lipids. In The Structure of Biological Membranes, P. Yeagle, ed., CRC Press, Inc., Boca Raton, FL. pp.157-174. 56. Flam, F. ( 1994) The chemistry of life at the margins. Science, 265: 471-472. 57. Kawaguchi, A., and Y. Seyama. (1984) Thermal control of fatty acid synthetases in bacteria. Biomembranes, 12: 279-301. 58. Russell, N. J. (1984) Mechanisms of thermal adaptation in bacteria: blueprints for survival. TIBS, pp. 108-112. 42 59. Fulco, A. J ., and D. K. Fujii. (1980) In Membrane F luidity. Biophysical Techniques and Cellular Regulation, M. Kates and A. Kuksis, eds, Humana Press, pp. 77—98. 60. Sprott, G. D., M. Meloche, and J. C. Richards. (1991) Proportions of diether, macrocyclic diether, and tetraether lipids in Methanococcus jannaschii grown at different temperatures. J. Bacteriol., 173: 3907-3910. 61. Yamamori, T., and T. Yura. (1982) Genetic control of heat-shock protein synthesis and its bearing on growth and thermal resistance in Escherichia coli K-12. Proc. Natl. Acad. Sci. USA, 79: 860—864. 62. Craig, E. A., B. D. Gambill, and R. J. Nelson. (1993) Heat shock proteins: molecular chaperones of protein biogenesis. Microbiol. Revs, 57: 402-414. 63. Craig, E. A. (1985) The heat shock response. CRC Crit. Rev. Biochem., 18: 239-280. 64. Goldstein, J ., N. S. Pollitt, and M. Inouye. ( 1990) Major cold shock protein of Escherichia coli. Proc. Natl. Acad. Sci. USA, 87: 283-287. 65. Mayr, B., T. Kaplan, S. Lechner, and S. Scherer. (1996) Identification and purification of a family of dimeric major cold shock protein homologs from the psychrotrophic Bacillus cereus WSBC 10201. J. Bacteriol., 178: 2916-2925. 66. Hebraud, M., E. Dubois, P. Potier, and J. Labadie. (1994) Effect of growth temperatures on the protein levels in a psychrotrophic bacterium, Pseudomonas fragi. J. Bacteriol., 176: 4017 -4024. 67. Russell, N. J. (1990) Cold adaptation of microorganisms. Phil. Trans. R. Soc. Lond. B, 326: 595-611. 68. Goodson, M., and R. J. Rowbury. (1989) Habitation to normally lethal acidity by prior growth of Escherichia coli at a sublethal acid pH value. Lett. Appli. Microbiol., 8: 77-79. 69. Foster, J. W. ( 1991) Salmonella acid shock proteins are required for the adaptive acid tolerance response. J. Bacteriol., 173: 6896-6902. 70. Csonka, L. N., and A. D. Hanson. (1991) Prokaryotic osmoregulation: genetics and physiology. Annu. Rev. Microbiol., 45: 569-606. 71. Cram, W. J. ( 1976) Negative feedback regulation of transport in cells. The maintenance of turgor, volume, and nutrient supply. In Encyclopedia of Plant Physiology, New Series, Vol. 2A, ed. U. Luttge, M. G. Pitman. Berlin/New York: Springer. pp. 284-316. 72. Frey-Wyssling, A. (1957) Macromolecules in Cell Structure. Harvard University Press, 43 Cambridge, pp. 1 12. 73. Sant'Ambrogio, G. (1987) Nervous receptors of the tracheobronchial tree. Annu. Rev. Physiol, 49: 611-27. 74. Macnab, R. M. (1987) Motility and chemotaxis. In Escherichia coli and Salmonella typhimurium: Cellular and Molecular Biology. F. C. Neidhardt, J. L. Ingraham, K. B. Low, B. Magasanik, M. Schaechter, and H. E. Umbarger, eds., American Society of Microbiology, Washington, D. C., pp. 732-759. 75. Brown, A. D. (1990) Microbial Water Stress Physiology: Principles and Perspectives. Chichester: Wiley. 76. Csonka, L. N. (1989) Physiological and genetic responses of bacteria to osmotic stress. Microbiol. Rev., 53: 121—147. 77. Tesone, S., A. Hughes, and A. Hurst. (1981) Salt extends the upper temperature limit for growth of food-poisoning bacteria. Can. J. Microbiol., 27: 970-972. 78. Troller, J. A. (1986) Water relations of foodbome bacterial pathogens—-an updated review. J. Food Prot., 49: 656-670. 79. K0, R., L. Tombras Smith, and G. M. Smith. (1994) Glycine betaine confers enhanced osmotolerance and cryotolerance on Listeria monocytogenes. J. Bacteriol., 176: 426-431. 80. McLaggan, D., T. M. Logan, D. G. Lynn, and W. Epstein. (1990) Involvement of y- glutamyl peptides in osmoadaptation of Escherichia coli. J. Bacteriol., 172: 3631-3636. 81. Boos, W., U. Ehmann, H. Fork], W. Klein, M. Rimmele, and P. Postma. (1990) Trehalose transport and metabolism in Escherichia coli. J. Bacteriol., 172: 3450-3461. 82. Cayley, S., M. T. Record Jr., and B. A. Lewis. (1989) Accumulation of 3-(N- morpholino)-propanesulfonate by osmotically stressed Escherichia coli K-12. J. Bacteriol., 171: 3597-3602. 83. Robertson, D. E., M. F. Roberts, N. Belay, K. O. Stetter, and D. R. Boone. (1990) Occurance of B-glutamate, a novel osmolyte in marine methanogenic bacteria. Appl. Environ. Microbiol. , 56: 1504-1508. 84. Botsford, J. L., and T. A. Lewis. (1990) Osmoregulation in Rhizobium meliloti: production of glutamic acid in response to osmotic stress. Appl. Environ. Microbiol., 56: 488-494. 85. Nicolaus, B., V. Lanzotti, A. Trincone, M. De Rosa, and A. Gambacorta. (1989) 44 Glycine betaine and polar lipid composition in halophilic archaebacteria in response to growth in different salt concentrations. FEMS Microbiol. Lett., 59: 157-160. 86. Franks, F., S. F. Mathias, and R. H. M. Hatley. (1990) Water, temperature and life. Phil. Trans. R. Soc. Lond. B, 326: 517-533. 87. Rose, A. H. (1976) Chemical Microbiology: An Introduction to Microbial Physiology, 3rd Ed. Plenum Press, New York. pp. 181-243. 88. Pirt, S. J. (1965) The maintenance energy of bacteria in growing cultures. Proc. R. Soc. London Ser. B, 163: 224-231. 89. Rose, A. H. (1976) Chemical Microbiology: An Introduction to Microbial Physiology, 3rd Ed. Plenum Press, New York. pp. 245-325. 90. Neijssel, O. M., and D. W. Tempest. (1976) Bioenergetic aspects of aerobic growth of K lebsiella aerogenes NCTC 418 organisms growing in chemostat culture. Arch. Microbiol. , 107: 215-221. 91. Russell, J. B., and G. M. Cook. (1995) Energetics of bacterial growth: Balance of anabolic and catabolic reactions. Microbiol. Rev., 59: 48-62. 92. Buchanan, R. E., and E. I. Fulmer. (1928) Physiology and biochemistry of bacteria. The Williams & Wilkins Co., Baltimore. 93. Weast, R. C., ed. (1985) Heat of Combustion. In CRC Handbook of Chemistry and Physics, 66th ed. CRC Press, Boca Raton, FL. pp. D272-277. 94. Ratledge, C., and S. G. Wilkinson. (1988) Microbial Lipids, Vol. 1, Academic Press, New York, pp. 1-22. 95. Gruner, S. M. (1992) In The Structures of Biological Membranes, P. Yeagle, ed., CRC Press, Inc., Boca Raton, FL, pp. 211-250. CHAPTER H DEVELOPMENT OF A THEORY ON THE SIGNIFICANCE OF FATTY ACID HETEROGENEITY 45 46 ABSTRACT The theory that the chain length distribution of the fatty acyl groups of bacteria is a Gaussian function the mean and variance of which depends on temperature was tested by examining this distribution in a permafrost bacterium that is capable of growth from temperatures close to freezing up to 30°C. The variance of the curve was a linear function of absolute temperature over most of the range with a second order deviation towards the upper end. This supports the idea that the chain length selection is a random process in which the distribution reflects the external temperature, or more correctly momentum distribution of molecules in the extracellular matrix, and is characterized by a smaller variance at lower temperatures and a higher one at higher temperatures. Fatty acid composition data from the literature for three other Gram-positive organisms were also used to support the theory. Having established that there is a consistent relationship between temperature and acyl chain distribution, then one can predict the fatty acid composition of a given organism over a temperature range. By adding a variety of chemical perturbants to Arctic strain 45-3, Bacillus subtilis PY79, and Sarcina ventriculi J K, we also show that there is a temperature equivalent for a given solvent perturbation in a particular organism. This is consistent with earlier isotope studies that indicated the mobility of the hydrocarbon chains is actually the critical dynamic factor that is influenced by temperature. The membrane acyl chain length alterations induced by chemical substances, such as solvents and antibiotics could, therefore, be explained by the structures of the perturbants and how the structures could influence the acyl chain packing. These results indicate that by knowing whether the 47 perturbant would fluidize or rigidify the membrane, the temperature tolerance of an organism can be increased (or decreased) by adding exogenous substances such as very long or short chain or fused-ring hydrocarbons to counteract the temperature effect. For extreme perturbations, the mean of the distribution curve shifted abruptly to form a new series of distribution curves about the new mean. For elevated temperatures, this is consistent with the increased value of kT. INTRODUCTION "Why not more quantitative data? Lipid data on the same strain, grown and analyzed under the same conditions in two different laboratories, or even in the same laboratory, may vary appreciably from a quantitative point of view. Thus, the most important facts to consider are the types of lipids that occur in a given organism and which of these represent the major types. Actual percentages will not be of much use. " M. P. Lechevalier [1] While this quote was made in 1982, it still seems to reflect our understanding of membrane acyl chain heterogeneity today. Researchers have shown that many organisms alter their membrane acyl chain lengths, unsaturation, branching, and cyclization in response to fluctuations in temperature. The percentages are graphed or listed in many combinations, but they still seem to just end up describing an overall trend of changes with no real relationship between the magnitude of the temperature change and the percentages determined. Here we show that there is, in fact, a regular (mathematical) relationship 48 between the acyl chain distribution of an organism and its growth temperature. The reason for the many types of acyl chain alterations that have been documented in response to temperature (and other perturbations) is thought to be an organism's response to maintain a necessary degree of membrane fluidity. This phenomenon is termed "homeoviscous adaption" [2]. Key questions that remain are: How do cells sense temperature? and, What is the connection between this observed membrane heterogeneity and temperature? All cellular metabolic pathways are series of chemical reactions, and these reactions all have activation energy barriers which must be overcome. The driving force to overcome these barriers is heat (infrared quanta). Temperature dictates the available heat that can be used for driving cellular processes, including the synthesis of chemical energy (ATP). One theory is that the membrane lipid acyl chains are able to channel this energy (kT) to the cell interior via special vibrational modes called longitudinal accordian modes (LAMS) [3]. This concept is depicted in Figure 2.1. The sun emits a continuous spectrum of electromagnetic (EM) radiation, which determines the thermal energy spectrum on earth. Though the infrared part of the spectrum is of interest here because it is absorbed as heat in matter. The most abundant quanta at a given temperature (T) has the magnitude kT. This thermal energy spectrum is then transferred to the matter (water, gas, and other molecules) surrounding cells and is stored in their rotational, vibrational, and translational modes. In biological systems water plays an important role in transferring this energy through collisions with other 49 .8285; 82878838 :8 803:8 838—08 .3 Emma 8:25 :8 85 9 8:888: 85 fl $88 $5. .3326 888 86:88 3:63:82 8:8 888 R8 885:88 85 :0 8:2: 8:28.53 :8on E 88?. :8 88582: 85 8 8:888: E $88 $5 82853 852 $838 :8 80658 8388:: km 8:8 2: w:_:::o::m $2822: 82w n 88w £2829: 2.3 u 883 mm 83$ 8:2: 2: .6 8:9: 3:8: 8:283:23 :8 A93 8:285; .30: 8:288: 85 E 888 a $88 :5. 8:88:88 n H :8 .3888 m.::w:§:om n « .588 :0 88:8: u > 3:388 8885 n a 88:3 ..5 u >: E 8288:: 85 mm: 888 8588: 28:: 85 833 :0 32m: 886:: 6:8:me0588 £88 :3 2:. £8.82: :8 :5 3 :52? .8: 25.552: .3 88E 531.88: BEE—zwfiz :2, 385:838 2: 88m 55.58% .388 12:85 85 a: 8.85:: 2: $538: Eauwfin— "fiN «Sumn— vibrational U) C (U I_.. 4...! + :9 > + 4.: O L— em radiation 50 \‘i‘; ,\ 3'7“”? ‘?\/h’J '3: i 2 Figure 2.1 51 molecules and solvent-lattice vibrations. The membrane, due to its location around the cell, has the opportunity to control the thermal energy spectrum that enters the cell. This is because collisions with molecules surrounding the cell and solvent-lattice vibrations result in energy being transferred to special vibrational modes of the membrane acyl chains called longitudinal accordian modes (LAMS), which is the extension and compression of the acyl chains. The net effect is such that the energy quanta are converted into highly tailored, specific, oriented waves on passing through the membrane and into the cell. This energy spectrum is then transmitted throughout the cell interior again via molecular collisions and solvent-lattice vibrations. However, the energy that can be stored and propagated by a given chain length is a function of the chain length, with longer chains storing higher energy quanta. Therefore an acyl chain length distribution is equivalent to an energy distribution. In a rigorous mathematical analysis of the relative energies of these LAM modes, Hollingsworth showed that both the LAM energy distribution in the membrane and its corresponding acyl chain length distribution were normally distributed, as described by the Gaussian density function: f(x) = { 1/[O(21t)"2] }e'l’2[(P-X)2/02 where f(x) is the probability, 0 is the standard deviation, p is the mean, and x is the acyl chain length. He also showed that the most prominent energy stored in the most abundant acyl chain length of the membrane was equal to kT at the growth temperature of the organism. 52 Thus it was proposed that altering of membrane acyl chain lengths allows organisms to transmit the maximal available thermal energy spectrum [3]. It was also shown in this paper that not only is the composition of membrane fatty acid chains normally distributed when ranked by chain length, but there is a relationship between the variance (02) of this distribution curve and temperature as well. Namely, as temperature decreases the distribution curve becomes taller and narrower (smaller 02). As temperature increases the curve becomes shorter and broader (larger 02) [3]. The molecular speed or momentum distribution of the molecules surrounding the cell, which is one aspect of the kinetic energy (translational energy mode) distribution of the molecules, also has a Gaussian type function described by the Maxwell-Boltzmann equation. This equilibrium distribution function {fo(p)} describing the momentum (p) of a dilute gas at temperature T has the form: fo(P) : Ae-(p-pfi)2/2ka where m is the mass of a gas particle, k is the Boltzmann constant, po is the mean momentum, and A is a constant involving the total number of particles (11), m, k, and T. Because the distribution function that describes molecular speed or momentum at a particular temperature has the same form, a relationship between temperature in the Maxwellian distribution function and 02 in the Gaussian density function is invited. This suggests several important questions, the main one of which is: Is there a linear relationship between 02 and temperature over the adaptation range of an organism? If temperature is sensed by alterations of the phase structure of membranes, then membrane-active chemical agents that 53 disturb the packing and fluidize or melt certain regions should have the effect of increasing temperature. On the other hand, those that cause a reduction or freezing of acyl chain motion should have the effect of reducing temperature. These effects are not limited to organic (hydrocarbon) soluble molecules. Hence antibiotics that interfere with the lipid headgroup packing thus disturbing their long range order should have the same fluidizing effects as increasing temperature. Therefore, there should be a temperature equivalence for a variety of membrane-active chemical perturbants. Modulation of phase equilibria is another aspect of membrane structure which may enable cells to sense temperature variations. Membrane components are not uniformly dispersed, rather they are organized into phases or domains which may differ in composition, morphology, or both. The chemical make-up of the constituents (including lipid headgroups, acyl chains, and membrane proteins) dictate the types and structures of the phases. These domains exist in a very complex equilibrium between each other and with the bulk of the membrane. The phase equilibria are extremely temperature sensitive. A small temperature change can result in dramatic alterations in phase structures. Since the phase boundary is a discontinuity, the specific heat should change abruptly across it and small changes in temperature should be amplified. A change in temperature would affect gel to liquid- crystalline and other phase transitions thus altering membrane fluidity. If the new dynamical state does not adequately capture the thermal spectrum, it does so eventually. This is because an essential characteristic of adaptation is that the molecular systems that capture the available heat energy at the maximum of its distribution are the ones that are most viable. 54 One can, therefore, equate a change in temperature with a change in some dynamical property of the membrane and vice-versa. In this study we use a Siberian permafrost bacterium grown at four different temperatures to determine the real dependence between 02 and temperature. We also demonstrate the generality of this relationship using data from the literature regarding three other Gram-positive organisms, Bacillus subtilis, Listeria monocytogenes. and Bacillus amyloliquefaciens. We also investigate the effects of several solvents and antibiotics on Arctic strain 45-3, Bacillus subtilis PY79, and Sarcina ventriculi JK to establish a temperature equivalence for the effect of chemical perturbations on membrane acyl chain distributions. MATERIALS AND METHODS Arctic strain 45-3 culture conditions Arctic strain 45-3 was cultured in 100 ml cultures of trypticase soy broth at 4°C, 14°C, 24°C, and 30°C with shaking at moderate speeds. Effects of cyclohexane and 1,4-cyclohexadiene on Arctic strain 45-3 Cultures (10 ml) of Arctic strain 45-3 were grown as described above except that 0.5 M and 1.0 M concentrations of cyclohexane (54 ul and 108 pl, respectively) and 1,4- cyclohexadiene (47 pl and 95 ul, respectively) were added to individual cultures. 55 Cyclohexane was obtained from Sigma, and 1,4-cyclohexadiene was obtained from Aldrich. Effects of cis- and trans-decalins on Bacillus subtilis PY79 Cultures (10 m1) of B. subtilis PY79 were grown to mid-exponential phase (LB media: 10 g/l tryptone, 5 g/l yeast extract, and 10 g/l NaCl; pH 7.0) at 37°C with shaking at 270 rpm. At this point 0.2 M (0.31 ml) cis—decalin and 0.2 M (0.32 ml) trans-decalin (both obtained from Sigma) were added to individual cultures, and then incubated at 37°C for another six hours. Cells were harvested by centrifugation. Fatty acid methyl ester analyses were done as described below. Effects of antibiotics on Bacillus subtilis PY79 A 2.5 1 culture of B. subtilis PY79 was grown, as described above, to mid-exponential phase, and then split into 5, 500 ml cultures. Then 5 ml of 10 mg/ml stocks of chloramphenicol (Sigma), streptomycin (Boehringer Mannheim), and tetracycline (Sigma) were added to three of the cultures. One was a 37°C control. The last was an ethanol control (5 ml added), since chloramphenicol was dissolved in ethanol. The cells were harvested by centrifugation. The total lipids were extracted two times in 41114 CHCl3zMeOHszO by stirring for 36 hours total. The organic layers were removed and evaporated to dryness. Fatty acid methyl ester analyses were done as described below. Fatty acid methyl ester analysis About 1-2 mg of cells from cultures grown under the various conditions were 56 methanolysed in 5% HCl in methanol at 75°C for 36 hours. The suspensions were evaporated to dryness under a stream of nitrogen, and 0.5 ml HZO, 1.5 ml hexane, and 1.5 ml chloroform were added. After vigorous shaking, the mixture was centrifuged and the organic layer was removed and concentrated to dryness. The resulting fatty acid methyl esters were analyzed by gas chromatography (GC) and GC-MS using a 30 m DB1 column. The temperature program was a ramp from 160°C to 320°C at 3°/min. The final temperature of 320°C was held for 5 min. The relative proportion of the fatty acid components were calculated from the integrated peaks. The fatty acid methyl esters were identified by GC-MS analysis using a Jeol JMS—AXSOSH spectrometer interfaced with a Hewlett-Packard 5890A gas chromatograph. Variance (02) calculations of acyl chain length distribution curves Variances (02) of the acyl chain length distribution curves, which are normally distributed were calculated using the Gaussian density function: fix) = {1/[O(2n:)”2]}e'1’21 0 27‘ 0,18 0.22 ‘ 017 017 ......., HY. .. 277 282 287 292 297 302 307 277 282 287 292 297 302 307 312 Growth Temperaure (Kelvin) Growth Temperature (Kelvin) y = (7.70 x 105)x2 + (4.29 x 102)x + 6.15 y = (~2.86 x 105)x2 + (2.45 x 107)x - 4.42 C 04 0 38 0.36 0.34 3 0.32 .5 0 5 ' > 0.28 0 0,24 0 22 0 293 295 297 299 301 303 307 309 311 288 293 298 303 308 313 Growth Temperature (Kelvin) Growth Temperature (Kelvin) y = (-3.38 x10*‘)x2 + (2.12 x10‘)x - 33.1 y = (-6.11 x1t'r5)x2 + (3.81 x 102)x - 5.59 Figure 2.6 68 length distribution and the growth temperature. Therefore knowing the temperature, one can predict the fatty acid composition over a temperature range for a particular bacterium. The effects of two solvents, cyclohexane and 1,4-cyclohexadiene, on the membrane acyl chain distribution of Arctic strain 45-3 was also determined. These two solvents were chosen because they interact with the hydrocarbon chains and because L. Be’rube’ in our lab demonstrated that cyclohexane had a rigidifying effect on S. ventriculi membranes while 1,4- cyclohexadiene fluidized them. These solvent effects were confirmed by further analysis using T1 NMR experiments (explained in detail in Chapter 5), in which these solvents mixed with DPPC (DL-a-phosphatidlycholine, dipalmitoyl) vesicles indicated the terminal methyl and methylene groups of the acyl chains had greater motional freedom in the presence of 1,4- cyclohexadiene, but were more restricted in the presence of cyclohexane. However, in the case of Artic strain 45-3 both cyclohexane and 1,4-cyclohexadiene acted as fluidizers. This is best shown in the GC chromatograms (Figure 2.7 A, B, and C) in which there is an increase in the amounts of longer acyl chains (namely anteisa-C17 and n-Cm) in the presence of both cyclohexane and 1,4-cyclohexadiene as compared to the 24°C control. This increase in saturated and longer chains fatty acids shows the solvents act like high temperatures and fluidize the membrane. The difference in effects on the Arctic strain 45-3 and S. ventriculi bacteria can be explained by the structures of both the solvents and the main membrane acyl chain structures. As shown in Figure 1.5, the greatest degree of acyl chain motion is at the terminus of the chain in the middle of the bilayer. The T1 NMR 69 Figure 2.7: Gas chromatograms of fatty acid methyl esters obtained from Arctic strain 45-3 cultured in the presence of two solvents. (A) 24°C control, (B) 0.1 M cyclohexane, and (C) 0.1M 1,4-cyclohexadiene. The main acyl chain components were (1) Cm (2) iso- C14 (3) n-Cl4 (4) C15,l (5) am‘eiso-C15 (6) iso-Cl6 (7) n-C,6 and (8) anteiso-Cn. 7O Figure 2.7 71 experiments discussed above, also show that the primary effect of these solvents takes place at these methyl termini of the fatty acids. In Figure 2.8 the packing interactions of cyclohexane and an (o-cyclohexyl acyl chain with a straight hydrocarbon chain are shown. The chair conformations of both these cyclic hydrocarbons enable them to pack closely in a nesting fashion with the methyl terminus of the straight hydrocarbon chain, effectively reducing the motional freedom of the chain resulting in a rigidifying effect. 1,4- Cyclohexadiene, which is a more planar molecule is not capable of this nesting arrangement with the straight acyl chains and as a result fluidizes the membrane. This is what is observed with S. ventriculi, which has primarily straight acyl chains. In the Arctic bacterium, however, the main membrane fatty acids are anteiso- and iso-branched, and as such neither cyclohexane nor 1,4-cyclohexadiene pack well with the branched acyl chain termini, and both therefore would increase the random motion at the methyl termini. 1,4-Cyclohexadiene was particularly toxic to the Arctic bacterium, such that the cells showed limited growth after its addition. This is probably due to the more planar conformation of 1,4-cyclohexadiene which disturbs the packing of the chain termini more, or to its greater diffusibility within the membrane because of the absence of stable packing arrangements along the hydrocarbon chain. Further substance perturbation studies were done using cis- and trans-decalin and three antibiotics (Figure 2.9) on B. subtilis PY79. The GC chromatogram for B. subtilis PY79 at 37°C is shown in Figure 2.10. The primary membrane acyl chains are the branched, odd chain length series. In Figure 2.11 the effects of cis- and trans—decalins on B. subtilis 72 Figure 2.8: Computer generated pictures showing the packing interactions between a straight chain hydrocarbon and cyclohexane and an w-cyclohexyl acyl chain from two different views (A) and (B). Note the trans conformations of the rings allows close packing in a nesting fashion with the methyl terminus of the hydrocarbon chain. 73 Figure 2.8 74 Figure 2.8 (cont.) 75 Figure 2.9: Structures of the chemical perturbants used studies with Bacillus subtilis PY79. (A) cis-decalin, (B) trans-decalin, (C) chloramphenicol, (D) tetracycline, and (E) streptomycin. 76 H Figure 2.10: Gas chromatogram of fatty acid methyl esters of Bacillus subtilis PY79 from a 37 °C culture. The main acyl chain components were (1) iso-Cl4 (2) n-Cl4 (3) iso-Cls (4) anteiso-C15 (5) iso-C16 (6) n-Cl6 (7) iso-Cl7 (8) anteiso-Cl7 (9) n-Cls. 77 0.80 --o--37c 0.70 1 - I- Cis-decalin 0.60 1 + Trans-decalin 0.50 1 0.40 1 Probability 0.30 1 0.20 1 0.10 1 0.00 I 12 13 14 15 16 17 18 19 20 Acyl Chain Length Figure 2.11: The effects of cis- and trans-decalins on the plot of the probability distributions of Bacillus subtilis PY79. 78 acyl chains distributions is shown, in which cis-decalin has no perturbing effect on the membrane while trans—decalin acts as a rigidifier. Again this can be explained by interaction of the perturbants with the type of membrane acyl chains in this organism. cis-Decalin has an L—shaped structure and as such the fluidizing effect of one ring is offset by the nesting capability of the second ring, and thus would not disrupt the membrane packing anymore than the branched acyl chains already cause. The two rings of trans-decalin, on the other hand, are both chair conformers and can nest between the hydrocarbon chains along their length causing an overall decrease in acyl chain motion. The effects of the antibiotics are also related to their structures. Chloramphenicol and tetracycline are largely hydrophobic and interact with the membrane acyl chains. Both of these antibiotics have planar ring systems and would not pack well with the membrane acyl chains. As expected these two antibiotics fluidize the B. subtilis membranes (Figure 2.12). In fact the large ring system of tetracycline fluidizes the membrane to such an extent that the mean of the curve is shifted towards longer chains lengths (similar to that seen by extremely high temperatures). Streptomycin, though, is a polar molecule consisting of sugar derivatives. As a result it most likely interacts with membrane headgroups such that the amidino groups, which are positively charged, interact with the negative phospholipid charges and stabilize the membrane. This headgroup stabilization would result in closer packing of the acyl chains, rigidifying the membrane. Again this predicted effect is demonstrated with B. subtilis, as seen by the taller, more narrow distribution curve for streptomycin as compared to the 37 °C control (Figure 2.12). 79 0.9 "OHS? C Control 0.8 1 - I- EtOH Control 0.7 1 - 1&- - Chloramphenicol 0.6 q +Streplornyern z. - ai- - Tetracyclin E 0.5 1 .D 3 2 0.4 1 n. 0.3 1 0.2 1 0.1 1 O T T T l I If T 12 13 14 15 16 17 18 19 20 Acyl Chain Length Figure 2.12: Effects of three antibiotics on the plot of the probability distributions of Bacillus subtilis PY79. 80 These results indicate that the effect of chemical substances on acyl chain length distributions can be explained based on their structures and the membrane acyl chain components of a given organism. As a result there is a temperature equivalent shift in membrane acyl chain length compositions based on whether the perturbant acts as a fluidizer or rigidifier of the particular membrane. The ability of organisms to adapt to environmental perturbations by alteration of their acyl chain distributions would be reflected in their growth rate under these new conditions. These results indicate that there might be a relationship between temperature, solvent concentration, and growth rate. In Figure 2.13 hypothetical growth rate curves for conditions of high and low temperatures and solvent concentrations. Such a three dimensional graph of these factors for a given bacterium would allow one to determine the solvent concentration needed to offset the temperature shift to allow the organism to grow outside its temperature range. A temperature offset could also be determined to allow a bacterium to grow in the presence of higher solvent concentrations. In living systems something must fuel the molecular motion of the cellular processes, since these processes are not self-sustaining. This driving force is the heat (kinetic energy), determined by temperature, available to the organism. Therefore, there has to be a match between the thermal energy spectrum and the cellular chemical properties, and this match is characteristic of a given organism. Organisms capture the available kinetic energy by collisions with molecules whose momentum distribution is a Gaussian type function, the mean and variance of which is determined by temperature according to the Maxwell- Boltzmann distribution law. The collisions between the surrounding solvent and gas 81 & ¥ temperature 0 [solventL “3. O 2. 61 n ‘32: Figure 2.13: A three dimensional graph showing hypothetical growth rate curves for conditions of high and low temperatures and solvent (fluidizing type) concentrations. 82 molecules and the molecules in the boundary structures of the cells form the basis of the heat exchange process. To effectively capture this kinetic energy distribution, the surface of the cell must have an array of molecules or some molecular system that can absorb the energy given up in these exchanges into some vibrational mode because the cellular chemistry is driven by these vibrations. So then the best adapted organism should have some energy harvesting system that traps this energy that is passed on in these collisions in vibrational modes with the same mean energy as in the collisions and over a distribution that is narrowly defined around the mean energy. Since the energy distribution of the collisions to start with is Gaussian as well, it makes sense that the vibrational modes are also Gaussian, with the most abundant acyl chain length in the membrane trapping the most abundant energy quanta stored in the matter surrounding the cell. The energy trapped by the membrane and propagated to the cell interior is used to drive the cellular chemical processes. Not only does this energy determine the rate of diffusion of molecules (and thus the collision rate), it also dictates various motions of proteins that determines the activity of these proteins, and as a result the types of cellular processes that can be carried out. Two types of large-scale motions observed in proteins are hinge bending motions and single strand motions, both of which may produce large relative displacements of atoms within proteins and are usually involved the biological activity of the protein. As will be discussed the frequencies, by relation to the time scale of the motions, of these motions are also of the magnitude kT. 83 In 1995, Hollingsworth showed that the frequency of the most abundant membrane acyl chain length corresponded to 215.4 cm’1 (in the spectroscopic unit of wavenumbers) or 6 psec", which corresponded to kT at the growth temperature of the organism [3]. This energy (as well as the rest of the energy distribution stored in the acyl chains) is then transferred to the cell interior, mostly to the water molecules. Recently terahertz (Thz) laser vibration-rotation-tunneling spectroscopy has been used to characterize the frequencies of the water solvent cage. The first calculations were for trimers of water molecules, and they found that the characteristic frequency (in wavenumbers) for one aspect of water vibrations was 40-50 cm’l [8]. For water tetramers this frequency increased to 2.04 Thz (67.9 cm") [9]. This frequency continued to increase with the size of the water cage to 2.4 Thz (81.2 cm-l) for pentamers [10] and 2.491 Thz (83.03 cm-l) for hexamers [1 1]. So one would expect that as this water cage approached infinity, this water vibrational frequency would get larger and eventually settle to a frequency probably in the range of 100-200 cm‘1 (since the rate of increase in frequency between the pentamer and hexamer is smaller), which would correspond to the characteristic energies transmitted by the membrane acyl chains through their LAM modes. This is also only one aspect of water cage vibrations, and there are certainly many others that would contribute to the total energy that is transmitted by the solvent molecules. This energy stored in the solvent molecules in the cell interior is now redistributed among the other low frequency modes of proteins and other molecules (such as DNA and RNA) for use in carrying out many of the cell processes. Many enzymes and binding proteins consist of several globular domains connected 84 by relatively flexible polypeptide chains [12-14]. The binding sites for substrates and ligands are typically located between these domains, and thus the time scale and relative motion of the domains would play an important role in the binding and activity of the enzymes (proteins). These large scale motions would involve collective motions with the concerted displacement of many atoms. Normal mode calculations done on the small protein bovine pancreatic trypsin inhibitor (BPTI) showed that some of these large scale motions were low frequency modes on the order of 3.6-6 ps'l (100—200 cm"), which typically involved motion of several neighboring residues. More global motions have frequencies less than 3.6 ps‘1 (100 cm") [15]. Adiabatic mapping calculations for hinge bending motions, one type of global motion, were first done for lysozyme, in which its active site is located in a cleft between two globular domains [16]. These calculations, and further improved calculations that take into account solvent damping effects, indicate that the hinge bending motion is greater than 10 ps (less than 0.1 ps"). In fact with solvent damping effects, the hinge motion is expected to be Brownian in nature with a typical fluctuation opening and closing the cleft by 0.1 nm and will persist for about 20 ps [17]. In regards to other types of proteins, fluorescence depolarization techniques were used to determine the time scale of hinge bending motions of immunoglobulins is in the range of 10-50 ns [18, 19]. Also, indirect evidence provided by X-ray crystal structures indicate that tRNA molecules may even have hinge bending motions [20]. Initial investigations of this movement for tRNAPh" have been done [21, 22], and molecular dynamics simulations show a periodic motion of the bending between the two arms of the tRNAPhc molecule which is 24 ps [23]. It is expected that solvent damping effects would increase this time scale to the characteristic nanosecond range 85 [12]. It has also been shown that several proteins have biologically important strand folding and unfolding motions, which often occur in peptide segments near binding or active sites. Upon binding of the substrate or ligand these segments assume ordered conformations in which they cover or otherwise interact with the bound molecule [12-14]. For example, the active site of trypsin, which consists of several polypeptide loops, is highly ordered; but in its inactive precursor, trypsinogen, these loops are disordered. The apparent characteristic time scale of this loop motion is about 11 ns [24], which is in agreement with studies done on the growth kinetics of alpha helices [25]. The time scale of these global motions of proteins and other macromolecules are limited by the viscosity of the solvent, which is water in biological systems. Thus solvent damping impedes these global motions, and hence their rates of motion are diffusion limited and are in a very low frequency range (less than 100 cm"). Based on all of this evidence, though some of it is indirect from various types of simulations, one can perceive that the energy transmitted by the membrane acyl chains is disseminated among the many vibrational modes of the solvent cage (and other molecules) in the cell interior. These numerous solvent-lattice vibrational modes store and propagate many frequencies of energy, in the low frequency range, to the proteins and other large molecules in the cell, which are then stored in these molecules in their various vibrational modes, thus allowing them to carry out the myriad cellular processes. In this chapter, focus is on the acyl chain length heterogeneity of membranes and their 86 function in transmitting the available thermal energy spectrum. However, membrane lipid acyl chain heterogeneity also includes differences in unsaturation, branching, and the presence of carbocyclic rings. It has been shown that functionalization of an acyl chain by single alkenyl, methyl, and cyclopropyl groups does not cause a significant difference of the acyl chain LAM frequencies [26]. These functionalizations, though, do affect the packing arrangement of the membrane acyl chains. Thus these alterations in acyl chain structure most likely affect other aspects of membrane motional dynamics as is predicted by the homeoviscous adaption theory. As a result these changes may play important roles in maintaining other aspects of membrane function, such as their selective permeability functions or control of certain enzymatic processes by altering the membrane domain compositions. REFERENCES l. Lechevalier, M. P. ( 1982) Lipids in bacterial taxonomy. In CRC Handbook of Microbiology, 2nd ed., Volume IV, eds. A. I. Laskin and H. A. Lechevalier. CRC Press, Inc., Boca Raton, FL, p. 436. 2. Sinensky, M. ( 1974) Homeoviscous adaptation--a homeostatic process that regulates the viscosity of membrane lipids in Escherichia coli. Proc. Nat. Acad. Sci USA, 71: 522-525. 3. Hollingsworth, R. I. (1995) Membrane lipid alkyl chain heterogeneity is a molecular transform of the available thermal energy spectrum. J. Phys. Chem, 99: 3406-3410. 4. Tadayon, R. A., and K. K. Carroll. (1971) Effect of growth conditions on the fatty acid composition of Listeria monocytogenes and comparision with the fatty acids of Erysipelothrix and Corynebacterium. Lipids, 6: 820-825. 5. Paton, J. C., E. J. McMurchie, B. K. May, and W. H. Elliott. (1978) Effect of growth temperature on membrane fatty acid composition and susceptibility to cold shock of Bacillus 87 amyloliquefaciens. J. Bacteriol., 135: 754-759. 6. Suutari, M., and S. Laakso. (1992) Unsaturated and branched chain-fatty acids in temperature adaptation of Bacillus subtilis and Bacillus megaterium. Biochim. Biophys. Acta, 1126: 119-124. 7. Weerkamp, A., and W. Heinen. (1972) Effect of temperature on the fatty acid composition of the extreme thermophiles, Bacillus caldolyticus and Bacillus caldotenax. J. Bacteriol., 109: 443-446. 8. Viant, M. R., J. D. Cruzan, D. D. Lucas, M. G. Brown, K. Liu, and R. J. Saykally. (1997) Pseudorotation in water trimer isotopomers using terahertz laser spectroscopy. J. Phys. Chem, 101: 9022-9041. 9. Cruzan, J. D., M. R. Viant, M. G. Brown, and R. J. Saykally. (1997) Terahertz laser vibration-rotation tunneling spectroscopy of the water tetrarner. J. Phys. Chem, 101: 9022- 9031. 10. Liu, K., M. G. Brown, J. D. Cruzan, and R. J. Saykally. (1997) Terahertz laser spectroscopy of the water pentamer: structure and hydrogen bond rearrangement dynamics. J. Phys. Chem, 101: 9011—9021. 11. Liu, K., M. G. Brown, and R. J. Saykally. (1997) Terahertz laser vibration-rotation tunneling spectroscopy and dipole moment of a cage form of the water hexamer. J. Phys. Chem, 101: 8995-9010. 12. McCammon, J. A., and S. C. Harvey. (1987) Global structural changes. In Dynamics of Proteins and Nucleic Acids, Cambridge University Press, Cambridge. pp. 137-150. 13. Bennett, W. S., and R. Huber. (1984) Structural and functional aspects of domain motions in proteins. CRC Crit. Rev. Biochem., 15: 291-384. 14. Petsko, G. A., and D. Ringe. (1984) Fluctuations in protein structure from X-ray diffraction. Ann. Rev. Biophys. Bioeng., 13: 331-371. 15. G6, N., T. Noguti, and T. Nishikawa. (1983) Dynamics of a small globular protein, in terms of low-frequency vibrational modes. Proc. Natl. Acad. Sci. USA, 80: 3696-3700. 16. McCammon, J. A., B. R. Gelin, and M. Karplus. (1976) The hinge-bending mode in lysozyme. Nature, 262: 325-326. 17. McCammon, J. A., and P. G. Wolynes. (1977) Nonsteady hydrodynamics of biopolymer motions. J. Chem. Phys, 66: 1452-1456. 88 18. Yguerabide, J ., H. F. Epstein, and L. Stryer. (1970) Segmental flexibility in an antibody molecule. J. Mol. Biol, 51: 573-590. 19. Holowka, D. A., R. E. Cathou. (1976) Conformation of immunoglobulin M. 2. Nanosecond fluorescence depolarization analysis of segmental flexibility in anti-e-l- dimethylamino-S-naphthalenesulfonyl-L-lysine anti-immunoglobulin from horse, pig, and shark. Biochemistry, 15: 3379-3390. 20. Westhof, E., P. Dumas, and D. Moras. (1985) Crystallographic refinement of yeast aspartic acid transfer RNA. J. Mol. Biol, 184: 119-145. 21. Harvey, S. C., and J. A. McCammon. (1981) Intramolecular flexibility in phenylalanine transfer RNA. Nature, 294: 286-287. 22. Tung, C-S., S. C. Harvey, and J. A. McCammon. (1984) Large-amplitude bending motions in phenylalanine transfer RNA. Biopolymers, 23: 2173-2193. 23. Harvey, S. C., M. Prabhakaran, and J. A. McCammon. (1985) Molecular-dynarnics simulation of phenylalanine transfer RNA. 1. Methods and general results. Biopolymers, 24: 1169-1188. 24. Butz, T., A. Lerf, and R. Huber. (1982) Intramolecular reorientational motion in trypsinogen studied by perturbed angular correlation of 199"“Hg labels. Phys. Rev. Lett., 48: 890-893. 25. McCammon, J. A., S. H. Northrup, M. Karplus, and R. M. Levy. (1980) Helix-coil transitions in a simple polypeptide model. Biopolymers, 19: 2033-2045. 26. Hsu, S. L., G. W. Ford, and S. Krimm. (1977) Longitudinal acoustic mode in polymers. II. Solution of the general composite perturbed elastic rod. J. Polym. Sci., Polym. Phys. Ed., 15: 1769-1778. CHAPTER III THE INTERRELATION BETWEEN TEMPERATURE-INDUCED SHIFTS IN CARBON METABOLISM AND MEMBRANE LIPID CHEMISTRY 89 90 ABSTRACT The available thermal energy, determined by temperature, drives the chemical processes carried out in living organisms. The ability to utilize a particular substrate to derive chemical energy (typically ATP) is dependent upon the transition state (activation) energies of all of the steps in the metabolic pathway for that substrate. If the available thermal energy is not enough to overcome the activation energy for one or more steps in the metabolic process, the pathway requires significant energy investment or priming, and/or the energy yield is too low, then the organism might not be able to utilize the given substrate effectively. This is true for the more highly oxidized substrates like carbohydrates. As a result, the balance of metabolism might conceivably shift to utilization of more reduced substrates, like amino acids and fatty acids. The utilization of which might require lower activation energies and certainly requires less of a chemical energy investment and can yield much more chemical energy per gram. The notion that more energy is required to support cellular upkeep at lower temperatures is well established and embodied in the concept of maintenance energy. Based on this premise we investigated the ability of Arctic strain 45-3, which thrives in extremely cold environments, to utilize glucose at 4°C and 24°C. One special advantage of such a strain is that it allows a substantial spread of temperature. Most organisms that are available in culture do not grow well below 15°C. Our results, using NMR spectroscopy and radioactively labeled glucose, indicate that this strain utilizes glucose much more slowly at 4°C than at 24°C. We also show that enzymatic control of metabolic processes, such as carbohydrate and amino acid metabolism, must be done by regulation of constitutively expressed enzymes because the total cellular protein profiles at the two 91 temperatures are essentially identical. The fact that this strain can use glucose more readily at 24°C is in agreement with our observation that more straight membrane acyl chains are produced at 24°C than at 4°C. These results support the idea that the available thermal energy regulates the balance of chemical processes carried out in the cell, and this balance determines the metabolic intermediate profile which is reflected in the membrane acyl chain structure. INTRODUCTION As temperature fluctuates the amount of heat (infrared quanta) available for use by cells changes in turn. This energy (kT), which is propagated into the cell interior by the longitudinal accordian modes (LAMS) of the membrane acyl chains, provides the necessary energy for various metabolic pathways [1]. In Figure 3.1 hypothetical energy profiles for two different metabolic pathways are shown. At high temperatures (Tl) there is most likely enough energy to carry out either metabolic pathway. At lower temperatures (T2), though, the available energy may not be enough to get over the activation barriers of the pathway indicated by profile A, so the cell has no choice but to use the metabolic pathway indicated by profile B. In general the more oxidized substrates, such as carbohydrates, require more chemical energy input in their metabolism. More reduced substrates such as amino acids and fatty acids require less energy input. The structure of the cell membrane reflects the type of metabolism carried out by the 92 Energy Reaction progress Figure 3.1: Hypothetical energy profiles for two different metabolic pathways are shown, a higher energy requiring profile in (A) and a lower energy requiring profile in (B). Note that at high temperatures (T,) there is most likely enough energy to carry out either metabolic pathway. At lower temperatures (T2), though, the available energy may only be enough to use the metabolic pathway indicated by profile B. 93 cell. Evidence of this is suggested by the literature documenting the greater occurrence of branching in membrane fatty acids at low temperatures [2, 3]. The structures of these fatty acids are determined by the acyl-CoA primers. For the branched fatty acids the primers result from metabolism of the amino acids, leucine, valine, and isoleucine [4]. The tendency to synthesize branched fatty acids at lower temperatures therefore could be interpreted to mean that amino acid metabolism is more favorable than carbohydrate metabolism at low temperatures. Changes in metabolism can be accompanied by changes in either the activity or the levels of synthesis in the enzymes involved. The controls of metabolic pathways could be accomplished by de novo protein synthesis, or induction or suppression of constitutively expressed enzymes. Changes in metabolism with temperature can also occur in the absence of both these two control mechanisms. This is because at higher temperatures, there may be enough activation energy to make some pathways active that are not active at low temperatures. Here we studied the ability of an Arctic permafrost bacterium to utilize glucose over its growth temperature range as well as monitor its protein profiles at two temperatures. This bacterium, Arctic strain 45-3, utilizes glucose much more slowly when grown at 4°C than at 24°C. The protein profiles for this bacterial strain at 4°C and 24°C are essentially the same, indicating that the control of different metabolic pathways likely occurs through thermal or constitutive enzyme control. 94 MATERIALS AND METHODS Culture conditions of Arctic strain 45-3 Siberian strain 45-3 was cultured in trypticase soy broth (TSB) at 4°C and 24°C with shaking at moderate speeds. Analysis of trypticase soy broth before and after bacterial growth Samples of TSB media before and after Arctic strain 45-3 was cultured at 4°C and 24°C (both cultures were grown to about the same optical density) were taken and concentrated to dryness. The samples were then resuspended in 0.6 ml of DZO. 1H NMR spectroscopy was done on each sample. Two dimensional DQF-COSY (double quantum filtered-correlated spectroscopy) and TOCSY (total correlated spectroscopy) spectra were also done for the unused TSB media sample. Rate of utilization of carbon sources by Arctic strain 45-3 Five milliliter aliquots of TSB media were made, and 0.2 [Ci of l4C-glucose, l4C- leucine, and l4C-valine (each) were added to the media. Then the media was inoculated with Arctic strain 45-3, and cultures were grown at 4°C and 24°C. Samples (100 pl) were taken before the cells were added and then every 4 hours for at least 36 hours (longer for the 4°C cells). The cells were centrifuged out and then 20 ul of the media were spotted on TLC plates. The metabolites on the TLC plates were separated using a solvent system of 7:3:1 1- propanolzammonium hydroxideszO. The plates were then exposed to a phosphoimager. 95 The intensities of spots were determined using the integration routines available in the IrnageQuantTM package. Total protein profiles of Arctic strain 45-3 grown at 4°C and 24°C Cultures of Arctic strain 45-3 at 4°C and 24°C were grown as described above. Total cell lysates were prepared from 4°C and 24°C cultures by adding two volumes of 10 mM Tris-HCl, pH 8.0 and 3 mg/ml lysozyme and incubating for one hour on ice. One volume of 2X SDS-PAGE sample buffer solution [0.125 M Tris-HCL (pH 6.8), 0.04% sodium dodecylsulphate (SDS), 10% glycerol, 5% B-mercaptoethanol, 0.0025% bromophenol blue] was then added. The samples were mixed well, boiled for 5 minutes, and centrifuged briefly to remove any precipitate. The lysates from 4°C and 24°C cultures were separated using Laemmli’s discontinuous buffer system [5] on 12% polyacrylamide slab gels containing 0.32% bisacrylarnide, 0.375 M Tris-HCl (pH 8.8), and 0.1% SDS. The stacking gel contained 4% polyacrylamide, 0.1% bisacrylamide, 0.125 M Tris-HCl (pH 6.8), and 0.1% SDS. Electrophoresis was carried out at 80 V for about 3 hours. Gels were stained in a solution of 0.1% Coomassie blue in 40% methanol/10% acetic acid, and destained in a solution of 40% methanol/10% acetic acid. RESULTS AND DISCUSSION Arctic strain 45-3 has predominantly branched chain fatty acids indicating the use of amino acid derived acyl-CoA primers for fatty acid synthesis. While amino acid metabolism 96 predominates in generating these primers throughout the growth temperature range, the amount of C16 straight chain fatty acids increases by almost a factor of six in going from 4°C to 30°C (Table 3.1), which is consistent with an increased utilization of carbohydrates. Also at 24°C and 30°C, C18 straight chains (not shown) are detected. Based on the reasons that the Table 3.1: Percent of fatty acid methyl ester derivatives in the total methanolysate of the cells grown at 4°C, 14°C, 24°C, and 30°C in order of elution. The numbered components correspond to the numbered GC peaks in Figure 4.12. Peak No. Structure % in 4°C % in 14°C % in 24°C % in 30°C Cells Cells Cells Cells 1 CW 1.30 1.41 0.00 0.00 2 iso-CI4 2.65 3.18 2.24 1.08 3 n-Cl4 1.55 3.02 2.78 3.74 4 C15,, 17.58 9.09 0.87 0.28 5 anteiso-C15 62.67 64.53 68.82 60.87 6 iso-Cl6 8.72 10.22 10.38 6.31 7 n-C16 3.11 5.43 8.61 17.69 8 anteiso-C ,7 2.42 3.10 6.31 10.02 pathway requires significant energy investment or priming or the energy yield is insufficient when the available thermal energy is too low, it appeared possible that carbohydrate metabolism might be less effective at lower temperatures than other forms of metabolism. We investigated the ability of Arctic strain 45-3, which survives in extremely cold climates, to utilize glucose and amino acids at 4°C and 24°C. 1H NMR of TSB media before (Figure 97 3.2A) and after growth of Arctic strain 45-3 at 4°C (Figure 3.2B) and 24°C (Figure 3.2C) are shown. The signals for the oc- and B-anomeric protons of glucose at 5.18 and 4.58 ppm, respectively, have almost completely disappeared from the media of the 24°C culture, while the relative decrease in these signal intensities for the 4°C culture is much smaller. To further investigate this bacterium's ability to metabolize glucose, l4C-labeled glucose, leucine, and valine were added to the TSB media for 4°C and 24°C cultures. Figure 3.3 shows the TLC separation of l4C-glucose, l4C-valine, and l4C-leucine for one time point and its corresponding analog profile indicating the relative intensities of the three spots. Figure 3.4 shows the extent of utilization of l4C-glucose by the bacterium at 4°C and 24°C. The time points at 24 hours and 67 hours for the 24°C and 4°C, respectively, are when the two cultures are at about the same optical density. The results from this radioactive label experiment confirm the NMR results, which indicate that Arctic strain 45-3 utilizes glucose at a greatly reduced rate at 4°C, while at 24°C the glucose becomes limiting very quickly. Since the trypticase soy broth medium is largely peptides and amino acids (17 g/l pancreatic digest of casein; 3 g/l papaic digest of soybean meal; and 2.5 g/l glucose), the percent reduction of l4C-valine and l4C-leucine was small. These results, however, support our hypothesis that the ability to use more oxidized substrates at low temperatures is impeded. Thus the bacteria shift the balance of their metabolic pathways to use more reduced substrates, amino acids for instance. To determine whether this shift in metabolism is reflected in variations of this strain's 98 Figure 3.2: ‘H NMR of TSB media before and after growth of Arctic strain 453. (A) Unused TSB media, (B) TSB media from 4°C culture, and (C) TSB media from 24°C culture. Most of the glucose signals are in the 3.2-3.6 ppm range, while some of the free amino acid and polypeptide signals are in the 1.0-1.5 ppm range. However, note the almost complete disappearance of the (l) a-anomeric [5. 18 ppm] and (2) B-anomeric [4.58 ppm] glucose signals in the 24°C culture, as compared to the unused media and the 4°C culture. 99 1 2 l l 1 2 1 2 l 1‘ 1 Figure 3.2 Geranium LINE 1 glucose valine leucine Figure 3.3: TLC separation of l‘C-glucose, l4C-valine, and 1“C-leucine for one time point and its corresponding analog profile indicating the relative intensities of the three spots. 101 50 251 Percent 201 151 101 ...--24c +40 0 1o 20 30 4o 50 60 7o 80 Hours Figure 3.4: The extent of utilization of l4C-glucose by Arctic strain 45-3 at 4°C and 24°C. The time points at 24 hours and 67 hours for the 24°C and 4°C cultures, respectively, are when the optical densities for the two cultures are about equivalent. 102 enzyme production, total protein profiles from 4°C and 24°C cultures were determined. These profiles are shown in Figure 3.5. The profiles for the two temperatures are essentially the same, showing no appearance or disappearance of different proteins at the two temperatures. The differences seen in some band intensities is not consistent upon repeated culturing of the organism at the two temperatures. In many gels of the total cellular proteins from subsequent cultures of this organism at the two temperatures, no significant differences in band intensities was seen. This indicates that regulation of the metabolic pathways in this organism must occur largely by induction or suppression of constitutive enzyme activity, or by thermal regulation in which higher temperatures may provide enough activation energy to make some pathways active that are not active at lower temperatures. As has been discussed previously, there is a relationship between the types of metabolism carried out by an organism and its membrane acyl chain structure. This relationship is represented in Figure 3.6. The balance between carbohydrate and amino acid metabolism, for instance, can be regulated by de novo enzyme synthesis, or by activation or suppression of constitutively expressed enzymes, or by thermal regulation as described above. The balance between these two types of metabolism determines the available acyl- CoA primers for fatty acid synthesis. An increase in amino acid metabolism, in particular the levels of leucine, isoleucine, and valine, would increase the pool of branched acyl-CoA primers. In contrast, an increase in carbohydrate metabolism would increase the pool of acetyl-CoA leading to straight acyl chains. 103 199 kDa m l 15 kDa Figure 3.5: Total protein profiles for Arctic strain 45-3. Lane (1) low molecular weight standards, (2) 4°C, and (3) 24°C. 104 Carbohydrate ‘ , Amino acid metabolism metabolism (more oxidized substrates) (more reduced substrates) (higher activation barriers?) (lower activation barriers?) (high initial energy investment) (low initial energy investment?) \ / enzyme suppression or de novo synthesis versus thermal regulation of constitutive enzyme activity 1 1 Available acyl-CoA primer pool for fatty acid synthesis Figure 3.6: Schematic representation of the relationship of the balance between two types of cellular metabolism, the potential thermal or enzymatic control of the metabolic processes, and the pools of available acyl-CoA primers for fatty acid biosynthesis. 105 The balance between these two metabolic pathways is controlled by the available thermal energy. Since carbohydrates are more oxidized substrates, they generate less energy per gram (669.94 kg-cal/g for glucose) than amino acids (855.6 kg-cal/g for leucine) [6]. Carbohydrates also require more energy to prime (in phosphorylation steps) and is oxygen dependent. In low temperature climates not only is the available thermal energy reduced, but the oxygen availability is also much less (due to decreased diffusion of oxygen through cold or frozen soils). So as temperature drops the highly oxygen dependent and energy costly metabolic pathways, i. e. carbohydrate metabolism, shut down in favor of utilization of the more reduced substrates, amino acids and fatty acids. This is also consistent with the fact that maintenance energy is higher at lower temperatures [7-9]. Therefore organisms that use energy sources that yield more energy per gram would be more successful at low temperatures as well. Therefore the balance of metabolic processes, which is determined by the available thermal energy, is reflected by the structure of the membrane acyl chains. At higher temperatures more straight chains are produced, while at lower temperatures more branched chains are made. Since the available thermal energy determines the balance of cellular metabolic processes, the outcome of competing cellular reactions is determined by kinetics (the activation energy barriers of the steps in the pathways) and not thermodynamics (the difference in free energies of the starting materials versus the products). In other words, if two chemical processes in the cell are competing, then the short-term outcome of the competition is determined by the activation energy barriers the two processes face. The 106 ability to overcome these barriers is determined by the available thermal energy spectrum that has been transferred to the cell interior. One reaction may have a lower activation energy barrier, but the resultant product has a higher free energy content. So these products are kinetic products, and they accumulate the fastest because of the lower activation barriers to their formation and are not determined by the final free energies of these products. If the cellular systems were closed systems and were in equilibrium, these higher energy kinetic products, if there is some reversibility in the processes, would revert to the lower energy thermodynamic products. However, cellular metabolic pathways are an intricate web of reactions, and the kinetic products are often entrained in subsequent reactions such that the entire process is determined by kinetics and not by thermodynamics (free energy differences). Also in living systems, energy is constantly being fed in as heat, light, and/or nutrients, and energy is constantly being given off in completely unpredictable patterns in swimming, crawling, or walking. Thus an equilibrium state of cellular processes, as would be required for thermodynamic control of chemical processes, is never reached. Consequently, the nature of living systems is determined by kinetics, not thermodynamics. REFERENCES 1. Hollingsworth, R. I. (1995) Membrane lipid alkyl chain heterogeneity is a molecular transform of the available thermal energy spectrum. J. Phys. Chem, 99: 3406-3410. 2. Russell, N. J., and N. Fukunaga. (1990) A comparison of thermal adaptation of membrane lipids in psychrophilic and thermophilic bacteria. FEMS Microbiol. Rev., 75: 171-182. 3. Suutari, M., and S. Laakso. (1994) Microbial fatty acids and thermal adaptation. CRC 107 Crit. Rev. Microbiol., 20: 285-328. 4. Kaneda, T. (1991) 180- and anteiso-fatty acids in bacteria: biosynthesis, function, and taxonomic significance. Microbiol. Rev., 55: 288-302. 5. Laemrnli, U. K. (1970) Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature, 227: 680-685. 6. Weast, R. C., ed. (1985) Heat of combustion. In CRC Handbook of Chemistry and Physics, 66th ed. CRC Press, Inc., Boca Raton, FL. pp. D272-277. 7. Buchanan, R. E., and E. I. Fulmer. (1928) Physiology and biochemistry of bacteria. The Williams & Wilkins Co., Baltimore. 8. Rose, A. H. (1976) Chemical Microbiology: An Introduction to Microbial Physiology, 3rd Ed. Plenum Press, New York. pp. 245-325. 9. Russell, J. B., and G. M. Cook. ( 1995) Energetics of bacterial growth: Balance of anabolic and catabolic reactions. Microbiol. Rev., 59: 48-62. CHAPT ER IV THE ADAPTATION OF ARCTIC BACTERIA TO LOW PRE-FREEZIN G TEMPERATURES INVOLVES MEMBRANE ALTERATIONS BUT NOT INCREASED PRODUCTION OF OSMOLYTES 108 109 ABSTRACT An Arctic bacterial strain isolated from the Siberian permafrost grown at 4°C produced undetectable levels of osmolytes such as glycerol, sugars, other sugar alcohols and glycine betaine in the cytoplasm. The relative proportions in the total carbohydrate and amino acid metabolic profiles between cells cultured at 4°C and 24°C using NMR spectroscopy showed no significant changes. There was a net reduction by a factor of 4.6 in the total amount of amino acids at 4°C compared to 24°C. In the cells cultured at 4°C, major changes in membrane chemistry and cell surface chemistry were observed compared to those grown at 24°C. There was a change in composition and a dramatic reduction in the amount of polysaccharide capsule formed when cells were grown at 4°C compared to 24°C. Because of this, cells from the lower temperature culture were hardly wettable. This would negate the effect of freezing water on the membrane surface. Cells grown at the lower temperature were smaller in volume by a factor of ~14. These results indicate that the major mechanism for low temperature adaptation in this organism is alteration of the surface and membrane structure. INTRODUCTION The processes bacteria use to allow them to survive and function at close to freezing temperatures and below is a mystery that has fueled much interest and spawned several theories. It is generally believed that there are two major physical or physico-chemical threats that the bacterium needs to respond to in these temperature regimes. The first is ice formation 110 within the cell which might lead to cell lysis because of the volume increase on expansion of water as ice is formed. The second is the increased salinity outside the cell as ice formation leads to the separating out of pure water (as ice) and a corresponding increase in salt concentration in the liquid state. It is thought that a general mechanism to counteract both of these phenomena would be an increase of solute concentration inside of the bacterial cell. A reasonable response then would be for the organism to increase the intracellular amounts of certain metabolites especially glycine betaine, glycerol, mannitol and sorbitol both as cryo-protectants and as osmolytes (1-3). There are certain to be other mechanisms at work to protect against freezing since as temperature decreases the solubility of most organic molecules and salts decreases. The decrease is rather dramatic in the case of phosphates and carbonates. This solubility trend is in the opposite direction to that required for freeze protection by increasing osmolyte concentration. There are other aspects of bacterial cell structure that can contribute to this adaptation. These include membrane structure, capsular polysaccharide structure and amount and cell size. For instance, the membrane can be protected from freezing water by increasing its hydrophobicity. This can be accomplished by changing the structure and reducing the amount of the hydrophilic capsule. A change in structure might lead to a microenvironment that reduces the tendency of water associated with it to freeze. Making the actual surface more hydrophobic by removing charged groups from the lipids will reduce the interaction between the actual membrane surface and external water. Psychrophilic bacteria (organisms that are adapted to these near-freezing and below 111 conditions) are very good systems in which to examine the mechanisms of bacterial adaptation to these extremes. In this study, we used a strain of bacteria that was obtained from the permafrost sediment core of the age 60- 100 thousand years deposited in the Siberian tundra zone on Kolyma-Indigirka Lowland (152-162°E and 68-72°N), see Figure 4.1. The permafrost sample was received by rotary drilling without any solution or chemical reagents. The temperature of the extracted core was no higher than -7°C. The methods of sampling, storage, transportation and contamination control were described previously (4). Using these methods, it was shown that the micro flora in the samples were there in situ and were not the result of contamination. The organism studied was isolated from a sample at a core depth of 4.0 meters and was named 45-3. The bacterium was a gram positive and according to its morphological properties and 16S rRNA sequence (Jim Tiedje, Center for Microbial Ecology, Michigan State University, Michigan; personal communication) was an Arthrobacter sp. The bacterial morphology existed as rods and V-forms in young cultures and coccus type in old cultures. The strain has a temperature range for growth of 4 to 30°C and an optimal growth temperature of 23°C. MATERIALS AND METHODS Culture conditions Siberian strain 45-3 was cultured in trypticase soy broth at 4°C and 24°C with shaking at moderate speeds. 112 .Q 28 Soc 038mm 88 a So: @2202 £3 m-mv Edam Alembic ES mouefiém: ecu—Bed axnmwmcEéEm—ov— 5325— 5893 -53: «o 0:8 «.655 2: E :Ewou wEEEam 3.58.3 «mamas—Lon car—2.5 “:0 9.53 113 3. 23E 0 RiVer [ya 0 fl 2% a 8985 f - A 582-3 . 114 Hydrophobicity analysis Cells from 4°C and 24°C cultures were diluted to the same optical density. Glass coverslips coated with chlorotrimethylsilane were dipped into the cultures for several minutes and then rinsed thoroughly in a stream of fresh media. The coverslips were placed on slides and the cells were then stained with several drops of 0.1% acridine orange for five minutes. The coverslips were again washed thoroughly using fresh media. The cells were viewed using rhodamine iso-thiocyanate (RITC) optics with excitation and primary barrier filters at 529 and 550 nm respectively. Total lipid extraction Cells from the 4°C and 24°C cultures were stirred with 60 ml H20, 130 ml CHC13, and 15 ml MeOH at 30°C for 12 hours. The layers were separated by centrifugation. The aqueous layers were extracted once more with 30 ml CHCl3 for 15 min., and the organic layers were then combined. The aqueous layers were again centrifuged and the clear solutions were kept. The organic layers were analyzed by NMR spectroscopy and by thin layer chromatography using 10:4:2:2:1 chlorofonn/acetone/methanol/acetic acid/water. TLC analyses were performed on silica gel plates. The spots were visualized by spraying either with orcinol or 10% phosphomolybdic acid in ethanol and heating at 120°C to visualize the organic components. Polysaccharide isolation The aqueous layers from the total lipid extracts were concentrated to 4 ml and 12 ml of ethanol was added. The polysaccharide which precipitated was recovered with a glass rod 115 and dissolved in 10 ml with water, and 10 mg MgClz, and 10 units each of RNase A and DNase were added. The solutions were kept at room temperature for 2 hours. They were then dialyzed for 4-5 hours, changing the water twice during that time, lyophilized and the resulting solids were weighed. These polymers were analyzed by NMR spectroscopy and by gas chromatography-mass spectrometry (GC-MS) after converting them to alditol acetate derivatives [5]. The remaining aqueous alcoholic solution was centrifuged to remove any precipitated solids. The supematants which contained free amino acids, carbohydrates, and the like were removed and analyzed by NMR spectroscopy and a sample subjected to amino acid analysis on a Waters Millipore system which included a WISP 7103 autosampler, a Waters 510 pump, and a Waters 440 absorbance detector. Fatty acid methyl ester analysis Total cells from cultures grown at 4°C and 24°C were methanolysed in 2% HCl in methanol at 75°C for 30 hours. The suspensions were blown to dryness, and 1 ml H20 and 2 ml hexane were added. After vigorous shaking, the mixture was centrifuged and the hexane layer was removed and concentrated to dryness. The resulting fatty acid methyl esters were analyzed by gas chromatography (GC) and GC-MS using a 30 m DB1 column. The temperature program was a ramp from 160°C to 320°C at 3°/min. The final temperature of 320°C was held for 20 min. Lipid headgroup analysis Approximately 5 mg of total lipids (isolated as described above) from cells cultured at 4°C and 24°C were dried, and then 0.5 ml of 2 M trifluroacetic acid was added. The 116 lipids were hydrolyzed at 120°C for 1.5 hrs. The lipid headgroups were obtained by adding 0.5 ml of H20 and extracting twice with 2 volumes of CHC13. The chloroform layers were removed and combined. The aqueous layers, containing the lipid headgroups, were dried and analyzed by NMR spectroscopy. Protein analysis by SDS-PAGE Total cell lysates were prepared from 4°C and 24°C cultures by adding two volumes of 10 mM Tris-HCl, pH 8.0 and 3 mg/ml lysozyme and incubating for one hour on ice. One volume of 2X SDS—PAGE sample buffer solution [0.125 M Tris-HCL (pH 6.8), 0.04% sodium dodecylsulphate (SDS), 10% glycerol, 5% B-mercaptoethanol, 0.0025% bromophenol blue] was then added. The samples were mixed well, boiled for 5 minutes, and centrifuged briefly to remove any precipitate. The lysates from 4°C and 24°C cultures were separated using Laemmli’s discontinuous buffer system [6] on 12% polyacrylamide slab gels containing 0.32% bisacrylarnide, 0.375 M Tris-HCl (pH 8.8), and 0.1% SDS. The stacking gel contained 4% polyacrylamide, 0.1% bisacrylarnide, 0.125 M Tris-HCI (pH 6.8), and 0.1% SDS. Electrophoresis was carried out at 80 V for about 3 hours. Gels were stained in a solution of 0.1% Coomassie blue in 40% methanol/10% acetic acid, and destained in a solution of 40% methanol/10% acetic acid. Electron microscopy Arctic strain 45-3 cells from 4°C and 24°C cultures were negatively stained with 1.5% potassium phosphotungstate and viewed by transmission electron microscopy. 117 RESULTS The morphologies of the cells cultured at the two temperatures were very different. The cells that were cultured at 4°C were waxy and oily. These cells would not form a suspension even after prolonged agitation. In contrast, the cells cultured at 24°C were fluffy and readily suspendable. This indicated that there was a significant layer of polysaccharide material on their surface and that the 4°C cells did not have much of a polysaccharide capsule. This was confirmed by the weight of capsular polysaccharide material obtained from the cells. Even though the packed volume of the 4°C cells was over twice that of the 24°C cells, the yield of polysaccharide in the latter cells was 2.4 times greater than from the 4°C cells. The cells grown at 4°C were also much more susceptible to osmotic shock in a hypotonic medium compared to those grown at 24°C. Hence these cells were completely lysed on storing at 10°C in distilled water in less than 4 hours whereas the cells cultured at 24°C were still intact after 2 days under the same conditions. Transmission electron microscopy revealed that there was a dramatic difference in cell sizes for bacteria grown at the two temperatures. Those grown at the higher temperature had twice the length and more than twice the diameter of the cells grown at the lower temperature. Since bacteria were rods at both temperatures, this represented a difference in volume of a factor of ~14 (Figure 4.2). Electron microscopy also indicated that the 4°C cells had only a microcapsule and had a very non-polar membrane surface since these took up the ionic stain only with difficulty. In agreement with these results, the hydrophobicity experiment showed that more cells from the 4°C culture (Figure 4.3A) attached to the 118 Figure 4.2: Electron micrographs (at the same magnification) of Arctic strain 45-3. Bar = 0.66 pm. (A) Cells cultured at 4°C. (B) Cells cultured at 24°C. Note that the 24°C are about two times the width and length of the cells grown at 4°C. _119 Figure 4.3: Fluorescent micrographs using RITC optics (at the same magnification) of Artic strain 45-3 stained with 0.1 % acradine orange attached to silanized coverslips. (A) Cells cultured at 4°C. (B) Cells cultured at 24°C. 120 silanized coverslips. than from the 24°C culture (Figure 4.3B), which indicated the 24°C cells have a more polar surface and were unable to attach to the coverslips. Proton N MR spectroscopy was used to provide a comprehensive determination of the relative amounts of the various metabolic intermediates elaborated by strain 45-3 at 4°C and 24°C. Analyses were carried out on the supernatant of the ethanol precipitation of the aqueous layer which would contain free amino acids, sugars, sugar alcohols, and nucleotides. Proton NMR spectra of these fractions from the 4°C (Figure 4.4A) and 24°C cells (Figure 4.48) indicated that the major components were amino acids. Two dimensional DQF-COSY (Figure 4.5) and TOCSY (data not shown) experiments showed that these were glutamate, valine, glycine, methionine, threonine, and a compound tentatively identified as 3- aminobutyric acid. These results were in agreement with amino acid analyses. From the NMR spectra, the relative proportions of the amino acids in each culture did not change significantly between cells cultured at the two temperatures, which was again confirmed by the amino acid analyses in which the relative amounts varied by about 1-4%. Signals for common osmolytes such as free sugars and sugar alcohols such as glycerol, sorbitol and mannitol, which have proton shifts in the 3.0 - 4.5 ppm region were noticeably absent or at most present in only trace amounts (Figure 4.6). Signals for the methyl group of glycine betaine (~3.2 ppm), another common osmolyte were also absent. Gas chromatography mass spectrometry analysis of the capsular polysaccharides obtained from the bacteria grown at 4°C and 24°C indicated that they contained fucose, ribose, mannose, galactose, glucose, and glucosamine (Figures 4.7A and B). There were 121 Figure 4.4: 1H NMR spectra of the cytoplasmic metabolites. (A) Cells cultured at 4°C. (B) Cells cultured at 24°C. The peaks are labeled with the one letter amino acid codes. The major components were glutamate (E), valine (V), glycine (G), methionine (M), threonine (T), and a compound tentatively identified as 3-aminobutyric acid (see also figure 5.5). Note that there are no significant differences in the relative pr0portions of amino acid components. 122 T Y Y I 7 1* f v r T r r v y I r r Y I Y r v v I v r v v I v I v r I v I v I I v 1*! '.-_' \ Figure 4.4 123 .880 Boa 255 5:2 one 05 55 @223 8a 33m $80 2:. .009 “a con—5.3 £8 59: mom—onus... flame—98.3 2: «e 55.58% VmOUd—On— Enema—35:. 95.“. "me 953% 124 m... 23E v a 3“). n v ’ TV. .If i! .p . v =2- .9. V was: . 8.. ..m. . as .... v a u. 1 a; V I" fabuwh. u. .n ...... “aile....3. 1&3... 31......5... u..?.«.........»t. 1.9.3... 18. .57 .I .o’ar’ uéofibsn ..o\\\.... ‘r r-Y—r—r‘r-r—i‘v Fr"! 1 r1‘1”'1'r"1"1 r r In In. .ro. 1m. I I O in O O 125 Figure 4.6: 1H NMR spectra of some common osmolytes. (A) glucose (B) glycerol (C) mannitol (D) sorbitol. 126 V I y—r f+1 Y Y Y —4 —1 4 O 7 6 5 C _" {fi- 1 Y fT v Y V r 1' v v Y 7 r v Y 1‘1 1 Y 4*,“ B l f' T I V r T r I V Y Y Y I V fl r I Y Y TfT ‘ l 5 3 2 pp. ... .- __ u L ll A....IL . u ._ 4].-.. —w . -w— - v—fi 7 -—~ ._..__' “V7 v v —~——— v—vw v.- v ‘ v v f 1 l v v v—r 1 v Y T—Y 5 I 3 I V V ‘7 Y I ff V Tfi ' Vfi'fi I V Y I ' 1' V f Y 1 I ‘ 2 1 PP Figure 4.6 127 Figure 4.7: Gas chromatograph profiles of the capsular polysaccharides from the bacteria cultured at (A) 4°C and (B) 24°C. The peaks indicated are ( 1) fucose, (2) ribose, (3) mannose, (4) galactose, (5) glucose, and (6) glucosamine. (Note: The peak seen at 58 min. was due to a transient shift in the baseline.) 2.5E6 2.8[8 1.5EB .BEB 5.0ES Abundance 1.8E8 .BES .BES .BES .QES Abundance m A 07 CD 128 3 A 4 2 l ,1J.i ,. 1. ,. .L,- . .. 10 20 3G 49 SB 68 Time(min.) 3 B 2 5 l 4 J M- r .. . 10 20 30 4B 50 80 7a Time(min.) Figure 4.7 k...’ 129 quantitative differences in the relative proportions between the cells from the two different cultures. Mannose was the major component in both cases. Ribose was the next major component in the case of the cells grown at higher temperature whereas galactose and glucose were the next major components in the cells grown at lower temperature. Another major difference was that the polymer from the 4°C cells contained a greater amount of glucosamine. Thin layer chromatography of the total lipids from the 4°C and 24°C cells showed three major components namely a fast moving component which was tentatively identified as triacylglycerol, diacylglycerol and free fatty acids. The latter two had retention times consistent with those observed for standards run under the same conditions. The levels of triacylglycerol were much higher in the cells grown at 4°C and appeared as a methanol insoluble oil that was soluble in chloroform or hexane. In the case of the 4°C cells, triacylglycerols accounted for more than half the total lipids. Diacylglycerol was the next predominant component, while only traces of phosphoglycerol were present. Other minor components were observed in the lipid preparations from the 24°C cells but the quantities were too small to allow characterization. Triacylglycerol probably serves as an energy source and not as a membrane component. Its presence, though, precluded a meaningful estimate of the total amount of membrane lipids obtained from the 4°C cells. Triacylglycerol is only sparingly soluble in methanol, however, and this was used as a way of separating it from the total lipids for NMR analyses. The recovery of lipids (excluding triacylglycerol) from the 4°C cells was ~6 mg per gram (wet weight) and about half this amount for the 24°C cells. 130 The 'H NMR spectra of the total lipids from the 4°C and 24°C cells are shown in Figure 4.8A and 4.8B, respectively. The two dimensional DQF-COSY and TOCSY experiments (not shown) indicated that the signals between 5.30 and 5.43 ppm were due to vinyl protons because of correlations with those at ~2.05 ppm. These were much more abundant in the lipids from the 4°C culture than from the 24°C culture. Connectivities between signals assignable to diacylglycerol and what later proved to be triacylglycerol were also evident. In the spectrum of the 24°C lipids, extra signals due to the glycerol group of phosphatidyl glycerol at 3.71 and 4.12 ppm were also observed. In agreement with the TLC analyses and the 1H NMR spectra of the total lipids, the 1H NMR spectra of the isolated lipid headgroups of the 4°C cells (after precipitating triacyl glycerol) showed definitively that the only lipid headgroup was glycerol (Figure 4.9A). The two doublet of doublets at 3.47 ppm and 3.57 ppm were assigned to the C1 and C3 protons of glycerol. The multiplet at 3.7 ppm was assigned to the proton on carbon 2 of glycerol. The lipid headgroup components of the 24°C cells were glycerol and phosphoglycerol (Figure 4.9B). The chemical shifts of the glycerol signals were the same as mentioned above. Two doublet of doublets at 3.65 ppm and 3.76 ppm were assigned to the Cl and C3 protons of phosphoglycerol. The multiplet at 3.84 ppm was assigned to the proton on carbon 2 of phosphoglycerol. From these data the proportion of phosphoglycerol was determined to be 40% by integration of the peaks. This method is much more reliable than charring the TLC plates and integrating densitometer traces. 1. m 131 Figure 4.8: 1H NMR spectra of the total lipids extracted from the bacteria grown at (A) 4°C and (B) 24°C. Note the much higher relative proportion of vinyl protons relative to the head group protons in the 4°C cells. TAG = triacylglycerol, DAG = diacylglycerol, PG = phosphatidylglycerol. 132 [Eu—0| n—UII I ppm lfilmulomol O.3§5.-.‘+> 223 \ 4 o . x - j ocn, ........... I ---.§3¢h .( s‘ ‘7‘ 42 H 221<—°14— 24! ’ : l .n I 42 : I - rue—ma— s 2 231 «AZ—2194i 311' : """"°'""""~j-MJ ooooooooooooooooooooo J I m . I I l 87. 101. 115. 129. 143. .. 100 60‘ 4a. I t ' 1 23 11 Z : s 4 l i" I. ‘ . .' I . . 200 300 400 500 600 Figure 5.2 163 19a 5 R 8 . O M 1 . : BB. MeO n "3 1 i or V I e 66‘ 9 j 74 '3 . are n 43. 97 d 38 a b n ‘ 2 7 ’ : 281 97 143 j 1 1 1 3171 1-4 2 9 27 ’ J 3 8 7 4 7 . so me 150 290 259 see 350 493 450 M/Z Figure 5.3: The electron impact mass spectrum of a 23 carbon fatty acid methyl ester resulting from coupling of n-heptane to a hexadecanoic acid residue from a S. ventriculi lipid component at the 00-1 position. The molecular weight is 368. The inset shows some of the major fragmentation patterns. 164 show that tail-to-tail coupling of exogenous lipids occurs without regard to its headgroup, the addition of hexadecylpyridinium chloride was also demonstrated. The 1H NMR spectrum of the total fatty acid methyl esters (Figure 5.4A), which included free hexadecylpyridinium chloride, regular chain fatty acids and transmembrane fatty acids, indicated that some of the hexadecylpyridinium compound had been converted to a reduced pyridine system since there were new signals in the region of 6.9 - 7.2 ppm. The total fatty acid methyl esters were separated by thin layer chromatography. A family of transmembrane fatty acids (visible by UV detection), now separated from the free hexadecylpyridinium chloride and regular chain fatty acid methyl esters, gave a mixture in which the 1H NMR spectrum indicated that the hexadecylpyridinium was coupled to native membrane lipids (indicated by the complex splitting of the methyl signals at about 0.9 ppm), but the aromatic ring had been transformed to a 1-alkyl-4-pyridone (Figure 5.4B). This was indicated by the disappearance of signals for the pyridine nucleus and the appearance of a narrow doublet at 7.72 ppm (2-position protons) and a multiplet at 7.54 ppm (3-position protons). The signals between 6.9 and 7.2 ppm were hardly visible in the spectra of the total extract prior to TLC indicating that they probably arose by oxidation of a dihydropyridine during handling. This is not surprising since the chimeric fatty acids were formed under reducing and anaerobic conditions. The infrared spectrum contained strong resonances for the C-H stretch at 2956 cm", 2928 cm", and 2859 cm". The peak for the ester carbonyl stretch was observed at 1725 cm“1 (m), while the C=C stretch was observed at 1600 cm'I (w). Electrospray mass spectrometry also supported the presence of chimeric fatty acid methyl esters containing the 4-pyridone nucleus. To test the theory that the bacteria reduced the pyridine ring, with subsequent 165 Figure 5.4: 1H NMR spectra of (A) total fatty acid methyl esters from S. ventriculi cultured with hexadecylpyridinium chloride, and (B) isolated fatty acid methyl esters containing hexadecylpyridinim coupled to native S. ventriculi membrane lipids. The signals at 8.1 ppm, 8.45 ppm, and 9.25 ppm in (A) are due to the pyridine nucleus of free hexadecylpyridinium chloride. Note the new signals in (A) in the region of 6.9-7.2 ppm which indicated that some of the hexadecylpyridinium had been converted to a reduced pyridine system. In (B) note the essential disappearance of signals for the reduced pyridine nucleus between 6.9 and 7.2 ppm, and the appearance of the narrow doublet at 7.72 ppm (2-position protons) and a multiplet at 7.54 ppm (3-position protons) indicate that during handling the aromatic ring was oxidized to a 1-alkyl-4-pyridone. 166 .rvrr vaIVV’yIv vvlvrwytyyvutrvvvy 7‘vlvavTvvyvlvvvarvryTvvrvl 7‘— u 10 9 O 1 6 s C 3 2 1 0 fi- Figure 5.4 167 oxidation to form the pyridone (probably during handling) (Figure 5.5), we reduced hexadecylpyridinium chloride using sodium borohydride to form a dihydropyridine, and then attempted to oxidize it to a pyridone using ceric ammonium nitrate (Figure 5.5). The reduction step was very successful but oxidation yielded a variety of products and 1H NMR spectroscopy indicated that only trace amounts of product corresponded to the desired material. In a prior study, it was demonstrated that Sarcina ventriculi carries out tail-to-tail coupling of its membrane lipids in response to agents lowering the pH [1]. It might be argued that the coupling response might have some ionic component which is acid catalyzed. An explanation that appears to be more consistent with the profile of this response, however, is simply that lowering of pH should suppress the ionization of the phosphate groups in the membrane lipids thus reducing their electrostatic cross-linking with calcium ions thus decreasing membrane stability. Such an argument would require that we also see coupling at high pH, since then the phosphate groups would be completely ionized. The resulting electrostatic repulsion should also destabilize the membrane thus triggering the process. Observing coupling at high pH would also rule out any acid catalysis. The gas chromatography profile of fatty acid methyl esters from Sarcina ventriculi cells cultured at pH 9.7 is shown in Figure 5.6. A family of peaks with very long retention times is seen in the pH 9.7 chromatograms. The inset shows a mass spectrum of peak 3, which corresponds to a C34,,oz,w-dicarboxylic acid dimethyl ester formed at high pH. This was the case for all of the other late eluting peaks in the profile. They were all readily identifiable as very long 168 fly an o {J NAD(P)H [O] —" —* l l \’V N N N BH4' H H O Ceric ammo- nium nitrate several Other I I —> + products N N ( Q15 (<) 15 Figure 5.5: Bacterial (proposed) and synthetic mechanisms of the reduction and subsequent oxidation of the pyridine ring of hexadecylpyridinium chloride. 169 £2.82 0882808 86250 no 88: 920 m .08 3:0 a w8880 .3 808.5. 8 £023 .m 80m .8 88.00% 888 .0mm8_ 80800—0 8. £508 .02: 0:... 88.00 35086 025.88.08.38 3.0 A3 .8050 35086 208 0.3888063... .30 Amv .8080 3508:. E08 0.388.806-38 SJ .8 .8080 3508388. -3 3.0 A: .8050 3508 20.. 02.08880 3:0 5. 80501508 20... 02.98280 3:0 ADV 80.0.0 3508 28 28.88 3.0 a. .882. be. So 2. a. 8:08.88 8.88 05. .5. In 8 08:50 8 $30.58.. .m. .8 $8.58 Ba: 88. 89a 3882...... 8080 .288 E8 .38.. .8 888388980 80 5% 08m?— 170 8 2:5 .Al as: :2820: we .4! N...‘ "1 "“‘1 J U N\: 00... 0mm corn r - aw? Gun SN 00.— . 8.5 “.13.: j. a n v N u v , 6.0. m 0 u m N n N h N . C v m 0 .av n 0 A. ._ . c . .uw u 8 . > . no . :8 L0 f a C {007%}; km.- ..8 _ u «.8 mm 0 m. Nflm .Oou 10109100 esuodso; 171 chain bifunctional species observed when these bacteria are cultured at low pH, high temperature or in the presence of organic solvents again demonstrating the universality of the response. To further test the hypothesis that tail—to-tail coupling is triggered by changes in the physical state of the membrane, Sarcina ventriculi was cultured in D20. The stronger hydrogen bonds in D20 should slow down the lipid molecular motions, which would be equivalent to a reduction in temperature. The proportion of transmembrane lipids should differ between cells cultured in H20 versus those cultured in D20. In these experiments, the pH was allowed to drop freely as a result of acetic acid production by the bacteria [8], usually reaching a minimal pH of 4.0. A reduction in pH is known to promote the formation of long chain lipids leading to a high proportion of transmembrane lipids. If the cells are cultured in D20, however, the reduced motional dynamics in the hydrogen bonding network in the lipid headgroup and in the various solvent cages should stabilize the membrane thus off- setting the fluidizing effects of the low pH. This is exactly what was observed (Figure 5.7). A high proportion of transmembrane fatty acids is seen in the cells grown in H20. In contrast, a very small proportion is observed in the cells cultured in D20. These results are consistent with the idea that substituting D20 for H20 is equivalent to a reduction in temperature causing an increase in membrane viscosity. The cells respond by maintaining a low level of long chain lipids. A similar kinetic effect of D20 on lipid composition has recently been reported [9]. In this study Bacillus subtilis was grown in D20 and the proportion of lipids with alkyl chain length of 16 carbons increased relative to those of 18 172 Figure 5.7: Gas chromatographs of fatty acid methyl esters from S. ventriculi cells cultured in (A) H20 and (B) D20. Note the high proportions of late eluting peaks in A compared to B. 173 Retention time Retention time Figure 5.7 174 or 20 carbons. The proposed mechanism for tail-to-tail coupling of membrane lipids is that hydrogens are removed from the w-l positions of two opposing alkyl chains forming radicals which quickly combine (Figure 5.1). Since these bifunctional fatty acid species are chiral, they should not be readily accessible to radical traps or scavengers. Radicals planarize rapidly, much faster than the timescale for diffusion from the cage in which they are formed [10-12]. This notwithstanding, it was of interest to assess the effects of various radical traps and scavengers on this system. Such experiments were important from another perspective. It was clear from previous experiments that oxygen had an adverse effect on the coupling activity. This would indicate either a radical process or the presence of an oxygen labile species such an iron-sulfur complex. Reagents which would allow us to test these possibilities were added to Sarcina cultures growing at 37°C, then the temperature was raised to 50°C to trigger the coupling reaction. The use of a radical trap is a very direct way of testing whether any accessible radical species are generated in the coupling reaction. This was done using 2-nitroso-2-methylpropane [13]. In these experiments, transmembrane cross- linking was not inhibited but, in fact, increased to levels (19%) much higher than observed in the 50°C control (Figure 5.8). Thiols too react very rapidly with radicals and would also be good probes to see if, during the coupling process, any radicals which might be generated are accessible. Two thiols, t-butyl thiol and n-butane thiol, were used and contrary to the nitroso methylpropane results, both inhibited coupling (Figure 5 .8). Iron-sulfur species are very sensitive to thiols [14-17] and this inhibition could be indicative of the presence of such 175 30- l % Coupling Relative to 50 c Control -40 _- 2-NMP t butyl tthI n-butane thiol 1 ,4-cyclohexadiene cyclohexane Figure 5.8: Results from radical trap/scavenger experiments indicating the percent of tail-to-tail coupling of S. ventriculi membrane lipids normalized to the 50°C controls. 2-Nitroso—2-methylpropane and 1,4-cyclohexadiene enhanced coupling, while the two thiols and cyclohexane inhibited the process. 176 a center and not the presence of an accessible radical center. To further clarify this point, 1,4-cyclohexadiene was used as a radical scavenger. This is a radical-sensitive diene that should have no effect on iron—sulfur centers. A similar result to the one seen with 2-nitroso- 2-methylpropane was obtained. The levels of cross linking actually rose (Figure 5.8). It appeared that 2-nitroso-2-methylpropane and 1,4-cyclohexadiene were not reacting with radicals but were further fluidizing membrane. 1,4-Cyclohexadiene is flattened and will not stack well between the membrane lipid chains. Cyclohexane, on the other hand, has a chair conformation and will stack well between the lipid chains. Since it would reduce the alkyl chain mobility, it could be expected to reduce the level of coupling considerably. This is exactly what was observed (Figure 5.8). The reduction of alkyl chain mobility by cyclohexane has also been characterized by fluorescence spectroscopy [18]. The relative effects of 1,4-cyclohexadiene and cyclohexane on hydrocarbon chain dynamics was investigated using NMR spectroscopy. The test system used was dipalmitoyl phosphatidylcholine. This is a comparatively rigid molecular system and the signals for the methyl and methylene groups at ~ 0.7 and 1.2 ppm are broad with very low intensities because of very short spin spin and spin lattice relaxation times. In the NMR (Figure 5.9) cyclohexane (spectra B and C) does not increase the fluidity of the lipids (rigidifies) as compared to the control (spectrum A). The methylene and terminal methyl peaks at 1.19 ppm and 0.72 ppm are still barely visible in all three spectra indicating no increase in T 1 or T2. 1,4-Cyclohexadiene, however, shows a marked affect on the lipids. Both the methylene and terminal methyl groups show increased mobility as seen by the increase in intensity of 177 Figure 5.9: 1H NMR spectra showing the effect of cyclohexane and 1,4-cyclohexadiene on the lipid packing arrangement of dipalmitoyl phosphatidylcholine. (A) dipalmitoyl phosphatidylcholine control, (B) 5g] cyclohexane, (C) lOul cyclohexane, (D) 5 ul 1,4- cyclohexadiene, an (E) 10 ul 1,4-cyclohexadiene. Signals for the methylene protons (1.19 ppm) and methyl groups (0.72 ppm) are indicated. 178 L L (E L L Figure 5.9 179 the peaks at 1.19 ppm for the methylene (seen in both spectra D and E) and 0.72 ppm for the methyl groups (which is more prominent in spectrum B), respectively. All of the results obtained here, therefore, indicate the putative hydrocarbon radicals are caged and therefore inaccessible to any external agents. It is clear that thiols are inhibiting cross-linking via another mechanism and not by radical scavenging. One possibility is that they are extruding iron-sulfur centers from the necessary cofactors such as ferredoxin [17], thus breaking the electron transfer chain and stopping the recycling of the hydrogen acceptor. The involvement of iron-sulfur centers is also consistent with the oxygen lability of the coupling process. The hydrocarbon chain coupling phenomenon demonstrated has no known precedent. The coupling conditions are mild and the regio- and stereospecificity is extremely high. It is independent of the nature of the functional groups on the distal ends of the carbon chains. Other systems containing this activity will, for sure, be identified and knowing how to optimize and control it will have important implications for synthetic organic chemistry. The stereoselective formation of carbon-carbon bonds between unactivated centers is viewed as the last frontier in carbon-carbon bond formation. ACKNOWLEDGMENTS I would like to acknowledge Seunho Jung, a past member of our laboratory, who did 180 the cell-free coupling of the heptadecanoic acid to S. ventriculi membrane fatty acids. I would also like to acknowledge Luc Bérubé’s, another past member of our laboratory, contributions to this work as well. He did the alkaline pH and D20 experiments. He also did the experiments in which S. ventriculi was cultured in the presence of cyclohexane, 1,4- cyclohexadiene, and tert-butyl thiol. REFERENCES 1. Jung, S., S. E. Lowe, R. I. Hollingsworth, and J. G. Zeikus. (1993) Sarcina ventriculi synthesizes very long chain dicarboxylic acids in response to different forms of environmental stress. J. Biol. Chem., 268: 2828-2834. 2. Jung, 8., and R. I. Hollingsworth. (1994) Structures and stereochemistry of the very long cam-bifunctional alkyl species in the membrane of Sarcina ventriculi indicate that they are formed by tail-to-tail coupling of normal fatty acids. J. Lipid Res, 35: 1932-1945. 3. Jung, 8., and R. I. Hollingsworth. ( 1995) A mathematical model explaining the molecular weights and distribution of very long chain dicarboxylic acids formed during the adaptive response of Sarcina ventriculi. J. Theor. Biol, 172: 121-126. 4. Bérubé, L. R., and R. I. Hollingsworth. (1995) Membrane lipid alkyl chain motional dynamics is conserved in Sarcina ventriculi despite pH-induced adaptive structural modifications including alkyl chain tail to tail coupling. Biochemistry, 34: 12005-12011. 5. Sinensky, M. (1974) Homeoviscous adaptation--a homeostatic process that regulates the viscosity of membrane lipids in Escherichia coli. Proc. Nat. Acad. Sci. USA, 71: 522-525. 6. Eibl, H. (1983) The effect of the proton and of monovalent cations on membrane fluidity. In Membrane F luidity in Biology, ed. R. C. Aloia, 2: 217-236. 7. Glasoe, P. K., and F. A. Long. (1960) Use of glass electrodes to measure acidities in deuterium oxide. J. Phys. Chem, 64: 188-190. 8. Canale-Parola, E. (1970) Biology of the sugar-fermenting Sarcinae. Bacteriol. Rev., 34: 82-97. 181 9. Hollingsworth, R. I. (1995) Membrane lipid alkyl chain heterogeneity is a molecular transform of the available thermal energy spectrum. J. Phys. Chem, 99: 3406-3410. 10. Doering, W. v. E., M. Farber, M. Sprecher, K. B. Wiberg. (1952) On the decarbonmonoxylation reaction. J. Amer. Chem. Soc., 74: 3000-3001. 11. Greene, F. D., and N. N. Lowry. (1967) 9-Decalyl free radicals. Decomposition of cis- and trans-9-decalylcarbinyl hypochlorite. J. Org. Chem, 32: 875-882. 12. Walborsky, H. M., and P. C. Collins. (1976) Cyclopropanes. XXXVIII. Effect of 1- substituents on the stereochemical stability of the cyclopropyl radical. J. Org. Chem, 41: 940-946. 13. Thie’ry, C. J., JP. L. Agnel, C. M. Fre’javille, and J. J. Raffi. ( 1983) Electron spin resonance spin-trapping analysis of y-induced radicals in polycrystalline oc-D-glucose. J. Phys. Chem, 87: 4485—4488. 14. Bobrik, M. A., L. Que, Jr., and R. H. Holm. (1974) Synthetic analogs of the active sites of iron-sulfur proteins. IV. Ligand substitution reactions of the tetranuclear clusters [Fe4S4(SR)4]2' J. Amer. Chem. Soc., 96: 285-287. 15. Que, L., Jr., J. R. Anglin, M. A. Bobrik, A. Davison, and R. H. Holm. (1974a) Synthetic analogs of the active sites of iron-sulfur proteins. D(. Formation and some electronic and reactivity properties of Fe4S4 glycyl-L-cysteinylglycyl oligopeptide complexes obtained by ligand substitution reactions. J. Amer. Chem. Soc., 96: 6042-6048. 16. Que, L., Jr., M. A. Bobrik, J. A. Ibers, and R. H. Holm. (1974b) Synthetic analogs of the active sites of iron-sulfur proteins. VII. Ligand substitution reactions of the tetranuclear clusters [Fe4S4(SR)4]2' and the structure of [(CH3)4N}2 [Fq S4 (SQ l-g )]. J. Amer. Chem. Soc., 96: 4168-4178. 17. Que, L., Jr., R. H. Holm, and L. E. Mortenson. (1975) Extrusion of Fe2S2* and Fe4S4* Cores from the active sites of ferredoxin proteins. J. Amer. Chem. Soc., 97: 463-464. 18. Sikkema, J., J. A. M. de Bont, and B. Poolman. (1994) Interactions of cyclic hydrocarbons with biological membranes. J. Biol. Chem, 269: 8022-8028. CHAPTER VI SUMNIARY AND PERSPECTIVES 182 183 The ability of procaryotes to survive in, or adapt to, many extreme environments is impressive. The fact that they can not only withstand such extremes of temperatures, pH, pressure, salinity, and high concentrations of detrimental chemical agents, but thrive under these conditions is extraordinary. One of the most commonly observed cellular change, however, in response to all of the different environmental perturbations is alteration of membrane acyl (or alkyl) chain structures. As the cell envelope membrane(s) is in constant contact with the environment, it makes sense that rapid alterations in membrane structure are necessary as conditions change. We have shown in this work that the extent of membrane acyl chain length heterogeneity observed response to environmental perturbations is significant. The selection of the acyl chain length distribution is a random process, in which the mean and variance of the distribution curve depends on the available thermal energy spectrum. We know that many plants and animals also make similar changes in their membrane acyl chains in response to temperature as well, for instance fish adapt to cold temperatures by a similar spectrum of membrane chemistry changes. This, then, seems to be a general response to temperature changes. Thus aspects of the relationship between temperature and acyl chain length distributions are likely to be universal. From an ecology and paleontology perspective, the relationship between fatty acid chain length heterogeneity of a given organism and temperature could be used basically as a temperature stamp. The fatty acid compositions of bacteria isolated from different soil depths (or from other organisms like fish) that have been preserved in some fashion (whether or not they are still viable), like the 184 Siberian permafrost bacteria, could give indications of global climate changes or catastrophic events that have occurred in a particular region as the earth has evolved. The fact that there is also a temperature equivalent for solvent and antibiotics effects is another important finding. The possible relationship between temperature, solvent concentration, and growth rate could be used to extend the industrial utility of microorganisms. Using a three-dimensional graph like the one shown in chapter two, the external operation temperature for microorganisms could be extended in both directions, by adding the appropriate concentration of a fluidizing or rigidifying solvent. It may also be possible to overcome drug resistance of some bacteria by determining the appropriate combination of temperature and drugs to kill the bacteria. It has been shown, in fact, that several bacterial species show temperature dependent sensitivity to certain antibiotics. Several Pseudomonas strains were more resistant to gentamicin at 20°C and 30°C, but became more susceptible at 37°C and 42°C [1]. In another study, all Yersinia kristensenii strains were more resistant to B-lactam antibiotics at 22°C than at 37°C [2]. The reason for the increased resistance to antibiotics at lower temperatures is probably due to the increased rigidity of the membranes such that antibiotics are less effective in intercalating into the membrane and so there is less uptake of the drug. Recent evidence for this was shown in a study in which strains of Stenotrophomonas maltophilia that were more resistant to aminoglycoside antibiotics at lower temperatures all had more rigid membranes than strains that were susceptible at those temperatures as shown by ESR spectroscopy [3]. It may also be feasible to kill cancer cells using the right concentrations of chemotherapy drugs and 185 localized heating of the tumor with lasers or probes. Many researchers use the profile of fatty acid compositions of microorganisms as an indication of the species, but these organisms are typically cultured under different conditions in the laboratory than found in nature. However, we have shown that the balance between metabolic pathways, for instance carbohydrate versus amino acid metabolism, could have a significant impact on the membrane acyl chain structures. This balance is determined by the available heat (kinetic energy) needed to carry out chemical processes. At lower temperatures the amount of maintenance energy increases dramatically and cells need to use carbon sources that require less energy to catabolize and yet yield more energy for maintaining the necessary cellular functions. These shifts in metabolism directly affect the types of fatty acid chains produced by changing the available acyl-CoA primer pool for fatty acid synthesis. This finding also supports the randomness of membrane acyl structures, since the cell has to use the primers available, a random selection process would favor whatever is the most prominent primer in the pool. Therefore it would be more appropriate to demonstrate that organisms produce the same spectrum of fatty acids under different conditions. This is preferable to basing the relatedness of organisms on similar fatty acid profiles when the organisms were most likely cultured under two very different conditions. Also since it is expected that organisms that are adapted to subzero climates would most likely use triacylglycerol stores for their primary energy source at low temperatures, as is shown by large amounts of triacylglycerol found in Arctic strain 45-3 at 4°C, they should have lipases which are very active at low temperatures. Isolation of these enzymes would 186 have many uses in industry. Work in our laboratory is currently being done to characterize the activity of lipases in Arctic strain 45—3. The findings that Arctic strain 45-3 does not produce compatible metabolites in response to cold temperatures indicates that it is not clear that there is link between extreme salinity and freezing. Also the current models on the ability of compatible solutes to confer cryotolerance as well are certainly not general. As such more serious physico-chernical models and theories for explaining the interactions between water and biological surfaces need to be developed. The model we propose, in which cell surface alterations are sufficient to provide cryoprotection, should also be expanded and subjected to a more physico- chemical analysis in more systems. The advantages of carbon-carbon bond formation at unactivated centers by S. ventriculi has many potential synthetic organic chemistry and industrial applications. The coupling of fatty acids in this manner could be used to make biodegradable plastics, liquid- crystalline devices, and allow the grafting of materials onto polymers to help in the biocompatibility of artificial implants. While the entire functioning enzymatic system will probably not be able to be isolatable from the cells, since its regulation in the membrane is likely very complex, characterizing aspects of the process would aid in understanding ways of optimizing the system to effect the coupling of the desired lipid species. Initial work done in isolating potential enzymes involved in this coupling process is presented in Appendix A. 187 A lot of work in studying adaptative responses of microorganisms to environmental perturbations seems to result in "catalogs" of types of responses, the levels of the responses, and in which organism they occur. Since these fine details of the adaptations seem to vary so much in bacteria, little (if any) general conclusions are drawn from these observations. In this work, we have hopefully provided new insights and theories as to the fundamental principles involved in adapting to fluctuations in the environment. Bacteria, and even higher organisms, have certain cellular functions which must be maintained regardless of the environment. The manner in which organisms adapt to perturbations vary based on the types of cellular processes which they are able to carry out. However, all of these chemical processes rely on the available thermal energy (and thus kinetic energy) spectrum. There is a circular relationship, though, between this available energy, the balance of metabolic processes, and membrane structure. But, by "tuning" the membranes through continual alterations in membrane structures to propagate the maximum of this available energy, an optimal balance of metabolic pathways can be achieved to maintain necessary cellular functions. The process by which organisms adapt to other environmental perturbations, other than temperature, results from similar perturbations in the relationship of kinetic energy, metabolism, and membrane structure. These changes are again detected by disturbances in membrane phase equilibria, similar to temperature effects, which is why similar adaptative responses are observed. So organisms typically have a general set of changes they can make in response to many stress conditions, rather than a specific mechanism of adaptation for a given type of environmental change. Thus the idea of specific receptors, channels, and pumps for individual antibiotics or other chemical species is highly unlikely. Hopefully the 188 work presented here will elicit new views, ideas, and models as to the generality of adaptation mechanisms to environmental perturbations and how they are tied to the types of cellular processes than can be carried out by the cell. Also the application of the principles we have developed for the adaptation and adaptability of bacteria to extreme environments, along with the use of certain enzymatic functions of extreme adapted organisms has a great deal of potential use in research and industry. REFERENCES 1. Papapetropoulou, M., G. Rodopoulou, E. Giannoulaki, and P. Stergiopoulos. (1994) Effect of temperature on antimicrobial susceptibilities of Pseudomonas species isolated from drinking water. J. Chemother., 6: 404-407. 2. Bejar, V., C. Calvo, and A. Ramos-Cortnenzana. (1986) In vitro susceptibility of Yersinia kristensenii strains to B-lactam antibiotics. Ann. Inst. Pasteur. Microbiol., 137A: 169-177. 3. Rahmati—Bahram, A., J. T. Magee, and S. K. Jackson. (1995) Growth temperature- dependent variation of cell envelope lipids and antibiotic susceptibility in Stenotrophomonas (Xanthomonas) maltophilia. J. Antimicrob. Chemother., 36: 317-326. APPENDIX A PREPARATION OF THE RADIOLABELED PHOTOAFFINITY PROBE FOR IDENTIFYING PUTATIVE CROSS-LINKING ACTIVITY 189 190 We have begun work to try to isolate the enzymes involved in the tail-to-tail fatty acid cross-linking mechanism in Sarcina ventriculi characterized in chapter 5. Below is described the synthesis of a membrane l4C-labeled photoaffinity probe, which was then incorporated into growing cultures of Sarcina ventriculi. Preliminary results of this experiment are then shown. Synthesis of a 1“C-labeled photoaffinity probe A l4C-labeled photoaffinity probe was synthesized as outline in Figure A. l. (1) The probe was made by combining 50 mg of the C,5-iodofatty acid with 40 pl (40 uCi) ”C- phenol and 20 mg cold phenol (final concentration of 1 uCi/mg phenol), plus 50 mg anhydrous potassium carbonate and 0.5 ml acetone. The reaction was stirred overnight. The reaction was stopped by adding 1 ml of water and glacial acetic acid until effervesence ceased. The product was extracted with chloroform and dried under a stream of nitrogen. The product was dissolved in 2 ml acetic acid and the solution cooled in ice while stirring. One milliliter of concentrated sulfuric acid was then added followed by the slow addition of 100 u] fuming nitric acid. The reaction became very dark orange/brown in color. After about 10 minutes it was diluted with about 10 g of crushed ice, and then 15 ml chloroform was added. The mixture was stirred, and the chloroform layer was then removed, washed with cold water, and evaporated to dryness. The nitrated product was then hydrogenated to form the amino derivative. This was done by taking up the product in 50 ml of absolute ethanol and adding a few drops of water and acetic acid. Ten percent of palladium on carbon (~20 mg) was added while passing nitrogen over the reaction. Hydrogen was supplied by L 191 attaching two hydrogen filled balloons to the system. The reaction mixture was allowed to stir for over 8 hours until the yellow color had disappeared. The palladium on carbon was removed by vacuum filtration through methanol washed celite. The product mixture was evaporated to dryness. The product was dissolved in 10 ml of 2 M sulfuric acid and 10 ml of absolute ethanol and cooled in a salt/ice water bath. Sodium nitrite (0.2 g) was slowly added. The reaction then continued to stir for 2 hours more. From this step on, the reactions were protected from light. Sodium azide (0.2 g) was slowly added, and then the reaction was stirred for one hour at room temperature, followed by a half hour at 60°C. The product mixture was evaporated only long enough to remove the ethanol, leaving the product in about 5 ml of water. The l4C-photoaffinity probe, containing the azide groups, were extracted three times with ethyl acetate and evaporated to dryness. The product was taken up in 5 ml of ethyl acetate and 50 ul was counted on the scintillation counter. The product was then dried again and stored at -20°C. The specific activity of the photoaffinity probe was about 55 mCi/mole. Growth of Sarcina ventriculi with the l"C-photoaffinity probe The ”C-photoaffinity probe was dissolved in ethanol and half of it was split between two exponentially growing cultures of Sarcina ventriculi (50 ml) at 37°C. The final ethanol concentration per culture was 0.2 M, and the cultures were wrapped in aluminum foil. Continued 37°C incubation for 6 hours. Cells were protected from light at all times. Cells were harvested by centrifugation, and washed with Dulbecco's phosphate buffered saline (10 mM phosphate, 0.15 M NaCl, pH 7.0) containing 0.9 mM CaClz, 2.7 mM KCl, and 0.5 mM 192 MgClZ. Cells were lysed by passage through a French pressure cell (American Instruments Co., Inc., Silver Spring, MD) and by freeze/thawing several times. The probe was then cross-linked membrane components by exposing the lysate (in a quartz cuvette) to a high intensity UV lamp for two minutes. Isolation of potential protein involved in the cross-linking mechanism Half of the l4C-lysate (1 ml) was separated by using a Sephadex G-75 column. The amount of radioactivity for the first 100 fractions was determined. Fractions containing the most radioactivity were lyophilized. SDS-PAGE gels and Western transfers The radioactive Sephadex G-75 fractions were lyophilized and separated on duplicate 12% SDS-PAGE gels using Laemmli’s discontinuous buffer system [1] on 12% polyacrylamide slab mini-gels containing 0.32% bisacrylamide, 0.375 M Tris-HCl (pH 8.8), and 0.1% SDS. The stacking gel contained 4% polyacrylamide, O. 1% bisacrylamide, 0.125 M Tris-HCl (pH 6.8), and 0.1% SDS. Electrophoresis was carried out at 200 V for about 1 hour. One gel was stained in a solution ofO. 1% Coomassie blue in 40% methanol/10% acetic acid, and destained in a solution of 40% methanol/10% acetic acid. The other gel was Western transferred to ProblottTM (Applied Biosystems) nylon membrane, which is suitable for protein sequencing, following manufacturer instructions: The nylon membrane was briefly wetted with methanol and then soaked in electroblotting buffer (10 mM 3- [cyclohexylamino]-l-propanesulfonic acid {CAPS}, pH 11 in 10% methanol). The gel was 193 soaked in electroblotting buffer for 5 minutes. The transblotting sandwich was assembled according to Bio-Rad manufacturer's instructions, and the transfer was carried out at 100V on ice for 1.5 hours. The nylon membrane was Coomassie stained with 0.1% Coomassie Blue R-250 in 40% methanol/1% acetic acid for less than one minute and then destained with 50% methanol and rinsed well with deionized water. The membrane was air-dried and then exposed to a phosphoimager. PRELINIINARY RESULTS Currently, work is being done to isolate components of the enzymatic system for tail- to-tail coupling of membrane fatty acids. A l4C-labeled fatty acid photoaffinity probe was synthesized (Figure A. 1) such that the reactive azide group would be positioned in the middle of the membrane bilayer. After growing S. ventriculi in the presence of this probe, the probe was cross-linked using UV light to membrane components. Since the cross-linking occurs in the center of the membrane, it is reasonable to expect that enzymes involved in this process must span the central portion of the membrane. Phosphoimages of Western blots containing proteins from the Sephadex column fractions containing the highest amounts of radioactivity show that only three of the void volumn fractions contain a detectable labeled protein (Figure A2). The other fractions may also contain labeled proteins, but either not enough protein was loaded onto the gel, or there is too little radioactivity to detect. In these three void volumn fractions, a single 15 kDa protein is detected. Since the fractionation range of Sephadex G-75 is 3 - 80 kDa, this protein was likely a large aggregate as it passed 194 OH OH OH 0H 0 on 0 _ 0H + 12:3; H2/Pd Name/Ho NaNa l4C-Iabelled. and unlabelled NO; NH: m‘c1 :4, Figure A.1: Synthesis of the l“C-labeled photoaffinity probe. Figure A.2: Phosphoimage of the nylon membrane containing proteins from several G- 75 column fractions. Only lanes containing proteins from column fractions 21, 22, and 23 had a radioactively labeled protein detectable by phophoimaging. The other lanes probably did not contain enough protein and/or enough radioactivity to detect. 195 through the column. This would be expected as integral membrane proteins are usually hydrophobic and thus they would tend to aggregate. This protein has been submitted for sequencing, though no results have been obtained to date. REFERENCES 1. Laemmli, U. K. (1970) Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature, 227 : 680—685. "‘illlillllllllllES