Thea“. mcmom STATE UNIVERSITY BRARIES l \ llllllllllllllll lllllllllllll l H 3 1293 01421 4930 l This is to certify that the dissertation entitled presented by Douglas Scott Burdette has been accepted towards fulfillment of the requirements for Ph. D . degree in Biochemistrx Datelgrch 19 . 1996 M5 U is an Affirmative Action/Equal Opportunity Institution 0-12771 — — EMF—gs _ LIBRARY Michigan State University PLACE ll RETURN BOX to remove thle checkout from your record. TO AVOID F INES return on or betore date due. DATE DUE DATE DUE DATE DUE -L_l::l Sicc THERMOANAEROBACTER ETHANOLICUS 39E SECONDARY-ALCOHOL DEHYDROGENASE: MOLECULAR BASIS FOR STABILITY AND CATALYSIS By Douglas S. Burdette A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Biochemistry 1996 I. Ala: 5- d“ u .J'NL ABSTRACT THERMOANAEROBACTER ETHANOLICUS 3913 SECONDARY- ALCOHOL DEHYDROGENASE: MOLECULAR BASIS FOR STABILITY AND CATALYSIS By Douglas S. Burdette The ath gene encoding Themtoanaerobacter ethanolicus 39E secondary-alcohol dehydrogenase (2° ADH) was cloned, sequenced, and overexpressed in Escherichia coli DHSa (~15% total prot.). The protein was purified to homogeneity by heating and precipitation. The 1056 bp gene encoded a homotetrameric recombinant enzyme (37.7 kDa subunits) that displayed a l70—fold greater catalytic efficiency toward NADP(H)—dependent propan-2—ol than toward ethanol oxidation. The 2° ADH site directed mutants C378, H59N, DlSON, D150E, and DISOC displayed 5 3% of wild type catalytic activity. These data and the wild type enzyme inactivation by dithionitmbenzoate (DTNB) and diethylpyrocarbonate chemical modification, supported sequence predictions based on Hydrophobic Cluster Analysis (HCA) that Cys37, HisS9, and AsplSO residues are catalytic Zn ligands. X-ray absorption spectrometry data supported the presence of a protein bound Zn with a ZnSl(N-O)3.4 coordination sphere. Induction coupled plasma emission spectrometry data of wild type and mutant enzyme Zn binding, implicated these residues in 2° ADH Zn liganding. Thus, a catalytic Zn atom and its immediate environment are important in 2° ADH activity. Analysis of the Gl98D mutant supported the HCA based prediction that this 2° ADH binds NADP(H) in a common nucleotide binding motif, a Rossmann fold. The substrate and cofactor binding sites in the 2° ADH are structurally distinct because catalytic Zn ligand mutations did not change cofactor affinity. The thermophilic 2° ADH) was optimally active at ~90°C and displayed a half-life of 1.2 days at 80°C. Analysis of temperature dependent unfolding of 2° ADH in guanidine hydrochloride (GuHCl) indicated the enzyme was very rigid (50% unfolded at ~115°C). Thus, the enzyme is nearly 100% folded at 90°C where it is maximally active. 2° ADH Activity increased 2-fold at 37°C to 75°C in low concentrations of GuHCl (120190 mM); and, activity remained in 1.6M GuHCl demonstrating its high resistance to chemical denaturants. Catalytic rate enhancement was also seen with weakly chaotropic (i.e., KNO3) and neutral (i.e., KCl) inorganic salts, indicating that specific ionic and not hydrophobic interactions are required for the catalytic rate enhancement. Linear Arrhenius plots for the oxidation of propan-Z-ol by the native and recombinant 2° ADHs from 30°C to 90°C suggested that the thermal activity relationships are related to binding high kinetic energy substrates and not to temperature dependent changes in enzyme unfolding. Consequently it is suggested that the 2° ADH has evolved high rigidity for stability and these molecular determinants are distinct from those which control catalysis. Copyn'ght by DOUGLAS scorr BURDBTI'E 1996 To Kate and to our families ACKNOWLEDGEMENTS I wish to thank my graduate advisor, Dr. Zeikus, and my committee members for their guidance and patience. I would also like to thank Dr. Robert Phillips, Dr. Robert Scott, and Dr. Bobby Ami whose collaboration on this project has provided greater depth to the research. I am specifically indebted to Dr. Claire Vieille, Maris Laivenieks, Cindy Petersen, and‘the rest of the Zeikus lab members for scientific guidance, technical assistance, and friendship. I also thank Chris Jambor, the Hupes, the Santoros, the Giegels, the Bordeners, the Castigliones, the Noons, and my family for giving me the strength to endure. I am especially grateful for the consistent support of my grandparents, specifically to Dora and Eugene Carbon whose kindness will live in my memory. Finally, I thank Dr. Kate Noon who inspired me daily to persevere. TABLE OF CONTENTS page LIST OF TABLES .............................................................................. ix LIST OF FIGURES ............................................................................ xi LIST OF ABBREVIATIONS ................................................................. xiii CHAPTER I Literature Review ............................................................................... 1 ALCOHOL DEHYDROGENASES Alcohol dehydrogenase structure and function ..................................... 2 Alcohol dehydrogenase industrial applications ..................................... 5 THERMOPHILIC ENZYMES Thermophilic enzyme sources and diversity ........................................ 8 Thermostability ......................................................................... 21 Thermophilicity ......................................................................... 23 Protein folding .......................................................................... 30 Protein unfolding ....................................................................... 32 Molecular mechanisms of thermostability Intrinsic factors ..................................................................... 34 Multiple substitutions and protein stabilization ............................. 44 Substitutions and modification of the thermodynamics of unfolding 45 Prolines in loop regions ....................................................... 46 Salt bridges ..................................................................... 49 Hydrogen bonds ............................................................... 50 Hydrophobic interactions and core packing ................................ 51 Covalent destruction and irreversible denaturation ......................... 53 Extrinsic mechanisms .............................................................. 54 Glycosylation ................................................................... 56 Salts ............................................................................. 56 Other chemical effectors ....................................................... 58 Pressure effects ................................................................ 59 Molecular mechanisms of protein thermophilicity .................................. 59 Genetic engineering of therrnozymes Modification of enzyme catalytic properties ..................................... 62 Thermostability and thermophilicity engineering ............................... 62 THESIS OBJECTIVES AND SIGNIFICANCE ............................................ 68 ACKNOWLEDGEMENTS .................................................................... 71 REFERENCES .................................................................................. 72 CHAPTER II Cloning and Expression of the Gene Encoding the Thennoanaerobacter ethanolicus 39E Secondary-Alcohol Dehydrogenase and Enzyme Biochemical Characterization ................................................... 89 Abstract .................................................................................. 90 Introduction ............................................................................. 91 Materials and methods ................................................................. 92 Mt St: wwknw Dawn. an: MAW). Results ................................................................................... 96 Discussion ............................................................................... 112 Acknowledgements ..................................................................... 116 References ............................................................................... 117 CHAPTER 111 Effect of Thermal and Chemical Denaturants on Thennoanaerobacter ethanolicus 39E Secondary-Alcohol Dehydrogenase Stability 120 Abstract .................................................................................. 121 Introduction .............................................................................. 122 Materials and methods .................................................................. 124 Results ................................................................................... 128 Discussion ............................................................................... 144 Acknowledgements ..................................................................... 147 References ............................................................................... 148 CHAPTER IV Mutagenic and Biophysical Analysis of T hermoanaerobacter ethanolicus Secondary-Alcohol Dehydrogenase Activity ................................................. 151 Abstract .................................................................................. 152 Introduction ............................................................................. 154 Materials and methods ................................................................. 156 Results ................................................................................... 159 Discussion ............................................................................... 174 References ............................................................................... 185 CHAPTER V CONCLUSIONS ................................................................................. 187 CI-IAP'I'ER VI Directions for Future Research .................................................................. 192 Sequence based mutagenesis and kinetics ........................................... 193 Structural Analyses ..................................................................... 202 Biotechnological utility ................................................................. 204 References ............................................................................... 205 APPENDIX A Sequence Comparison Based Predictions of Important Catalytic Domain and Cofactor Binding Domain Amino Acids in 1° and 2° ADH Structures .................... 206 APPENDIX B Computer programs used to calculate (hyper)thermophilic versus mes0philic total amino acid composition and the theoretical Arrhenius data for thermophilic and mesophilic enzymes ........................................................................ 214 APPENDD( C Construction and Kinetic Characterization of Thennoanaerobacter ethwwlicus 39E 2° ADH Proline Deficient Mutants ........................................................... 239 Materials and methods ................................................................ 241 Results .................................................................................. 242 Discussion .............................................................................. 246 References .............................................................................. 248 Ci In To; LIST OF TABLES page CHAPTER I Table 1 - Thermozymes from thermophilic organisms ..................................... 10 Table 2 - Thermozymes from hyperthermophilic organisms .............................. 14 Table 3 - Amino acid composition comparison of enzymes from mesophilic, thermophilic, and hyperthermophilic prokaryotes ............................ 20 Table 4 - Kinetic constants for type II xylose isomerases purified from mesoPhiles and thermophiles .................................................... 28 Table 5 - Kinetic constants for glutamate dehydrogenases pmified from mesophiles and thermophiles .................................................... 28 Table 6 - Representative site directed mutagenesis studies of enzyme thermostability mechanisms ...................................................... 35 Table 7 - Comparison of xylose isomerase thermal parameters with the optimal growth temperature of the respective microorganism ......................... 55 CHAPTER II Table 1 - Comparison of primary structural similarity between horse liver and bacterial alcohol dehydrogenases ................................................ 102 Table 2 - Effect of enzyme preincubation temperature on the activation energy for propan-2-ol and ethanol oxidation by T. ethanolicus 39E 2° ADH ..... 111 CHAPTER III Table 1 - Role of protein precipitation in T. ethanolicus 2° ADH thermal inactivation ............................................................... 136 CHAPI'ER IV Table l - Oligonucleotide primers for PCR amplification of mutant ath gene DNA .......................................................... 157 Table 2 - Pmification of the recombinant wild type 2° ADH expressed from plasmid pADHBKA4-kan ..................................... 160 Table 3 - Effect of site specific amino acid substitutions on T. ethanolicus 2° ADH activity .................................................. 164 Table 4 - Effect of the Glyl98 to Asp mutation on T. ethanolicus 2° ADH nicotinamide cofactor preference ..................... 175 CHAPTER VI Table 1 - Comparison of conserved 2° ADH amino acids with the corresponding 1° ADH residues from HCA based sequence alignments ..................... 194 Table 2 - Comparison of conserved N ADP+ dependent 2° ADH amino acids with the corresponding NAD+ dependent 1° ADH residues deduced from HCA based sequence alignments ......................................... 198 Table 3 - Added thermophilic 2° ADH proline residues compared to the mesophilic 2° ADH deduced from HCA based sequence alignments ........ 199 APPENDIX A Table 1- Comparison of aligned amino acids involved in ADH catalytic domain structure and function ................................................... 209 Table 2- Comparison of aligned amino acids involved in ADH nicotinamide cofactor binding ................................................................... 211 LIST OF FIGURES page CHAPI'ER I Figure l - Arrhenius dependence of reaction rate on the temperature and activation energy of mesophilic and hyperthermophilic enzymes ............ 26 Figure 2 - Role of proline residues in protein structural stabilization .................... 48 Figure 3 - Flow chart of potential steps toward engineering enzyme thermophilicity and thermostability .............................................. 65 CHAPTER II Figure 1 - Restriction map of the T. ethanolicus 39E 2° ADH clone (pADHB25) ...... 99 Figure 2 - Nucleotide sequence and deduced amino acid sequence of T. ethanolicus 39E ath and of the downstream open reading frame ....... 101 Figure 3 - HCA comparison of thermophilic and mesophilic 1° and 2° ADHs centered on the putative nicotinamide binding motifs .......................... 104 Figure 4 - Arrhenius plots for the recombinant T. ethanolicus 39E 2° ADH between 25°C and 90°C ............................................................ 110 CHAPTER 111 Figure 1 - Recombinant T. ethanolicus 39E 2° ADH thermophilicity ..................... 130 Figure 2 - Recombinant T. ethanolicus 39E 2° ADH thermostability ..................... 132 Figure 3 - Temperature dependence of 2° ADH precipitation .............................. 135 Figure 4 - Effect of GuHCl on T. ethanolicus melting temperature ....................... 138 Figure 5 - Effect of GuHCl on T. ethanolicus fluorescence ............................... 141 Figure 6 - GuHCl and temperature dependence of 2° ADH activity ...................... 143 CHAPTER IV Figure 1 - SDS-PAGE of wild type recombinant 2° ADHs ................................ 163 Figure 2 - T. ethanolicus 2° ADH EXAFS analysis ........................................ 166 Figure 3 - PCR-based ath gene site directed mutagenesis scheme ..................... 168 Figtue 4 - SDS-PAGE of 2° ADH mutant proteins ......................................... 170 Figure 5 - Comparative 1° and 2° ADH active site models ................................. 181 CHAPTER VI Figure 1 - Position of the non-conservative amino acid differences between 1° and 2° ADHs relative to the horse liver 1° ADH catalytic site .............. APPENDD( A Figure 1 - HCA comparison of the B. stearothennophilus 1° ADH, the T. ethanolicus 2° ADI-I, and the C. beijerinckii 2° Adh amino acid sequences to that of the horse liver 1° ADH .................................... APPENDD( C Figure 1 - Thermophilicity and thermostability profiles for T. ethanolicus 2° ADH Pro residue mutants Part A ........................................................................... Part B ........................................................................... 196 208 1° ADH 2° ADH DEPC D'I'NB EUI‘A EPR EXAFS GAPDH GuHCl HCA ICP MAIDI ORF PEG ABBREVIATIONS primary-alcohol dehydrogenase secondary-alcohol dehydrogenase alcohol dehydrogenase diethylpyrocarbonate dithionitrobenzoate dithiothreitol reaction activation energy (ethylenedinitrilo)tetraacetic acid trisodium salt Electron paramagnetic resonance X-ray absorption spectrometry Fourier transformed infra-red spectroscopy glyceraldehyde-3-phosphate dehydrogenase Guanidine hydrochloride Hydrophobic Cluster Analysis induction coupled plasma emission spectrometry matrix associated laser desorption ionization mass spectrometry open reading frame polyethylene glycol Polymerase chain reaction ribosome binding site Enzyme melting temperature in the absence of denaturant Enzyme melting temperature in the presence of guanidine Chapter I Literature Review Adapted from an article published in Ann. Biotechnol. Rev. 2 ALCOHOL DEHYDROGENASES Alcohol dehydrogenase structure and function Alcohol dehydrogenases (ADHs) are integral to both prokaryotic and eukaryotic metabolism. Categorized as 1° or 2° based on their displaying higher activity toward 1° or 2° alcohols, these enzymes typically act on a broad range of substrates. The 1° ADH isolated from mammalian liver is proposed to oxidize alcohols and possibly steroids during their detoxification [l]. ADHs are also involved in catabolic alcohol consumption by acetic acid batceria [2]. More commonly however, catabolic ADHs mediate the terminal electron transfer in solventogenic fermentations by yeasts [3—5] and bacteria [1]. Alcohol formation by these organisms provides the oxidized nicotinamide cofactor necessary for glycolysis, eliminating the need for an exogenous terminal electron acceptor. Catabolic 1° ADI-Is have been identified in mesophilic (optimal growth between 25 °C and 50°C) [1,6], thermophilic (optimal growth between 60°C and 80°C) [7-9], and hyperthermwhilic (optimal growth above 80°C) [10,11] microorganisms. However, catabolic 2° ADHs have only been reported in mesophilic [12,13] and thermophilic [8,14-16] bacteria. The mesophilic bacterium, Closm'dium bezjierinckii, expresses a 2° ADH during propan-Z-ol production whereas, the thermophilic Thennoanaerobacter sp. and Thennoanaerobacteriwn sp. use a 2° ADH to form ethanol - a 1° alcohol - from acetleoA [7,8]. These thermophilic bacteria ferment sugars to ethanol and acetic acid (neither butanol, acetone, nor propan-2-ol have been detected) and express a functional 1° ADH concurrently with the 2° ADH. The Thermoanaerobacter ethanolicus 39E 1° ADH is proposed to function in electron transfer between NAD(H) and NADP(H) and in ethanol consumption while the ethanol producing 2° ADH can function to reduce both acetleoA and ethanal [8] using the generalized enzymatic reaction: ii 9" n—c—cu3 + NADPH+H+ —> n—é—crt, + NADP” r‘r ADHs are predominantly nicotinamide cofactor dependent Zn metalloproteins although Fe [17] and ferredoxin linked enzymes have been reported [18]. Theorell and Chance proposed an ordered 1° ADH kinetic mechanism where cofactor binding is followed by substrate binding, hydride transfer, product release, then cofactor release [1]; however a semi-random mechanism for product and cofactor release has also been reported [19]. Isomerization of enzyme bound to reduced cofactor is believed to be the slow step in liver ADH catalysis [20,21]. Cofactor binding is proposed to induce a slow enzyme conformational change after a rapid initial binding step, completing active site pocket formation. Even the NAD+ molecular geometry required for horse liver ADH catalytic activity has been determined [22]. Zn's catalytic role as a Lewis acid in the electrophilic ADH mechanism is well established, with the metal mediating direct hydride transfer between substrate and cofactor [23]. Kinetic and spectrophotometric measurements provide direct evidence that the alcohol or carbonyl substrate binds to the catalytic Zn atom [23-25]. This may explain both the broad ADH substrate specificities and why decades of research into 1° ADH structure-function relationships have failed to identify residues responsible for substrate binding and discrimination. ADHs are either dimeric or tetrameric, composed of identical or nearly identical 35 kDa to 45 kDa subunits. Both 1° and 2° ADHs have been reported to contain catalytic Zn atoms and Cys residues critical to catalytic activity [1.8.15]. The exact protein structure has been determined for horse liver 1° ADH by x-ray crystallography [26] but no 3- dimensional 2° ADH structure has been reported. X-ray structural data confirmed that 1° ADHs bind a structural Zn atom with 4 Cys residues and a catalytic Zn with it His plus 2 Cys residues. The broad specificity of liver 1° ADH is attributed to a wide and deep hydrophobic binding pocket seen in the x-ray structure [1]. Nicotinamide cofactor was 4 shown to associate with a Rossmann fold in the 1° ADH structures as reported for other NAD(P)(H) binding proteins [27]. The identification of Rossmann fold consensus sequences {Gly-Xaa-Gly-Xaa-Xaa-Gly-(Xaa)13-m[negatively charged amino acid for NAD(H) dependent or neutral amino acid for NAD(P)(H)]} [28] provide sequence based predictions of cofactor specificity linked to this particular structural fold. Stereoselectivity is both a hallmark of enzymatic catalysis and an area of great biotechnological interest. ADHs are highly enentioselective catalysts despite their characteristically broad substrate specificities. In pyridine dinucleotide containing enzymes, stereospecificity results from the cofactor attack angle according to Prelog [29]. Carbonyl reduction can result from hydride addition to the Re face, following Prelog's Rule, or to the Si face, following Anti-Prelog's Rule. So based on the orientation of the carbonyl relative to the cofactor and which cofactor hydrogen is transfered, there are 4 possible mechanisms for enantiospecific ADH reaction (El-E4) [30]. The E1 mechanism is defined as the pro-R cofactor hydrogen (HR) on the substrate Si face, E2 involves cofactor pro-S hydrogen (Hs) attack on the substrate Re face, HR transfer to the substrate Si face results from an E3 mechanism, and reduction on the carbonyl Re face by the cofactor H3 describes mechanism E4. The horse liver and yeast 1° ADHs operate by the E3 pathway, producing (S)-alcohols. E1 reactions catalyzed by L. kefir and Psuedomonas sp. (SBD6) ADHs form (R)—alcohols [30,31]. Overall structural similarity among ADHs is presumed based on their similar gross morphologies and functional characteristics despite their dissimilar peptide sequences [1]. The proposed structural similarity between ADHs and the difficulty in determining residues involved in substrate specificity makes comparative analysis between enzymes with overlapping but difl'erent specificities important. Cloning the genes encoding these enzymes would provide adequate supplies of protein for structural studies and the means to perform site directed mutagenesis experiments. Mesophilic (GenBank Acc. No. D90004, L02104, M91440, X 17065, X59263), thermophilic (GenBank Acc. No. D90421), and Alct 5 hyperthermophilic (GenBank Acc. No. 851211) 1° ADHs have been cloned and characterized. The 2° ADH hour the obligately aerobic mesophile Alcaligenes eutrophus has been clomd and characterized [32] and the meSOphilic C. berjen'nckii 2° ADH gene sequence has been deposited in GenBank (Acc. No. M84723) but no characterization has been published. Alignment of peptide sequences derived from the translated gene sequences indicated the conservation of putative catalytic and critical core residues among the A. eutrophus and liver enzymes [32]. However, an alignment between the A. eutrophus 2° ADH, liver 1° ADH, and T. brockr‘i 2° ADH (The T. brockr'i peptide sequence was determined by Edman degradation of purified protein fragments) led Peretz and Burstein to conclude that insuficient similarity existed between the 3 sequences to make comparative structural predictions [33]. These conflicting conclusions indicate that the validity of 2° ADH structure-function hypotheses based on structural comparisons between 1° and 2° ADHs awaits assessment by genetic manipulation. Alcohol dehydrogenase industrial applications While industrial organic syntheses have typically been dominaed by chemical- synthetic processes, the rising public environmental awareness (reflected by new legislation) plus the recent developments in enzyme biotechnology suggest that enzymatic and mixed chemo-enzymatic processes will progressively be substituted for polluting or toxic chemical processes. Four enzyme groups - carbohydrases, lipases, proteases, and oxidoreductases - have a high potential for synthesizing peptides (e.g., pharmaceuticals, neuropeptides, and specific peptides for research purposes) [34]; flavors and fragrances (e.g., benzaldehyde, naringin) [35]; non-peptide polymers (e.g., polyesters, polyphenols, polyacrylates) [36]; chiral compounds [37]; and surfactants (e.g., monoglycerides, sugar fatty acid esters, alkyl glucosides) [38a]. Typically known for their hydrolytic activity, lipases and proteases can, in specific environments, be used as synthetic enzymes. Enzymatic syntheses present numerous advantages over chemical syntheses: they are 6 usually highly specific (i.e., enantio-, regio-, and stereo-specific), environmentally friendly, and their products are usually easily biodegradable. Chiral alcohols are common in biological effectors (eg. the neurotransmitter norepinephrine) and chiral alcohols can be used as stereospecific reaction centers in bioactive compound manufacturing. With the top ten optically active drugs representing sales of $10 billion (U .8. dollars) annually and with the recent FDA policy focused on pharmaceutical enantiomeric pmity [37], strategies to produce chiral compounds as either drugs or as synthetic intermediates have growing indusuial potential. ADHs, cuantioselectively active on a wide range of substrates, have been the focus of substantial research into enzymatic synthesis of industrially relevant chiral compounds. The horse liver 1° ADH is optimized for the interconversion of aldehydes and 1° alcohols but it will stereospecifically reduce ketones [1]. Mesophilic liver 1° ADH has already been used in the analytical scale production of chiral cyclic alcohols, polyalcohols, alcohol containing aldehydes (cg. L—glyceraldehyde) or ketones (cis-cyclohexanone-Z-ol), lactones, and organosilicates [see 38,39]. Similarly diverse and numerous potentially valuable chiral bioconversions have been performed in vitro and in vivo with yeast ADH [see 38,40,411. Resting cells of the hyperthermophile S. sulfataricus were shown to stereospecifically reduce ketones and to express a 1° ADH with activity toward ketones and 2° alcohols [42]. Ketones and 2° alcohols however, were poor substrates for this enzyme. The mesophilic, (IO-specific ADH from L. kcfir has high activity toward alcohols near cyclic or aromatic structures [31,43]. A range of chiral alcohols was also created using a mesophilic (R)- specific 1° ADH from Pseudomonas sp. (ATCC 49794) [30]. The potential for enzymatic conversions in specialty chemical synthesis and racemic resolution is widely recognized [38,44—51]. However, the expensive cofactor requirements (NAD, NADP, FAD) of oxido-reductases such as ADHs may limit their applications [37,52]. Several systems [51-54], have been demonstrated for cofactor recycling, but current biotransformations are typically performed in whole cells, where the cellular 7 machinery retains and regenerates the cofactor [37,40,41,52]. The specific cases of chiral diol synthesis using a host of microbial cells have been compiled and reviewed [55]. Many important specialty chemicals and their precursors are sparingly soluble in water, requiring a solvent stable biocatalyst. Horse liver ADH was shown to be catalytically active in numerous organic solvents with specific activities similar to that in aqueous solution [56,57]. The activity if horse liver 1° ADH was also examined in water-oil microemulsions to enhance conversion rates for substrates with low water solubility [58]. The hyperthermophile S. solfataricus 1° ADH [10], and its resting cells have been used in chiral biotransformations [42]. This enzyme has high stability in the presence of solvents [59], but its low catalytic efficiency toward 2° alcohols and ketones might limit its applications. The importance of biocatalyst longevity and stabilty to their large scale utilization makes necessary the development of strategies to stabilize mesophilic enzymes or the isolation of highly stable analogs to them. Inuinsically stable and active at high temperatures, thermophilic enzymes offer major biotechnological advantages over mesophilic enzymes: (i) once expressed in mesophiles, thermozymes may be purified by heat treatment [60—62]; (ii) their thermostability is associated with a higher resistance to chemical denaturants (such as a solvent or guanidine-HQ); (iii) performing enzymatic reactions at high temperatures can allow higher reaction rates, higher subsuate concenuations, and lower viscosity; and (iv) there is a higher product yield during certain reactions due to chemical equilibrium shifts with high temperatme. Isolating and characterizing thermophilic enzymes therefore, provides natural examples of structurally stable enzymes with high active temperature ranges that are potentially robust industrial biocatalysts. Comparisons of these thermophilic enzymes to their mesophilic analogs will also provide insights into what molecular interactions confer high folded protein stability. The extensive ADH structure and function research literature have been well reviewed [1]. AI-. mu“ 8 The emerging ideas regarding protein thermostability, thermophilicity, and folding molecular mechanisms will be reviewed here. THERMOPHILIC ENZYMES Thermophilic enzyme sources and diversity Originally, thermophilicity was a property associated exclusively with spore forming bacteria. Thermophilic enzymes were believed to be unstable, with high protein ttn'nover rates explaining why thermophiles did not grow faster than mesophiles [66]. However, Thermus aquaticus, a non-sporulating thermophile, was shown to express inherently thermostable enzymes [63], overturning these hypotheses. Subsequently, most thermophiles and hyperthermophiles [64—69] have been shown to possess inherently stable enzymes that function at temperatures above the organism's optimal growth temperature [65,70]. Enzyme thermostability is the protein’s capacity to resist irreversible thermal inactivation, and is commonly reported as the enzyme’s half-life at a given temperature [71]. Enzyme thermophilicity is defined as the temperature at which the enzyme is optimally active [71]. An enzyme is thermophilic if it is optimally active at temperatures above 60°C [71]. Although thermophilic enzymes typically originate from thermophiles and hyperthermophiles, some mesophiles produce enzymes active and stable above 60°C (e g. pancreatic ribonuclease A). Mesophilic enzymes are optimally active between 20°C and 60°C [71]. Mesophilic enzymes, typically from mesophiles, include most eukaryotic enzymes and those from mesophilic bacteria and archaea. All known thermophiles (optimal growth from 60°C to 80°C) are microbial including bacteria, arehea, and some blue-green algae which grow at temperatures up to 60°C. The predominance of prokaryotic life forms at thermophilic temperatures is consistent with the appearance of prokaryotes while the Earth was much warmer, with eukaryotic life forms evolving much later. Thermophiles have been isolated from hot catchments inch; sprigs, soils, shal‘ (ii) microbially se‘. and (iii) industrial sludge sysrems, o diverse as their . htmmphic‘ Ch: among Others. Tables 1 rhcrmophmc and mmmbiliq' y Itflecg the lack Specific 111cm 0d Cases, it inCluds Potential blotec thermophiles \t Enzymes haVc mmstablc “ growth tcmPCT Sfiingenuy for alcoho1 dehyd [GAPDHs], e function of m thumophllic ‘ mbstmtc Dr C D°3Di usually haVe ] 9 environments including: (i) natural volcanic environments (continental solfataras, hot springs, soils, shallow marine and deep-sea hot sediments, submarine hydrothermal vents); (ii) microbially self-heated environments (e. g., manure, coal refuse piles, compost piles); and (iii) industrial environments (e.g., food industry effluents, hot-water lines, sewage sludge systems, oil drilling injection water systems). Thermophiles, as physiologically diverse as their mesophilic counterparts, include species that are aerobic and anaerobic, heteromophic, chemoorganotrophic, chemolithotrophic, automophic, and phototrophic among others. . Tables 1 and 2 list enzymes which have been characterized and/or cloned from thermophilic and hyperthermophilic organisms, respectively. Thermophilicity and thermostability properties are included where available. The heterogeneity of this data reflects the lack of consensus on the way to guage these properties (we have proposed specific methods for standardizing these data [71]). This list is not exhaustive, in most cases, it includes only examples of each enzyme type and it focuses on enzymes with potential biotechnological applications. An extensive list of enzymes purified from thermophiles was published by Coolbear er al. [70] and detailed descriptions of individual enzymes have also been compiled [64,70,72]. Thermophilic enzymes are inherently thermostable with optimal activities at temperatures near the original organism’s optimal growth temperature. The proximity to organism optimal growth temperature holds more stringently for enzymes within a single su'uctural family (i.e., a—amylases, proteases, alcohol dehydrogenases [ADHs], glyceraldehyde-3-phosphate dehydrogenases [GAPDHs], etc.) than across a range of proteins, suggesting that stability is partly a function of the match structural fold. Note that the Optimal activity temperatures for some thermOphilic enzymes, oxidoreductases in particular, have not been determined because of substrate or coenzyme (e.g., NAD, N ADP) instability. Despite being synthesized in a mesophilic host, recombinant thermophilic enzymes usually have kinetic and thermal stability characteristics identical to those of the native rr\r=\elr\ Ii... hi- ..itl .Illllttl lit-istll: ilhlmufldddlll idflllll A. “mu—N. ._.+ tr newcomeui can . . - . REG . - .. . 3 0 5w. NIQNU we ArU 8v .3.\\93 §~kt§h§¢§~b¥< gzkogsozé MN 31 I. bore 3 v o o mmm32%e§.m 8.? as .m 6 6.3389. 6.8 33.8 35...: ESE? <75 2.“: as .m 6 65355. 9.3.35.8 8-83.3 secesficc a. Essence. magma—age... R m .o g. .a. cafiwcesfie unease .e 2. Quebec 838:8 §to~§o§a§§o§ 88262 85.6 6&8 agencies... .axo .m .U §~§§ goagogtofi .28? 583m 3 vegans. oaen 658$; 23.836 senescence; 28.. 6.3588 6.3 Eaasaescea. 33% 3.20 8M 23 8303523 an? 6.83%... «53.2 6.8:.62 Sausage: geesecéaé mmmfioaemmeexe 38888 35:88:55 36:28:55 Age 539% 282—3 .62 com 30:00 gem gem :32."qu gem 2:35”... 3:53.505 5?... BEES—Eon... .n 2.2:. Swami 2. Cf; ». 3 n. _ Uta i \1 §U>Um 0:33 .Ra. §.~\U~UEQKV~§§X»EVRN §=u~3=38~kfl=< N... :53 .m .0 .w. U%_E Cw W: 35950 006$» QENQQSSM K 0vdca~==~a0~>ne< Done 2:“ . «e w- Uonxxt 3:033. 66.. .L\Uocm-hn §~§§§3he§vktgvi~ gxfithéxb 0vé§~3=3nrv~kwt< 11 we be we 99 com on new we 2: 2. com 8 .2: no“ .2: 9:08 QB 8m 9: 08 a ma 3 3m .3 3 mm mm 3 md .55 .m .0 £98 .m .U .m .HQXD .m .0 .m .55 .m .0 .axo .m .o d: .hQKO am .0 .m .HQXO a” .0 .m m .HQKO am .0 .m m §e§u§ .u 9858 3:308 aimsszssgm 9%“me «£508 608V .3. §§§o§§§s 5% 258 *8 09:83: Oévgaaasfié .u 0% cu afibeugaau mange mm 9.2. §~8§~ g 935:. 2 08,503 .a. 338m 00828 w BoonSS <3 3.5.»: was} mm 3392 §§€§u§ s 5 N_— ——— 0-— .LQXO . I‘Sun AN—rvflu 2E W ATV UochEE ON hu-Hdu ”slung: *8 WEE CS 3002. Odin vowels ~ 6:“ .Rn :QSCQMN < a Vg “ .§h. §EKN~$N ”(VS m .Kh: SERV‘rN h§\\-\Q§N~\\A~K~\V\h 5‘ 9.6.5.29: Omacmugé 022:»qu OWQCQ-~==QAVOZ 12 0Q mm“ as 000 «Q 000 mag NS -~ ofi 03.x: b: .m: v: m: 0: N: g: a: mo— m0 .098 .m .0 m m m m m .098 .m .0 0.00% 2 SUEZ Ea .0EESBw .+~w—2\+N:E+v 055m 038m 0owQEE 3 Ba“ 0380 0.0%}... 0350 0% we 0ontfia mm oootfi 0cm» 3 a 2 000.00 0.600 gowA 0.228 0.3.: .8 055... 306 283 0.0%.... a 5 0.822 306 28 0 +0 0.238 cm 306 28 0 +0 0.0%.? 8 0.83 .00.. 268 $8 0.00.5.8 0.5.00.8 00% 002. 0.00.00.00.05 0330.2. $03.8 $30.8 8.0:: 3.3% DOE 0.2 0030.8 3% 0.3.00.8 0.0% 0.800 053095052003. .m 53.000030000005030 .0 Cost §03m§§00§05 ~§§03§02 33%....050880 .m Eaimcumuu .h ggoéfiuaogwi .0 fluuboguua .0 §0ub0§008 H in §E§E§E0§ 6.80 338%.. 0583955 E.» 300:8 3&0...” 2.9.2 .0... 3.....5 «0 3:80 05.3%.505200: .m 32 .5. 0.580 <32 .0. 35.2: «$3 .5. 352: §~3§~E050§0~u .m 8038M 083x 008088m 829.800 085. Sgafioaoofimoam 00805:? 038085 805:? 00005 00.2030 0020833408002m mmm<03 nz< $030028— .mmm<>q 38.83“ 30283“ 9.8.53; 0082000003 38035. 85—352 0353:... 8.83:5 Oman—musk— 820.80 08:08.0 8083:0002 .tbntkyficcu :02. E: SE £8.22 :C 02...:5222Q £5 k0 hymum 0:” 4 WNNJ t3.v.§.¢«yk~w~\.ub~\h §~\\.fiv§\»\30 \QVUCCOK a F. 1:: .5: 4m .L a: v: §>$~§E§S K 3.22:9: 961*] am. R 0533. Qt 3003" an: uiuhuabu K 0255.94 0201* .000 new hN _ .NC 575 .m .U ~7va +9 0005 No. azhzlhvoomw ufifihkzlkwihgvs uh u§g~ 95th IN—.FN- CLXU .V. NU 5: P. . hm unabguw usurvhkwgvbhhu. uh GREECEN Oh°~sfiufl 13 @8298 9:55:88”: ”.98 "8803623 22%29méanonfibobc "mm—.20 ” 2389655523508 ”E ”Egg 055808 “82 522—8885 um 8:3?6 9585 as»? "8 68538 um 60:20 no dam—Boga 83 8: ms— :88». no Rufiaaa $5 mo .0168 05. 9 Anna gubcguuuu .55»ch bbauom a 55: .55 .w .o E 2 £3 855 s afiefias x ad 3 Oman? c.5502 mm: 95:38.5 afiefi as? ~UOU ad 5% .55 .m .0 «cm: +v 955 Ne 2:308 SumEsasE§ g 035:3 as? «2.3— .55 .m .0 E. 053808 gogcguua K E 803K I ll-lnlld'H‘ ha‘lfiéh q %u&%<&UDQHZC~=K: . . «95.x: TEEEXNES 530:“ 35.55% . b 52.537.5be 5.. «2.5% a . a Z guy. {vamp—”flu“. 02:05:”.— Utrmhcl Sui—3M5 ”SSW . . . . . . nEa~=¢ML° 0~=IEOEL0£Pm0QH£ Eng EOEHNQELOCE. .N mine-3B 14 w: >3 9; a: 3: m: m: 3.3: N: .33: 2 3; w»: .5262 m2 we 08 V9 mmfiNmfi S 9 .oZ mom .55 .m .U .55 .m .0 m0 9.0% 2 UoooHEN 93:3. 2: 00358 3.2 camsoooSNA woman 9822 008% NE DonSBm.m UoOmEN Gone :2 H8 03% 000% UooaA DonoA gives 33.5% 98% 330%? 35.502 odmnaoafi Dona 008A 33de 608A ”.mmeuoow wwnfig Dom? 608: .383 §Em§© $33 .m 398:3 .: Gone Eugx fingegs: Aoomwvwefiiuxaafixszgax ~85. K Come §§ afieigwfi Coco: .59: “.3885 60% gurus swag»: 3985 H §ufi8§8 .m. 2.82% .m 608: vmm £22.: .5 .38th .m 608: §uot§§uu8§£ 6&8 asks: 398$st Comm not §utu~§8 azasebam 820»ng @503on NE ofiuowofimm 3688 5:696 emanowofimm Ao>uo§bnv~b aaaomeumm $920 $956 $920 a682 Geog 333“an .3292 using... Ba 0083 has 028a $8“ 53x0 =8 0.6%: Ban 2505 *8— OomQEEOm 000258 cm 0 a fi 0 can? 9E 92 3E .98 008A nomeuoo? fiasco? nsmgboc? cameo...» -NvOm .-~v0m 2 SS 5 33:3,? .3. 2m.5°8A =3. 2 9S6 Agave: Steven? mgamg DORA 3g .m fleas s guns g 333 a. 3.3% .k 95.29: .5 $3 .m ”333% .< 6.8 :fiméafimogué Egg .3 E3 .3 0.3 533 3.333.»: 3.2% .3 SE2 .5 gtg K a Saba—on <75 Agufiboq :63 fl unsung—om <29 gfigofiosaé‘v 3.353 .5. mmmkN.\. .\< Omucmu~ 0~EOUK~WC£Q8£Q-N 3.5m TV Ecsuaczv -NVE £.¥.._~¥ .2 n; + .-~..Ow .2 N . owonucuhigo.‘ «1.: AUOCOV O_£.Zm $87.. \WCIQOCOQ AUONQV ~\N\§ \Q E5105 RV hm» me — hibhthNK-g 38°U°§Mu§lhflvQ -N OWQLUENOAN 3 gag—E gages—mggé 35¢ 35.3275 own—>38 Aaazoogfiv 0333.8 A§=o§§ 82550 mmmafiomA—w: 38: agaoamsfi.~ gamma—Em .325de E 3.53.8 <75 do. Ni —I— Ox- : 00v. . Ecru $9.4m. no.2: ongmg 35.2 35:53 onaxznd Dunc—>5: CN 3: 3303 99¢.th 009 ~ O.h:n~\uv: nah‘wxxx. .V vaccxeaxxah §h§~§\. .nN .Qu. QankENSN 605 my K..~\.U.U b§UUQ§U§.N 859632vo :5 .1" mu 5 DEE AV 2810:8050: §Umn83~o -5 OfiflnmfiogNOI Olefivmmogzo 17 3208628 22%29338385 "man—<9 §§§-_§é -mz ”Hm—23.010 nauounogfiofigoficoifiofiézwz "gamma 69.8898 359.5888 ".55 63838 ”m 628—0 "0 § 98:58 2 9.218 Bangs 03588.32? :2 55 .m .o 935$ SECOR 5388.5 38089.. as? awfi u: v: .889: .m 8805:? 8520 5: 99:5 9.8.3 33.3: 9.505% 8.8920 53 0&2 o. 2&8 .2333 .2 > “$888699 5.5 mm 983 :z< .mmm scrambled structures [scheme 1] (active) (inactive) (often precipitate) Representations of the natural log of enzyme residual activity plotted versus time are usually linear. fitting a pseudo-first order equation {Equation (1). with k=rate constant and t=incubation time} . ln(residua1 activity) = do [1] Equation (1) contains no protein concentration term, suggesting that the rate-determining step is intramolecular with respect to protein. The only way to ensure that a pseudo-first order rate law reflects a truly intramolecular process however, is to measure thermal inactivation at various initial protein concentrations, verifying that 1: remains constant. Temazic ant thermal inac' Prote protein mole. thermophilic recombinant and in differ: native enzm Ihcrmosta‘oih Whfional 1 Im'crsihle n dcnamranon, Min chum infO‘TDation 11 (9pm? 304 m8 [203]. 22 Tomazic and Klibanov’s model is consistent with an intramolecular rate~determining step in thermal inactivation if the natine to non-native protein step rate determining. Protein stability is due to numerous ionic and nonionic interactions within the protein molecule and between the protein and the environment [198]. While some thermophilic enzymes have been shown to be stabilized by glycosylation [199.200], most recombinant thermozymes - expressed in mesophilic hosts, in the absence of glycosylation, and in different cellular environments - remain as thermophilic and thermostable as the native enzymes [141,143]. This observation demonstrates that thermophilicity and thermostability are encoded in the peptide sequence; they are neither a consequence of post- translational modifications. nor of non-covalent interactions with cellular components. Irreversible inactivation of mesophilic proteins typically results from irreversible denaturation, due to the disruption of numerous non-covalent interactions rather than to protein chemical decomposition [201]. Thus, the original protein conformational information remains intact in many denatured proteins. The protein stabilizing energy (typically 30-65 kJ mol'1 [202]) is lO-fold smaller than the magnitude of the opposing forces [203]. Since each additional hydrogen bond or salt bridge can contribute approximately 2-20 kJ mol'1 to the stabilizing energy, and since the possibilities to add hydrogen bonds or salt bridges in a protein are vast, it is likely that the protein covalent modification rates, rather than unfolding, determine the theoretical upper limit of protein thermostability. Ahem and Klibanov [204] reported that peptide depolymerization becomes significant at temperatures above 100°C. but oxidative decomposition of cysteine. asparagine. and glutamine residues has been shown to limit the thermostability of some enzymes to temperatures below 100°C. The identification of enzymes active at temperatures above 120°C suggests that the non-covalent protein structural interactions are strong enough to allow stability to approach the theoretical maximum temperatures defined by peptide bond destruction. Thermophilic While th from 65 to 70°C Thermophilic am with variations ir conditions of hi; directly involved denamration are . umsr be more res amines similar 1 atom positions lm Attire Emptraturt Differential scanni Monably Constar absent: 0f Signific mm 3" tllJiCa Mop-la] ”Gilliam 23 Thermophilicity While the system kinetic energy increase (~RAT) is the same from 25 to 30°C or from 65 to 70°C, the total molecular kinetic energy is greater at high temperatures. Thermophilic and mesophilic enzymes similarly resist the structural changes associated with variations in their molecular energy, but thermophilic enzymes must do so under conditions of higher total kinetic energy. Therefore, while protein denaturation is not directly involved in enzyme thermophilicity. structural interactions which oppose denaturation are very important for thermophilicity. In other words, thermophilic enzymes must be more resistant to denaturation, while having dynamic structural states allowing activities similar to those of their mesophilic counterparts. Because of sub-A variations in atom positions known to be tolerated in a functional active site, it is believed that, within its active temperatrn'e range, an enzyme maintains its average su'ucture within strict limits. Difi'erential scanning calorimetry measurements indicate that protein heat capacities remain reasonably constant within their catalytically active temperature ranges, indicating an absence of significant structural changes. Arrhenius plots for thermophilic and mesophilic enzymes are typically linear. suggesting that mesophilic and thermophilic enzyme functional architectures are similarly, tightly controlled tluoughout their respective temperature ranges (significant su'uctural changes that alter the functional architecture are expected to cause non-Arrhenius behavior). Biphasic Arrhenius plots reported for some mesophilic and thermophilic enzymes [82,145,205] represent an important exception to the typical Arrhenius-like behavior, however. discontinuities are not a specific trait of thermophilic enzymes. In the cases of yeast and Thennoproteus tenax GAPDHs, it has been proposed that the discontinuity extent indicates the degree of enzyme thermostability and thermophilicity [145]. It is critical however. that such interpretations of Arrhenius data are verified by careful biophysical and kinetic experimentation. More likely, the similar Arrhenius plots obtained for mesophilic and thermophilic enzymes suggest that all proteins respond similarly to temperature. In their res; names are more personal communic mesophilic counter temperatures is obs temperate rigidiz} 1027\in at these ten: mural change or more Vmax (kw, throughout their ac: MOM? Pmdicts “mm. a set 11'. Nuanandp a fun M15 that a mom Emily mt than a: minimum AUhCI mmnophilic analog I magi-'8 COnthing - Recent “'0' T. ’lfhnosuzfun-gem Change) (— . . _ 24 In their respective temperature ranges, psychrophilic, mesophilic, and thermophilic enzymes are more rigid at low temperatures than at higher temperatures (G. Petsko, personal communication). Thermophilic enzymes are significantly more rigid than their mesophilic counterparts at room temperature [198], however, increased rigidity at low temperatures is observed in all enzyme categories. It has been proposed that the mesophilic temperature rigidity of thermophilic enzymes [198,203] was responsible for their low activity at these temperatures [140]. This implies that a catalytically significant enzyme structural change occurs between low and high temperatures. However, the increasing enzyme Vmax (kw) values with temperature seen for thermophilic and mesophilic enzymes throughout their active ranges are typically consistent with the Arrhenius relationship. This relationship predicts that the percent of maximal enzyme activity measured at some temperature. a set number of degrees below the optimal activity temperature, is predominantly a function of the reaction activation energy (Ea) (Fig. 1). Thus. the theory predicts that a more thermophilic enzyme would have only a slightly broader temperature- activity curve than an enzyme with a lower optimal temperatme and the same Ea. Therefore, the Arrhenius equation can explain poor low temperature activity for thermophilic analogs to mesophilic enzymes while assuming that only the system thermal energy is controlling the reaction rate (there is no catalytically significant enzyme structural change). Recent work in our laboratory demonstrated that E. coli. B. stearothermophilus, T. thermosumm'genes. and T. neapolitana xylose isomerase Arrhenius plots were linear [206]. Thus, the percent of maximal activity at any temperature below the optimal temperature for an enzyme depended principally on the activation energy of the reaction. These data clearly showed low thermophilic enzyme activity at mesophilic temperatures. Also. the excellent agreement between the observed temperature-activity dependence and Arrhenius theory suggested that throughout the active temperature range for each enzyme no catalytically significant structural changes occurred. Figure l. Arrhenius dependence of reaction rate on the temperature and activation energy for mesophilic and hyperthermophilic enzymes. Tmax is definedas the temperature formaximalenzyme activity andEais thereaction activation. energy. Relative activity (7 Relative actlvrty 26 1.2 1 O I Ea = 80 Idlmol, Tmax = 90°C ' 'i I: Ea=80kllmol.Tmax=37°C 0 ‘ e Ba:20kJ/mol,Tmax=90°C .30 0.8- o Ea:20kJ/mol,Tmax=37°C '0 o.° I 0.00 0.: f 0.66 0 ° q: o r .. O I n 0.4- I .D In '1: 0.2- '3 0.0 ::;:::::::IIIII-IIIIIIIIIIIIII'.."5.- . .150 -100 -50 0 Change in temperature from maximal (°C) The Arritcr been attributed to a rigid inefficient ca Although a discor: discontinuities can other phenomena r. til‘fifanphilic enzy: 27 The Arrhenius plot discontinuities reported for some thermophilic enzymes have been attributed to a catalytically significant enzyme structural change from an excessively rigid, inefficient catalyst to a less rigid sturcture optimized for enzymatic activity [140]. Although a discontinuous plot is not absolutely consistent with theory, Arrhenius discontinuities can result from temperature dependent changes in the reaction slow step or other phenomena not linked to enzyme structure [207,208]. However. whether thermophilic enzyme Arrhenius plot discontinuities indicate a temperature dependent change in enzyme structtne and whether such a change can account for the low mesophilic temperatm'e activity of thermophilic enzymes, remains to be determined. As kinetic data accumulate on thermophilic enzymes. it becomes evident that, despite their activity at high temperatures, thermophilic enzymes catalyze reactions with Vmax and Km values similar to those of their mesophilic counterparts at their respective optimal temperatures. Thermophilic enzyme-catalyzed reactions were initially expected to have high catalytic efficiencies based on extrapolations of mesophilic enzyme rates and the assumption of similar substrate affinities for the mes0philic and themrophilic enzymes. Instead, thermophilic and mesophilic enzyme Vmax values at their respective optimal temperatures are typically similar. as is typified by a series of xylose isomerases (Table 4). Substrate affinities for thermophilic and mesophilic enzymes at their respective optimal temperatures are also usually similar (Table 4.5) [215]. These similar Km and Vmax data yield similar optimal temperature catalytic efficiencies among analogous thermophilic and mesophilic enzymes (Table 4). While higher reaction rates with increasing temperature can be attributed to increased rates of collision and a larger fraction of the substrate population possessing kinetic energies above the reaction activation energy, the temperature dependence of Km is less clear [208]. Even for the simple Michaelis-Menten case (equation 1) ..... 21.11. L. Eases K ghcfibxhfiocce K u . .- evil :3 mo. hm. 2.3mm Cmvfi. OK; We hN~ CK: Cc. m 564 he Hutuktsgxadgus K 33m Cd? 95? .0 .3 cc SEQRENSQéfi .9 ON. Cm OCR. O MNtNCN \\00 .rQ :.:=.—ZE\: =:=\: g‘gmog do. N — .32 HOW ~ A Ev—bcov— flag! a .22....) EDMOTAK 50.x Eva ”ha-Egg Egg ICB:SQOELOS. awn-I tfl-cfln-OIOE 50.: unfit-Luann IQQILUEOI- 00°~Hl hi UQNH he |~=I~I=09 U~sfia-ufi .‘ 0~A~I.~. .592 €255 8%: =5 .338 a WEE as EB .+£z Us 592 E 3325 a guage £936 883. 222 2:. a ‘2 Son 3%: $03,638 amassi 83H. a we o2 M: o 2. no N... o _ m2 2 8d «:3 8-0 m8 3. 33H .m. E E ”85 83. w: E N .a 388.5%“ 5 mm 86 So 3 E 3 Eggs .m. SN 2 mod ES 2 .1. 3m 33:3 g SN 82 28.9 $3. 2 3. E 93%;» .2 :~ Sm 2.3 23 3o 3 on §S§Eb .m as 2 ~35 35 2 Ed 3 a8 .m g 632%. 52 fix €28“ Macaw E92 592 +£z sesaaoxo.~ assuage :5.an 925 SM ‘ 28 8.23.52. .23 8.22805 E9: 35...:— uouauouobumsou 82:33» .8.— 3528 3.25— .m ~35. I cocoa». ”ammo—3? 2 82m“? 8 29250 Bu mow—5.3x a 9: $2 $3 .2 8 . Us :2 Wu» onmw do ca 3:33.: & hfi odmu 82 .2 no Euuubguua H 5 3.2 $3 .2 no gummiassus s SN adv 8% .3. co 32QSE§°§2 .m m2 on com o nu.~u :8 .m Egg; 3. 3 Aug 3 8 .oz mom EvSaoM “SM §E> 081x .8“ EM gig. Ewing cue—£35.05 EB 3.2182: E9..— uoEEn 833E8— 82? fl 2:. .5.— 353289 2.05% .v 933—. r , on the I‘m VaIUC. C for the temperatur: Symbols A1, A4,; activation energies respectively. “1111 Mined m'thout a With Ifirnper'cmme. 1 “311333111128 for be dchydmgcnase (Ta “31033 r0 mesophi cmPOlattion of an IElationship yields Sltbsmc affinities Sim"Jar substrate af ”1} my 0r cmDIOy a 33m dcscfibed bv WYUC late enhan 29 k1 k2 E+S<—>ES—>E+P (1) 511 the Km value, composed of 3 fundamental rate constants k], k], and k2, yields equation 2 for the temperature dependence of Km. Km = (A-1A2/A12)e[(2Eal - Ea-1 - Bad/R11 (2) Symbols A1, A4, A2, Eal, Ea.1, and Bag represent the preexponential factors (A) and activation energies (Ea) for the reactions corresponding to rate constants k1, k_1, and k2, respectively. While the specific effect of temperature on enzyme Km values cannot be predicted without a great deal of mechanistic information, these values are expected to vary with temperature. Reported Km values were indeed significantly higher at higher temperatures for both T. maritima xylose isomerase [190] and P. furiosis glutamate dehydrogenase (Table 5). If thermophilic enzymes are exact structural and functional analogs to mesophilic enzymes excepting their greater resistance to denaturation, the extrapolation of an enzyme's Km values to different temperatures using the Arrhenius relationship yields results inconsistent with the assumption of generally similar enzyme substrate affinities at mesophilic and thermophilic temperatures. The observations of similar substrate affinities for thermophilic enzymes at high temperature and mesophilic enzymes at low temperatures therefore, argue that thermophilic enzymes must expend more energy or employ a different mechanistic strategy to so efficiently bind substrate. The circe effect described by Jencks contends that excess energy from substrate binding is used for catalytic rate enhancement [216]. Increased partitioning of this total energy to maintain substrate binding affinity versus catalytic rate enhancement in thermophilic enzymes or their structural limitations reducing the total efficiency of substrate energy utilization therefore, may account for the activity and affinity differences between Arrhenius theory prediction and observation. increasing therrnor stability and redue Protein folding Thennoph; mmnes fold prep: bl Similar mechan; Wm: stability a; “lime Chamctcris' and involvcs the m; We is believed . mm is Stabiliu lie-tum Parts of the Staggering number c rapidly in“) ”lei! acz; Stiles 0f wen‘dcfine. 30 and observation. Reports of enzyme mutants that have consistently lower kw values with increasing thermostability supports a direct structure-function link between folded protein stability and reduced efficiency of substrate energy utilization [217,218]. Protein folding Thermophilic and mesophilic proteins are typically similar and thermophilic enzymes fold properly at mesophilic temperatmes arguing that both classes of proteins fold by similar mechanisms. The link between protein folded structure, enzyme activity, and enzyme stability argues that understanding protein folding may yield insights into these enzyme characteristics. Protein folding begins as the peptide is synthesized at the ribosome and involves the rapid condensation of particular regions, or nuclei, into native-like states. Folding is believed to be driven primarily by hydmphobic forces, and the native protein structure is stabilized by a variety of hydmphobic, covalent, and coulombic interactions between parts of the protein and between the protein and the solvent [219]. Despite the staggering number of theoretical structures they may occupy, the fact that proteins fold rapidly into their active conformation indicates that they reach their final structure through a series of well-defined intermediate states. Rapid, cooperative protein folding suggests that partially folded intermediates are not stable and that the native state includes very few structural conformers of similar energy. This conclusion is supported by the conserved structures seen in x-ray crystallography and NMR Extending this conclusion to protein folding energetics suggests that the free-energy well containing the protein native state is swep-walled and deep, comprising most of the free energy range that describes the actual folding pathway. Freire et al. [220] proposed that at least two factors are specifically related to the unfavorable energetics of exposing complementary surfaces to the solvent in the partially folded/unfolded states: (i) the driving energy in protein folding and (ii) the protein stabilizing forces. The authors define complementary surfaces as surfaces which are not I solventexposed i unfolded states. 7| partly foided’un f0j aqueous solution cl states. Their mod; SOIVCHL and that t" PTOtein characteris Mildlm protein ii r’3gvdless of their I ”ho“ also nored : Wm of m: SOCor aIldhave a highly J eXistencc of mes: SI 3 l solvent-exposed in the native state, but which become solvent-exposed in the partly unfolded states. These result from regions of the folded core remaining condensed in partly folded/unfolded proteins that expose surfaces uniquely present in these intermediate states. Their model predicts that protein folding intermediates are highly unstable in aqueous solution due to the exposure of hydrophobic complementary surfaces to the solvent, and that these states are poorly populated. This CORE model relies on universal protein characteristics, not on structural motifs (at—helices and B—sheets) specific to individual protein folds, and is consistent with the observation that most proteins, regardless of their secondary structural composition, fold in a two—state process. These authors also noted reports that protein condensed phases which retain a “...significant percent of the secondary structure content of the native state, exhibit considerable flexibility and have a highly disrupted tertiary structure” are stable under certain conditions. The existence of these stable condensed cores indicates that secondary structural element interactions cannot, by themselves, control the protein folding pathway and folded protein stability. Haynie and Freire proposed a thermodynamic model to determine the conditions maximizing the stability of folded protein intermediate states [221,222]. Their model predicts that, in the presence of a denaturant, intermediate-state stability is independent of the free-energy variation between the intermediate and the native state, and that, in the absence of denaturant, the intermediate state AH is derived from the thermodynamic parameters of the unfolded state alone. This model implies that, thermodynamically, protein folding is defacto a one-way process. The successive steps depend only on their thermodynamic properties relative to the previous state and not to the subsequent one. Furthermore, contrary to the current belief, they predict that the two-state protein folding process relies on a small entropy contribution to the intermediate state stability. Examining protein folding as a function of the total energies of the folded and unfolded states further complicates this scenario [223]. At 25°C and high dilution it has been established that non-polar solute transfer into water was opposed mainly by entropy and not enthalp} driven. However transfer to water tanperanne-inde; highest between t transfer is propogt loosely appmxirn Ma DOD-polar rl t“literature. Abo significam There “My and enthaj apmemamly ch anta] ”511]“ 0fpamnomng mu. “mm in mCSO' 32 and not enthalpy. Under these conditions, protein folding can be considered to be entropy driven. However, because of the large, positive heat-capacity change of non-polar-solute transfer to water [223,224], the AH and A8 for this process cannot be considered temperatme-independent. The AG of non-polar-solute transfer into water is predicted to be highest between 130°C and 160°C, where TAS~O [223,224]. At these temperatures the transfer is proposed to be completely enthalpy driven. Considering that the cell cytoplasm loosely approximates a dilute water solution, and that a protein loosely approximates a small, non-polar molecule, protein folding can be considered entropy driven only at room temperanne. Above this temperature, both entropy and enthalpy contributions become significant. Therefore, at temperatmes where many proteins are folded (37-100°C), entropy and enthalpy contributions to the folding free-energy are expected to be approximately equal [224,225]. These predicted energy contributions are consistent with experimental results [225], and may help explain why, despite the temperature dependence of partitioning entropy and enthalpy, thermophilic proteins are correctly folded when expressed in mesophilic hosts. Protein folding is therefore, a robust process resistant to entropy and enthalpy variations. Unfortunately, folding intermediates do not necessarily contain local, condensed conformations identical to their counterparts in the native enzyme, sothefreeenergiesofstrucnualintermediatescannotpredictthepardalfieeenergyofthe corresponding region in the native protein [226]. Protein unfolding Protein unfolding is fundamental to protein stability with the Gibbs energy of unfolding a direct measure of the folded protein stabilizing energy. Protein destabilization by thermal energy and chemical denaturing agents has been well documented [227]. Extensive irreversible denaturation (i.e., loss of active architecture that is not recovered by the removal of the denaturing force) is far more common than extensive reversible denaturation (where active structure is regained upon removal of the denatming force). Since the system : reversible unfold: irreversible protti.’ image of prorein ft folding intermedia in protein thermal folding. For syster' pathway man be t* un‘olding must era Mess at equilibnt processes does n0t . completely unfoldet for both folding and Molding experimc: folded enzyme stabi 33 Since the system is usually chemically or thermally altered to initiate protein unfolding, reversible unfolding is not necessarily a mirror image of protein folding. The existence of irreversible protein denaturation also indicates that irreversible unfolding is not a mirror image of protein folding. Thus, unfolding intermediates do not necessarily represent folding intermediates. For example, the molecular aggregation and precipitation common in protein thermal denaturation did not occur for these protein molecules during proper folding. For systems at equillibrium, the flux through each specific fundamental reaction pathway must be the same in both directions. Thus, the mechanism of reversible protein unfolding must exactly mirror the folding mechanism under the same conditions. Not a process at equilibrium however, this microreversibility principle describing equilibrium processes does not apply to irreversible protein denaturation. The AG between the completely unfolded and native states in the same environment, however, must be identical for both folding and unfolding. Thus, studies of protein folding energetics using data from unfolding experiments can yield direct insights into the molecular determinants controlling folded enzyme stability. But, extrapolations from these data to the protein folding mechanisms may be misleading due to the potential for folding versus unfolding asymmetry. Protein denaturation is directly related to thermophilicity and thermostability. Enzyme activity usually increases with temperatme until it falls precipitously above the temperature of maximal activity. This rapid loss of activity is consistent with the loss of an enzyme’s active structure by denaturation. Many enzymes, however, have long half-lives at temperatures above their highest active temperatme. For these enzymes, the recovered activity upon cooling into their active temperatme range has been attribuwd to reversible, incomplete unfolding [228,229]. Denaturation is usually rapid and cooperative but the observation that denatured proteins in solution retain some condensed structure indicates that denatured protein is not necessarily complete unfolding. Mechanistically, denaturation is the loss of native 3-dimensional structme sustained by solvent molecules invading the unfoidint protein. protein solvent in Molecular met Intrinsic Fact0 The reteu‘. mesophilic hosts pepu'de. These ir. entropy of unfold Others. The char. fumtionally Sinai COutpatrisons to cl diminution NC mlymes and me: hllth'bd to be dll 34 unfolding protein. This hydration reaction (in aqueous solutions) must originate at the protein solvent interface, its surface. Molecular mechanisms of Thermostability Intrinsic Factors The retention of thermophilic properties by enzymes recombinantly expressed in mesophilic hosts argues that these pr0tein characteristics result in part from the nature of the peptide. These intrinsic factors include sequence specific amino acid replacement, altered entropy of unfolding, hydrophobic core packing, and loop region architecture, among others. The characterization of thermophilic and thermostable enzymes, structurally and functionally similar to well characterized meSOphilic enzymes, allows the use of comparisons to determine the factors involved in stabilizing enzyme architecture against denaturation. N 0 universal mechanism explains the differences between thermophilic enzymes and mesophilic enzymes. Thermostability and thermophilicity properties are believed to be due to subtle changes throughout the amino acid sequences of thermophilic enzymes. An extensive comparative amino acid analysis by Argos et 01. led to the conclusion that thermal stability was related to (i) increased internal hydrophobic amino acids and decreased external hydrophobic amino acids, (ii) the replacement of Gly, Ser, Ser, Lys, and Asp by Ala, Ala, Thr, Arg, and Glu, respectively, and (iii) to helix stabilization by more exclusive use of amino acids commonly found in helices [230]. Not all amino acid substitutions alter the function or stability of a protein. Specific interactions and residues, rather than all‘amino acids, contribute significantly to protein structural stability [231]. While single substitutions can increase the stability of an enzyme over 10°C [232], the thermostability intrinsic to thermophilic enzymes most often results from multiple amino acid substitutions [233.235]. The effects of single and multiple amino acid substitutions on enzyme thermostability are shown in Table 6. The examples were d. . . l I <‘.\ link-fie v O - . ,. e A 3.5% Easesguegsgi guests. :CEWKEEA... «your. E .25: 5.: 663;. < .5932. 35 5:45.. 232:3: “your. .e dear. EE MEG-m > 53.00 35:: 55.5.9037. U.£C—.QC._€.A£ Don.“ .fi 55¢. 55 VMQM ~ 05:0: 25:: OED-9' 30; Uohn 2. 53¢ 5.: in 4 NBMNN DON.“ nun .DSI cm:- NVASM U~h¢h~ gwfiggugt-d nl na=vm279cou — ugnvhl 3% N van-EKU lEI-p-II-BE Hu-nnaioRSELU-uu United-u... kc Iain-93‘! I‘lmv-Ibhw-s as Bubbhhhv .IUI . be! §i ‘I Ii“..hnil§..' BI t‘ .\ a .r. 35 «.2 3: ca .368 098% e832 8:82: 3. «madam—MB»? B ee§8§5 8:20 menses a E a.» 8» «a mg 88 «8:82: Wm OH 23% 83—823 035m .2. ban—88:55 .5 some 5- «nus—82M 33838: cox—.55— eswp.» a 96: :8 means—E. vie—=2 C as- 6.: «mm..— ade 0:29:55 Ge .5... 0e ages in 956 £288: moo imam toga: 2.55 3» 3% >89: 38 BaaeEosaaasm $83.2 we «a E... eases: 853852855 «gnaw a 3m 3m 5.: 3338: fig confine—55 cons—8 €35 0:33:55 85.35 05 gas» 53352255 2 Gov «F v.3 03.2» flan—082 we- a 3.. smegmgeafiizézafie 2:- Gangxiezaze 2:- Gmmagxxxgéi ca. Gnomeuxxxauéx: :8 Cd K: Egan gap—vans 85—A— . . s ., _. _ 8.....«828383 35:88:55 88265 3:: mg geese—5.553295 2 389 53333»: «See». co 5328623365 2.. "6L 8.2. m8? 23 excavate 333% 35333.»: g Passage 39m < asgefiusea 3832 came?» as 89m > .33 as: 238% usage caneéase 3.9m _ use»: an»: 059 33 vegan... as 89m A 6E 09% a .58» as 99m 08m genie 28.8380 .8 £2: A 582: «En—:28... 3.358535 2539 me 8:53 3839...:— 382... a...“ 3:85.689: 6 «3:. I I (PM. x: C. m .c E Cd Cd E ---- «.3» ME. ~06 0; EV. £2.33: 006 cd 5:. 800 Baku-D ho EOE—5133‘: 0>38t9¢2C0 “N. — Mia LHNN. \QVMM\ ha huauun—Ofin—OnuA—a UCG Ont-:0) CmBSU 0‘3n A L3:- _IDu-v Anvov 3 3| .. Ofléb MAI: ~ UAR- EG—Uh haugauv Ia \n-u-aCQ-‘u-h-EOE A - .3c<<. -=.—.< :a h=la 0:: :0 .5 is V 36 1mg: milk—23:3 3 ms§ea§u 3 ms:— 2 banana 3 5% 2. Ann. o.» mxgzmini .oéeaos 9m 22 5a 538§853§§u§5 2 =32 o§_ 538553 38% 2a.: 3 "Gov 5.2 .5: 23:3... .935 841 3- fl: 8.? 9:- «5. 8a. 3- E Fm- 2n .3. own- 3 9:. SN- 3. an. Ra- q». 96. 8a- an- a: $4- .3. <>m m2- 2- m>m m2- 2. >2. “2. 2. <2. 8.? «a- at. 3.9 8- . Sm an? 2- >6 3.... 2- a one. «.9 E 8.... 8. ma 2.9 v6. 5.. 8.9 2. 2... ed 2. E E. E. as an. :3 2 .5 $828. 85 3 E. 28 BEE mo «Susana. 9358 mm; fin HF Sea 5 afiaofiogusssgfififiu 9-3... .85 GL . a E. 93:2 xufiaaaafisg: .03. a? 82 223.3036 2 ~..N.~.-..--N\ :KCs \ 0~=~HNAV§€ :=:=-~ _ In: 1-1-11:14 111-1.--- 1...:. :4 J LC< bass“ 31 E.- c 835s ”is; 28 38.3.3 a 1n.- 3- a 2v 3- m «5&8 £2 flags” 3- S- 23. 253825 aaEEaEB-m .88: 3mg 3 m.— 5a EEq oESRE. monta- 2§m n33..- \§_m\<§n «2:33 fiazzfiga a... a 3 3222:; 2,52 2a 2.3 2 Sega—m 535398..» 832.. 2a a mud-.35 $2385 33.:- 5.24 «23 Susanaaéfi-a : 8.33 52> SE n83... u 6.» e2 52m 2.38::- § 2 «.95 2- :- $8 2 8 .54. . 29 9o 3. 083 26?: saga-852° 5%? 3 S- S. 92 8262 @526 3%.2 2&8 a». :- 3. 0886:..— co 8%? Egabafim foe .85 60V . .3.me I i .03. :2. $35: ‘ 9308.36 filmm- So 30538353§3 v.3- "Gus-2» <§E $.85: 13.33: {c 1-3.2: (323'... I h.- o!.«.«-.-. 1 r? mad mam-FR A..._....:r. Uriah-b... p. — .C L.~.3< \V.7..H\ r=uuhul h: rut-7:95 uvaul I... 1.3.: .2-Xu. _ A - VF. - I A. ch :n-huf‘ VéflL- 5.-~A.H.H¢kl\- s35 38 .52: 2. L‘ ad- NS- One usage E. fizz 2963253 3. S. mac—o 3- 2- <85 €3,853. 3 ma 3 ma RE 3.: mica 28 8.22: €52 - cine: E. SE 2 M3 4 23.2 033a 28.3.. 8%? 3 3 I- <80 8262 “.536 828.. 2&5: a». 3 ma 3 ma 8E 2o 858 330553 a. 553m Gov 5.2 858». a. w-MNW J32 a; 3- 29E 3 Anna 3. 5E 3:3qu «Fa—ca 88 33:05 AN 31 >mnE 2- {WE 28 as...“ 8. ma; 22.. bfififieuécaaaos : 2- :33 2- 53 "Snagsa £53 2. >93 5a 1:- u 6.9 5% <84 25982 a. . Lnflw- vam an .23.... 9.2 2 a _ mm. o.: an €55§5$E<899§~§ 2:- hfi an €53 0%;ng Z- c.: a u «33 Una-ERUPEQBE 05% no. to c u 533 033585802 29.. 53950508... 88.. 22.. 2a 2- 3- : u sag-323 :53 592.. 23235382.. 2. 3. an .EERQHRU 8255383 Ho 853 be th- 65 6b 5.2 8585." g 238 9 as»: 0395 a .o 8u§< a 8... 2n:- aafimoaoaa «3 g 53 525.. 9.8.2 as assuage an. n O.» Eh m8... 258: E. ICIIII‘EI ab>::..x=x..: 7.2. .2532: ”xv-Wax- \NNQNEHNQQQ 12.71332. 53 vanity—.49 tow—:5 by: :3:- NmuK-u- - -> Rao- 1.. RCC_-~..:= 97:3:25 hc «DEF—Wu A- Nth-V: “N n ~> NWM..~!\ v.r-.N- I «.9ov .2.F< 2~NC~2 Oat-«WIFSV .7.»- 39 ”6. 055.3 + «AM 4 84 3444448” 53844 3 45:24:44 «4 «£44844 «6. n?- 5> ohm- 48> 3.0. m3”: 84- m8? $8. m8< £35” 5224 «2. 82 Evanssgsusasifia 8d- 842 30305844? 8 54434433 488445 R. 4 - 949‘ 4:348 44 :54. mfinaagu 29: m4 mm- 4- ma¢< 28 44432.4 05 :4 44:on .8424 a «5.95 3.4.- mm}. 44nd :4- u 004 6.2 m4v< 258». E- 3 SN: swan-8524353.. +46 4.83 00: 38.2 8.388 05488. E L444. 484 4: 4 .484 mu. 4- Ann 4E4 4 4 44.5484 wed. 4mn4§4z>4§ 8.4V. 4444;42334 «Wm. 182%: 4 4> 54.44.80.408 05 8.4 5588 $6- 44 4 338442 4:44.844 28 482 58.5488 3 8.44 4mm 4En434 Ed. 444 4>\m84 3.4442388 98445::— wmd. 48435434 33424-6 38440 4.8ng 84. 45> 804 m4 Sousa:— 344442: .40 484444 44 2.41 ~44 44> menu an- n Gov EH< .4842 2.3895 v.4. . .. - i . - a filial. n. t. the“. NXNN 0:.EANCU by... )2: 4.4-2.4.4.7 4.3.9 A::»:&...-: c.3- - nhP‘CL. «8.?sz bvdwpuztfih ~§u7v= nurvAv- rib-IE A: Av‘ut— h: unavm-m-u-d< AV.’ fl ALUCV hvf,ph.<€ ghfl~c< b!\\n\\\ngkl\(\thk:lv\b. h‘ .4847: 39. 4.“ a:«anggmflmfiugmoamaafizae~34 3388 E4525? 2: an n42«8.448444453434848844834?444E434 3834835.. 98440253 3: 9m 44344438442343.4834“:8.44244 an. 45 mahamoamamommkoagfisaafiu .844» 4852:0255 .8442. my. 4.4 mcbamwonmaaamfiaaafi .5 4.. £32 4.5 25.4. =4 8.4m 2a a 2.. in 4§4§m§=m§~§ 34.23 3 mSNmEEmENE £25852. 885.4 5%: 3 44.4 fizmaafi was to 0852. 88549588414 .44.. «.4 .4484 084.4482» 4.4445 N234 Gov 8.4.44 6.46340 .3488 .m 434- 4W8 2. on: 2- £8. 3. 448.4. 3. one. .3 >34. n4 2%.. 9m .59 4.4. 484. 50850584435” 8&5” 5224 S "484. 68 24.39828 05.352883: 3 4.48.4 43.2%.“..8: 4.» u 604 8.2 4494. afigsafi .m becamfipspa 4mm... 828%?2244 8825 4§¢zvoaseq§§ 232.5%... .23 84.38.. 9.3434 «4 "6.4 8.2 m8? afiggud 0% an .4344 35355.3” 3:85.528 34. .4844 84.384.554.834 4.4443» S. .48» 44.3 aims 9a 4.434425554454544 2:. .484. 4%.4240899533: §.§§spa4o§§< an ugcnaqa m$< ”sagas 1 lil I- I‘lfil‘ NI)» - .4 IVAEQAI ~ IS; 53.0 K 31‘ A...-.a-.X:2: 1.7.:- P — file :5 .uov I.....p.-hrv--rv trU\.-.a-:-r — T: V A— N. n I Ace; EFFQ \ NI}; :5» :23». TIAofiytav» -m $7.225- \\.u...» LN 41 an :8 5.2- sgaéuaafifi 445 8.4 445 3.- $5 4.452.444.4425? 444.4485 4.384444 2 9 4.~ 43> 2 9 4.“ .4445 4442844 84.48 2.844444% 44.24 24547. 544544444 944444.. 54460 44 2 444.2 4: 8.42M .448 .m 4.64 «.6 24395444 4&4 «4 4.434442 34 44.4 22434 44.44 S 9434 2 3 9454 3 .44. S454 48¢» 4498445 a6 QN- 25.”— 2:22.38 42. 242844 044444.482: 4.44. 4.4- 34544444 9304444444855 344544 444.484.0544 4.4 «.44 $44444 448343 «4485444480 844044448 a6. WV. 24% 84.82.558.442 84442 444 2422.4 44.4 4.4- 4444444444452 TEomoSuaEm 044444448244 444 84.344443844440444. 44. 4d- ed. A4444 40 u «44.4 44.44 «.4 BaoToBusno ”8852444 44.43435“ 44.2- 344. 554.3446 T350469 "$4 a; 2242 82.448443444424444 44.44 to 4.323444 4% E- EngomoguE 445N425” an 3 04444 T 44945 u 84 933444 844 «4 44844444444409 344044448 n6 4mm mug..— T 0.4.4095 u 2 8884485 84434433243 4.4. «.4 4844 "34 uI HA8 Fights-Luzscc oh O‘RHQrF ‘ "“5“...- .~ -m. " 55 E5 :08 Ba 58,—. a vogue 83 .058 098% a wedge Eng 05 Bob pa chat 858038 538» gauge 83390 a vogue .0: "Q2 8 as 8m 3 8 Egg .5 omega: mm 3 92 mm we nagmezthEeua .m 8 8 a. 8 8 ”vignettes g omega—85 mm mm e. V mm 5 a8 .m amazemen— mwe 00V 602 a is 9 mm Beacon—22 . a 89 03%: has a, e egos—emu 083% Sago 53533me 3392.8.— 2: «e 2532.83 5E...» 3:53 2: 5t» fleece—Ema .322: 320.5% one—mu he eaten—ecu S 035—. 56 are specifically bound by proteins. It is not surprising, therefore, that numerous enzyme thermostabilities are altered by such effectors [298]. Substrate molecules have long been known to stabilize enzymes, presumably by interaction in specific binding sites [250,266], and these effects are not reviewed here. Chemical cross-linking or immobilization, glycosylation, inorganic salts, and high pressme also stabilize enzymes. This discussion of these topics is confined to the current mechanistic theories and provides representative examples supporting or opposing these theories. Glycosylation - An increasing number of bacterial extracellular enzymes have been shown to be glycosylated (for example, C. thermocellum cellulosome components and T. saccharolyn’cum endoxylanase [298a-300]). Most of these glycosylated enzymes retain their catalytic and stability properties when expressed in mesophilic hosts (not known to extensively glycosylate recombinant proteins). Also, a major source of thermal destabilization for the A. missouriensis glucose isomerase has been shown to be the glycosylation of a Lys residue by glucose - a glycoside as well as the glucose isomerase substrate [264]. While some glycosylated proteins have been shown to be more stable than their non-glycosylawd forms [199,301], glycosylation is not a thermostabilization method commonly found in nature. salts - Inorganic salts stabilize proteins by either a specific salt effect, where a metal ion interacts with the protein in a conformational manner or a general salt effect, which mainly affects water activity. Ca2+ binding by ct-Amylases is an example of the former mechanism. The tat-amylase catalytic site is located in a cleft between two domains (an [oz/[313 barrel and a large loop). Coordinated by ligands belonging to these two domains, Ca2+ is essential for the a—amylase's catalytic activity and thermostability [302]. Xylose isomerases bind two metal ions (either C02+, Mg2+, or W”). One cation is directly involved in catalysis (the catalytic metal) and the second stabilizes the folded protein (the structural metal) [303,304]. The two metal-binding sites can have different specificities, and replacing one cation with another often significantly alters enzyme acuity, subsm ijsoz'tme was ] protein. relativt geeifrc salt eff size, coordinati A Study K3PO4, N33PC reduce the enzy (1837:35ng cffa mutation is t 0““ Wipitatt Convenient Way mosiabinry Whilt the five 6 varied from en; in the PTCSCDCC . $13me by 0.1 1110:: efficiently W“: effw gamed by the i [>1 H pOIaSSiLm 57 activity, substrate specificity, and thermostability [303,305]. Ca2+ stabilization of lysozyme was proposed to be due to a decrease in the entropy of the Ca2+-bound unfolded protein, relative to the entropy of the unfolded protein in the absence of Ca2+ [298]. This specific salt effect was attributed to the characteristics of the specific binding sites (i.e., size, coordination number, and the nature of the liganding residues) [65]. A study of GAPDH thermostabilization by salt indicated that the relative effects of K3PO4, Na3PO4, K2SO3, NaZSO3, KCl, and NaCl were consistent with their abilities to reduce the enzyme solubility in aqueous solvent. Their action was attributed to their decreasing effect on water activity, a general salt effect [76]. Ammonium sulfate protein precipitation is a common application of the general salt effect for enzyme precipitation. Once precipitated, enzymes are typically also stabilized. This procedure is a standard, convenient way to store enzymes. Thauer and colleagues studied the effect of salts on the thermostability and activity of five M. kandleri methanogenic enzymes [153-155,164,183]. While the five enzymes are activated and stabilized by salts, the extent of the salt effect varied from enzyme to enzyme. CHO—H4MPT formyltransferase was optimally stabilized in the presence of 1.5 M K2HP04 while the Ego-dependent CI-12=H4MPT reductase was stabilized by 0.1 M KzHPO4. K+ and NH4+ cations typically improve enzyme stabilities more efficiently than other cations. Of all the anions, 8042' and HPO42' had the strongest activating effect [164]. Enzyme stabilizing salt requirements however, are not always satisfied by the intracellular salt concentration. M. kandleri's intracellular salt concentration (>1 M potassium plus 1.2 M 2,3-diphosphoglycerate) [164] seems to favor CH2=H4MPT reductase activity (optimal at 2.0-2.5 M salt), but this intracellular salt concentration is far from being the optimal enzyme stabilizing concentration (optimal between 0.1 and 1.5 M salt) [155]. These studies illustrate the variety of situations encountered: (i) similar enzymes from different organisms do not have the same stabilization and activation requirements, (ii) enzymes from the same organism are not equally affected by an tiiirnmntrti‘i man‘s int: lithinnihtsis httnttn charge. (13231306). ins:- protcin desrabil: the interface btt‘ This hrpothesis I sat concentratin M) I‘Oquired to 5 Mar uncharget mtSOphiiic em "“33" Wrens: “lth the less pt than the lime” {hm PmdiCt: “Kit Partially Other dminutes bi Mammy!“l ”Elam to r ”gamism ; know} [0 stab mmmues ( Wins [3071 mitinp HOw 58 environmental (i.e., cellular) factor, and (iii) thermophilicity and thermostability are not necessarily controlled or favored by the same factors. The hypothesis that the general salt effect was due to direct electrostatic interactions between charged amino acids and salt counterions was not corroborated by experimental data [306]. Instead, a smeared charge repulsion mechanism was proposed: Salt-induced protein destabilization was due to nonspecific charge repulsion and to ionization changes at the interface between complementary peptide smfaces in the partially unfolded protein. This hypothesis agrees with the well-documented protein stabilizing effect of intermediate salt concentrations, in particular with the relatively high salt concentrations (1.0 M to 2.0 M) required to stabilize some thermophilic enzymes [306]. It also agrees with the reduced polar uncharged residue content in some thermophilic enzymes, when compared to their mesophilic counterparts [85,230]. At high temperatures, the system‘s higher kinetic energy increases the potential for exposing buried surfaces to the solvent. Salt interactions with the less polar residues in the partially unfolded thermophilic enzyme are less favorable than the interactions with the more polar mesophilic enzyme residues. Furthermore, this theory predicts that greater solvent ionic character will stabilize folded proteins relative to their partially unfolded forms. Other chemical effectors - Breitung er al. (1992) [164] found that the differences between M. kandleri, M. thermoautotrophicum, Archaeoglobus fidgidus, and Methanosarcina barkeri CHO—H4MPT formyltransferase activation by salts were directly correlated to the intracellular 2,3-diphosphoglycerate concentration in the different organisms. Polyalcohols such as glycerol, sorbitol, mannitol, sucrose, and starch and are known to stabilize proteins [306]. Hyperthermophilic bacterial cytoplasms contain thernmmines (polyamines) that have been shown to thermostabilize nucleic acids and proteins [307]. Organic polymers such as polyethylene glycol (PEG) also stabilize folded proteins. However, they can also induce precipitation and even inactivation. Pressu reported [30M crysulhne solid Unfolding a pro Stable than the i compact , foide pressure has on thermophiles h; from 200 to 401 from organism Nations are pf at high pressurt MOleClllar m 59 Pressure effects - The ability of high pressure to stabilize proteins has been reported [308,309]. The observation that folded proteins have densities similar to crystalline solids is the theoretical basis for pressure as a general stabilizing force. Unfolding a protein would increase its volume, making a partially unfolded protein less stable than the folded protein at increased pressures. Thus, at higher pressures, the compact , folded protein form is more stable. Hei and Clark [308] have studied the effect pressure has on enzymes from barophilic organisms. They showed that some barophilic thermophiles have enzymes which are also barophilic (i.e., optimal activity at pressures from 200 to 400 atrn) [308]. Hei and Clark also characterized barOphilic enzymes isolated from organisms that were not, themselves, barophilic [308]. Since numerous chemical reactions are performed at high temperatures and pressures, demonstrating enzyme stability at high pressmes is potentially important for enzyme biotechnological applications [309]. Molecular mechanisms of protein thermophilicity ‘ The molecular basis for enzyme thermophilicity is not well understood, and the literature on this subject is limited. Linear Arrhenius p10ts for numerous enzymes indicate that if there are structural changes in the catalyst throughout the active temperature range, they are either not catalytically significantor they are offsetting. The discontinuous Arrhenius curves observed for catalysis by some enzymes [82,145,190] indicate that, in these specific cases, temperature dependent, functionally significant structural changes may occur. These discontinuous Arrhenius ewes were believed to be characteristic of thermophilic enzymes, but, as mentioned previously, not all thermophilic enzyme Arrhenius plots are discontinuous and some mesophilic enzyme Arrhenius plots are also discontinuous [92]. Explanations for Arrhenius discontinuities have been proposed that do not involve altering the catalyst’s structure (e.g., change in the slow step of the reaction) [207 ,208]. Generally, enzyme activity in its active temperature range is well described by the Arrhenius equation, and can be derived from the temperatme for maximal activity, magnitude of equation, hovt Thus, an enzy the theoretical protein unfold independend) Most : thermophiiieit thtrtrrostabiiit Whig site if. inmascd its t] d” abStnce or and militant en Minty of the Minty 00nd; its Own tempt enlime therm thermophilicit iMimi)" of 2 t 60 magnitude of maximal activity, and the activation energy for the reaction. The Arrhenius equation, however, cannot predict the optimal temperature for an enzymatic reaction. Thus, an enzyme’s active temperature range appears to be determined at the lower end by the theoretical limitation on activity determined by the Arrhenius relationship. Partial protein unfolding or covalent chemical modifications appear to determine the upper limit, independently from the Arrhenius equation. Most studies on enzyme thermostabilization do not comment on the enzyme thermophilicity. Only a few reports [250,310] describe mutations that altered both enzyme thermostability and thermophilicity. Kuroki et al. [250] demonstrated that adding a Ca2+ binding site in human lysozyme by site directed mutagenesis both stabilized the protein and increased its thermophilicity from 70°C (native enzyme) to 80°C (mutant holoenzyme). In the absence of Ca2+, the mutant enzyme thermophilicity was reduced to 65°C. The native and mutant enzyme catalytic rates were superimposable up to the temperature for maximal activity of the wild-type enzyme (~70°C). Above that temperature the mutant enzyme activity continued to increase, with the same apparent Arrhenius behavior, until it reached its own temperature for maximal activity. The mutant behavior suggests that the wild-type enzyme thermophilicity was limited by its thermostability, and that the higher thermophilicity of the mutant was due to an increase in its thermostability. In one study, insertion of 2 to 4 amino acids into Caldicellulosiruptor saccharolyticus xylanase generated stability-only mutants and stability-plus-optimal temperature-altering mutants [310]. Insertion of a single Pro-Arg sequence reduced the enzyme's specific activity more than 2- fold, the thermostability 4-fold, and the optimal temperature by 20°C. These mutations support the hypothesis that, in this case, thermophilicity was also limiwd by folded protein stability. Other C. saccharolyticus xylanase mutants, displayed reduced activity and reduced stability (similar to the Pro-Arg mutant) while retaining native-like thermophilicity (70°C), suggested that the molecular bases for protein thermostability and thermophilicity can be structurally distinct. The Tomazic and Klibanov thermostability model (scheme (1), 61 and [197]) can potentially explain the observed difference between the two classes of C. saccharotyticum xylanase mutants. The mutations which only reduce enzyme thermostability would increase the rate of scrambled structure formation from the enzyme non-native state, while the mutations that reduce both thermophilicity and thermostability would result fi'om the shift of native to nonnative enzyme at lower temperatures. This analysis predicts that a class of thermophilicity-reducing mutants that do not alter thermostability could be generated if the normative state (resulting from reversible unfolding) were stabilized with respect to the scrambled structure (the enzymes would only reversibly inactivate). While the relationship between enzyme kinetics and protein thermophilicity] thermostability remains unclear, it has been proposed that the poor activity of thermozymes at low temperatures is the result of excessive rigidity. This rigidity is believed to be necessary to maintain active enzyme architecture at high temperatures [198,203,215]. Data from crystallography, deuterium exchange, proteolytic susceptibility, and other experimental approaches have demonstrated that folded thermozymes are indeed more rigid at low temperatures than their mesophilic counterparts [198,203,215]. Tchemajenko er al. [206], however, reported linear Arrhenius plots for the four xylose isomerases they studied. Since these plots were perfectly fit by the Arrhenius equation, increased low- temperature protein rigidity was not required to explain the poor low-temperature activity of thermophilic enzymes [206]. The temperatme dependence of thermophilic enzyme activity might, instead, only be determined by temperature-dependent substrate kinetic energy variations. Therefore, determining the importance of protein flexibility to thermophilic enzyme activity will require comparisons of corresponding thermophilic and mesophilic enzyme flexibilities and activities throughout their active temperature ranges. 62 Genetic engineering of thermozymes Modification of enzyme catalytic properties The accumulating protein structural data justifies the use of enzyme engineering as a rational research approach to make enzyme catalytic properties fit industrial processes. Knowing the L-lactate dehydrogenase structure and catalytic mechanism, Holbrook and colleagues attempted to systematically modify the B. stearothennophilus L-lactate dehydrogenas into a malate dehydrogenase by a point mutation proposed to better accommodate oxaloacetate in the substrate pocket [311], then to design a non-specific a- hydroxy acid dehydrogenase from the lactate dehydrogenase [312]. This non-specific a- hydroxy acid dehydrogenase mutant enzyme remained thermostable, and could be used for various chemical syntheses. A similar approach was used by Meng et al. [313] to switch the substrate preference of T. thermosulfun’genes xylose isomerase from xylose to glucose. Arnold and colleagues used PCR-mdiated random mutagenesis to enhance subtilisin B activity in polar organic solvents [313a,313b]. Using several screening steps, they eventually selected a multiple mutant that was 256 times more efficient than the wild- type enzyme in 60% dimethylformamide. A similar PCR based technique, was successfully employed to enhance the Z. mobilis 1° ADH thermostability [217]. These studies provide proof of principle for the general application of directed enzyme design strategies using mutagenesis. Thermostability and thermophilicity engineering Research has shown that while thermophilicity and thermostability can be altered independently, they are often structtn'ally related. Numerous molecular interactions might be responsible for these properties, however, only a few are of genetic engineering value: (i) protein sequences contain redundant information for proper folding, and protein structures can often accommodate amino acid substitutions without significantly altering the catalytic efficiency; (ii) a small number of amino acid substitutions throughout the protein can sigru‘frem protein are in ones) is critic stabil‘zation i sufaee 100p 5 analysis has it tremophilic ; 311mg protei hehces by cap same engineer successfully sr methods (eg. H dinnhona] Str identified (e. g., inc“mined 1. MIME Pepi applications (>9 mm Start 63 can significantly alter its thermophilicity and thermostability; and (iii) certain regions in a protein are more labile than others—stabilizing these regions (rather than the more stable ones) is critical to improve protein thermal properties. Four general strategies for thermal stabilization have been identified: more efficient protein core packing, tit-helix stabilization, surface loop stabilization, and prevention of chemical degradation. Recent crystallographic analysis has indicated that more-rigid core structure (lower Brvalues) is a hallmark of thermophilic proteins [124,198]. However, systematic engineering of core rigidity without altering protein function is beyond the capabilities of current technology. Stabilizing cr- helices by capping or by introducing alanines has had spotty success, and suffers from the same engineering problems as do the core packing strategies. Loop engineering has successfully stabilized proteins [233,243,248254258,259,272]. Sequence analysis methods (eg. HCA) now reliably predict surface loop regions, even in the absence of 3- dimensional structural information. Specific stabilizing amino acid substitutions have been identified (e.g., substitution of surface lysines with arginines and introduction of prolines in constrained loops, and salt bridges), making this approach a good engineering tool. Preventing peptide chemical degradation is also important for very high temperature applications (>90°C). Trends toward reduced Cys, Asn, and Gln contents in hyperthermophilic enzymes have been identified in nature [303], and deamidation has been shown to be a main factor responsible for irreversible enzyme inactivation [197]. The appropriate strategy for engineering protein thermostability or thermophilicity depends on the project goal, on the available structural information, and on the thermal mechanisms limiting enzyme stability or activity (Fig. 3). Selecting an initial enzyme with thermal properties as close to the target as possible clearly increases the chances of successful engineering. The probability of engineering success is further enhanced by having the initial enzyme's high-resolution 3-dimensional structtne as well as a protein in the same class that has thermal characteristics close to the desired protein. Comparing thermophilicity and thermostability properties indicates whether the limiting step in protein Figure 3. Flow chart of potential steps toward engineering enzyme thermophilicity and thermostability. 65 Design goal Thermophilicity Thamostability Reversible protein inactivtion Irreversible protein inactinipn (Thermostability > Thermophilicity) (Then-nostabrlrty = 'I'hmnovhlllcrty) N0 precipitation Precipitation 0?me Precipitate Precipitate active inactive Prevent structural clung ' Thor-mo bilicity General: 1) Random mutagenesis with selection 2) sequence comparison directed mutagenesis 3) substitute Pro residues into loops 4) replace Lys with Arg in loops 5) chemically aosslink V ' 6) immobilize 1) Immobilize Complete Partial 7) add stabilizing chemicals to the system 2) Chemically modify a) protein specific a) succinylate Lys residues i) salts (N (:1, KCl. etc.) b) intramolecular croaslink ii) substrates/effectors 3) add polyalcohols to system b) reduce solvent entropy i) glycerol ii) Hofmeister series ions 1) add divalent cation binding site 2) reduce bond strain by mutagenesis 3) till internal cavities by mutagenesis 4) stabilize helices i) 11le to cap helix C-terminus ii) Glu/ASP to cap helix N-tcrminus iii) Ala to stabilize helix v 0 C IV I 1) addGlytoloops WW 1) reduce themunber ofAsn. Gln. and Cys residues 66 inactivation is the initial partial unfolding or ineversible structural changes (such as precipitation or chemical degradation). If, in the temperature range of interest, the protein thermostability half-fife is longer than the activity half-life, the reversibly partially unfolded protein is not immdiately irreversibly inactivated If, however, these two half-lives are similar, irreversible inactivation follows the initial unfolding almost instantly. The protein concentration in the target application should be estimated based on the protein function (receptor, enzyme, antibody, etc.), to provide a concentration range for these characterizations. Determination of the thermostability half-life at different initial protein concentrations within this range will indicate if inactivation is intramolecular (concentration independent) under these specific conditions. Having increased enzyme thermostability allows applications of repeated thermal cycling (e.g., PCR), as well as permitting pre-application high-temperature processing steps (e.g., processing enzymes in food or feed additives). To increase enzyme thermostability, engineering efforts should focus on preventing irreversible protein inactivation. Strategies to prevent protein aggregation (e.g., chemical modification of lysine residues with succinic acid [197], addition of polyalcohols to the solution [306], or immobilization on an inert matrix) might significantly thermostabilize the protein. This approach would need to be undertaken if (i) enzyme precipitation coincides with inactivation and the precipitated protein does not retain complete activity, or (ii) if aggregation is undesirable for the target application. If these measures are insufficient or inappropriate for the design goal, then prevention of inactivating conformational changes through peptide modifications such as intramolecular cross-linking or through system additives (cg. salt or glycerol) that stabilize the folded protein structure should be examined. Salts have been shown to stabilize proteins through specific binding, charge shielding, and modification of the system entropy. The addition of organic chemicals such as glycerol and polyethyleneglycol has also been shown to stabilize some proteins, but they often reduce activity and can destabilize some proteins. The specificity of intramolecular 67 cross-linking is difficult to control and often reduces protein activity but it covalently tethers protein regions analogously to disulfide linkages. Either random or site directed mutagenesis of the initial protein may be used to thermostabilize proteins. Random mutagenesis has been successfully used to increase enzyme thermostability and resistance to solvent-induced denaturation. Hageman and colleagues [314] developed a system in which a gene encoding a mesophilic enzyme is introduced into a thermophilic host, and variant enzymes with increased thermostability are selected during growth at increasing temperatures. Heat-stable kanamycin nucleotidyltransferase mutants were obtained using a B. stearothennophilus strain that expressed a mesophilic kanamycin nucleotidyltransferase to select kanamycin resistant variants at increasing temperatures [239]. Although attractive at first glance, this approach is limited by the absence of genetic tools (i.e., shuttle vectors, transformation methods) available for most thermophiles and hyperthermophiles, as well as by the limited enzymatic activities directly selectable in B. stearothennophilus cultures. The cunent lack of thermophilic cloning hosts also makes classic genetic complementation by thermophilicity mutants a significant challenge. Practical use of random mutagenesis requires a powerful mutant selection procedure or specific structural information (to limit the size of the target thus, reducing the total number of possible mutants to be tested). Site-directed mutagenesis usually requires protein structural information. Strategies that include substitutions with residues present in structurally similar proteins with thermal properties similar to the desired mutant require sequence information from both proteins. If an enzyme targeted for hyperthermophilic applications (>80°C) is irreversibly inactivated, reduction of the number of noncatalytic Gln, Asn, and Cys residues can be used to stabilize the protein structme, using the single peptide sequence. Loop stabilization by introducing Pro or Arg residues also requires only knowing the protein sequence. Exact structural information from crystallography, NMR, or homology modeling allows the engineering of multi—residue motifs such as metal binding sites or disulfide bridges (for moderate 68 temperature stabilization) to stabilize the enzyme structure, and allows specifically tailored design schemes such as cavity filling mutations, helix stabilization by alanine insertions or capping (eg. His at the c-terminus), and alleviation of specific bond strains. Deuterium exchange, measured by NMR, can precisely identify the more labile protein regions to stabilize [315]. Muheim et al. [316] created a dihistidine metal-chelating site on a surface B—sheet of cytochrome c, and cross-linked it with the metal complex Run(2,2'-bipyridine). This cross-linking increased the meltingltemperature of the mutant enzyme by more than 23°C. Increasing enzyme thermophilicity requires preventing thermally induced unfolding. Since enzyme thermophilicity is often limited by thermostability, engineering thermophilicity usually requires enzyme thermostabilization. Also note that, for an enzyme that precipitates upon hearing, if precipitation is not the first inactivation step, preventing aggregation is unlikely to enhance thermophilicity. THESIS OBJECTIVES AND SIGNIFICANCE Because archaeal and bacterial thermophiles were the first organisms to have evolved the study of these organisms and their enzymes is expected to provide insights into the origins of biocatalysis on Earth [317]. With their remarkable thermostability and activity, thermophilic enzymes provide a model for studying protein thermostability and biocatalysis at high temperatures. Structurally similar to their mesophilic counterparts, thermophilic enzymes are powerful tools for developing enzyme structural or functional models through comparative analysis. Their high-temperature activity and high stability make thermophilic enzymes excellent candidates for industrial enzymatic applications. Therefore, the fundamental scientific examination of thermophilic life may also yield novel catalysts of tremendous biotechnological value. 69 Numerous highly similar thermophilic 2° ADHs have been isolated from thermoanaerobic bacteria [7,8,15,16] but despite more that a decade of research since the first enzymes were purified in 1981 [14,15], no report of the cloning of a gene encoding a thermophilic 2° ADH has been published. There is also no record of either mesophilic or thermophilic 2° ADH characterization by mutagenesis. The only thermophilic 2° ADH peptide sequence avaliable was determined by Edman degradation [33]. Alignment of the liver 1° ADH and the T. brockii 2° ADH peptides by Peretz and Burstein [33] suggested that the structural Zn binding loop was absent in the 2° ADH, but they concluded that insufficient similarity existed between the A. eutrophus, horse liver and Thennoanaerobacten'wn brocla'i enzymes for further comparative analysis. Earlier, Lamed and Zeikus reported that T. brockii 2° ADH activity was both cysteine and Zn dependent [l5] and Bryant and Ljungdahl measured Zn specifically bound to the T. ethanolicus (ATCC 31550) enzyme [14]. All thermophilic 2° ADHs identified are tetrameric with reported subunit molecular masses near 40 kDa. This coincidence of enzyme structural properties and the similarities seen in kinetic comparisons of these NADP(H) linked 2° ADHs led N agata er al. to conclude that all thermophilic 2° ADHs were extremely structurally similar [16]. The greater than 75% sequence identity between the mesophilic C. beijerinckii and thermophilic T. brockii peptides further argues that all catabolic 2° ADHs may share significant 3-dimensional structural identity. These thermophilic and mesophilic 2° ADH peptide sequences therefore, make an excellent system for determination of protein thermostability structure-function relationships through comparative analysis. However, reciprocal mutagenesis experiments to verify the sequence based predictions require expressed clones of the genes encoding both proteins. The T. brockii 2° ADH has been demonstrated to reduce aldehydes and ketones on aliphatic molecules containing up to 10 carbons [318]. This enzyme requires ketones on the second or third carbon, displaying higher activity toward unbranched noncyclic molecules. The enantiospecificity was lower and (R)— specific in the reduction of very Cb: DOLE: 70 short ketones but greater than 90% and (S)- specific for reduction of ketones with aliphatic chains greater than 5 carbons in length. The T. brockii 2° ADH, like the horse liver and yeast 1° ADHs, produces (8)-alcohols using the E3 pathway [30]. For the T. brockii 2° ADH, lower reaction temperatures also yielded higher enantiomeric excess [319]. Enantiomeric excess at equillibrium is a function of the Gibbs energy ratio between the (R)- and (S)- specific reactions and this relationship indicated that the variation of T. brockii 2° ADH reaction stereospecificity with temperatme resulted mainly from entropy considerations. An ordered kinetic mechanism with cofactor binding followed by substrate binding, hydride transfer, product release, then cofactor dissociation, similar to the liver 1° ADH mechanism, was identified for T. brockii 2° ADH catalysis [320]. The T. brockii enzyme is both thermostable and thermophilic [15]. Its optimal temperature for activity was reported to be above 70°C with significant inactivation rates occrn'ring at temperatures above 90°C. A discontinuity in the Arrhenius plot for propan-2—ol and butan-2-ol oxidation by this enzyme was reported in the same study. The T. brockii 2°ADH has already been used in the analytical scale production of a chiral constituent of civet (a perfume additive) [321] and of an insect pheromone [322]. Its potential value for other syntheses has also been examined [323]. Thermophilic 2° ADHs have high temperature optima (80 to 95 °C), high thermostability, and reduced sensitivity to oxygen (they are only reversibly inactivated, unlike 1°ADI-ls) [8]. The extent of T. brockii 2°ADH stability in the presence of organic solvent is still controversial. In the original 1981 study [323] the enzyme remained 80% active after 15 min at 52°C in the presence of 40% propanol-Z, whereas it was reviewed by Cowan [59] to rapidly inactivate at temperatures above 45°C and in the presence of 10% organic solvent. T. brockr'i enzyme catalysis was determined in vitro in supercritical 002 indicating that it is active in nonaqueous conditions under high pressure [324]. The T. brockz'i enzyme was also active toward both highly water soluble ketones and substrates with low water-solubility in a range of water-oil emulsions containing either cationic or neutral surfactants [325]. A tried Q flint 7 1 range of potential chiral transformations has already been demonstrated for ADI-Is. 2° ADHs, optimized to interconvert ketones and 2° alcohols, are better suited to the production of chiral alcohols that 1° ADHs, optimized to interconvert aldehydes and 1° alcohols. Therefore, continued research into the molecular mechanisms responsible for 2° ADH thermostability, thermophilicity, and substrate specificity will advance our understanding of these general enzyme properties specifically using an enzyme with potential as an industrial chiral catalyst. The research objectives presented in this thesis advances knowledge on thermophilic 2° ADH 2° ADH structure-function relationships in 5 ways. First, cloning and characterizing the gene encoding the T. ethanolicus 39E 2° ADH will make mutagenic analysis of thermOphilic 2° ADH structure-function relationships possible. Second, stable overexpression of the properly active recombinant thermophilic 2° ADH in E. coli will provide sufficient quantities of easily pmified protein for structural analysis. Third, characterization of enzyme thermostability and thermophilicity will allow sequence comparisons between this thermophilic enzyme and the similar mesophilic enzyme from C. beijerinckii for identification of 2° ADH thermal property molecular determinants. Fourth, site directed mutagenesis and biophysical analysis will test the predictions of 2° ADH catalytic architecture based on comparison to those documented for 1° ADHs. Finally, the biochemical mechanism of thermoinactivation and the forces stabilizing active the T. ethanolicus 2° ADH structure will be elucidawd. ACKNOWLEDGMENTS I gratefully acknowledge Dr. Claire Vieille for her part in writing and editing the original manuscript. I also am grateful to Dr. Claire Vieille, Dr. Vladimir Tchemajenko, Dr. Mariet Van dc Werf, and Maris Laivenieks for their helpful discussions. I acknowledge Mr. Chris Jambor for critical review of parts of the original manuscript. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 72 REFERENCES Branden, C.-I., Jomvall, H., Eklund, H. and B. Fururgren (1975) in The enzymes, 3rd ed., vol. XI part A" (P. D. Boyer, ed.) pp. 103-190. Academic Press, New York. Gottschalk G. (1986) Bacterial metabolism, 2nd edition. Springer-Verlag, New York. Schimpfessel, L. (1967) Arch. Int. Physiol. Biochim. 75, 368-369 Schimpfessel, L. (1968) Biochim. Biophys. Acta 151, 386—?. Lutstorf, U. and Megnet, R. (1968) Arch. Biochem. Biophys. 126, 933-944 Sund, H. and Theorell, H. (1963) in The enzymes, 2051 ed., vol. 7, (Boyer, P D., Lardy, H. and Myrback, K., ed.), pp. 25-83 Academic Press, New York. Bryant, F. 0., Weigel, J. and Ljungdahl, L. G. (1988) Appl. Environ. Microbiol. 54, 460-465 Burdettc, D. S. and Zeikus, J. G. (1994) Biochem. J. 302, 163-170 Kolb, E. and Harris, J. I. (1971) Biochem J. 124, 76P-77P Rella, R., Raia, C. A., Pensa, M., Pisani, F. M., Gambacorta, A., De Rosa, M., and Rossi, M. (1987) Eur. J. Biochem. 107, 475-479 Ma, K., Robb, F. T., and Adams, M. w. w. (1994)App1. Environ. Microbiol. 60, 562-568 Ismaiel, A. A., Zhu, C.-X., Colby, G. D. and Chen, J .-S. (1993) J. Bacteriol. 175, 5097—5105 Steinbfichel, A. and Schlegel, H. G. (1984) Eur. J. Biochem. 141, 555-564 Bryant, F. 0. and Ljungdahl, L. G. (1981) Biochem. Biophys. Res. Corn. 100, 793-799 Lamed R. J. and Zeikus, J. G. (1981) Biochem. J. 195, 183-190 Nagata, Y., Maeda, K. and Scopes, R K. (1992) Bioseparation 2, 353-362 Neale, A. D., Scopes, R. K., Kelly, J. M., and Wettenhall, R. E. H. (1986) Eur. J. Biochem. 154, 119-124 Bleicher, K. and Winter, J. (1991) Eur. J. Biochem. 200, 43-51 Ainsley, G. R. Jr. and Cleland, W. W. (1972) J. Biol. Chem. 247, 946-951 Brooks, R. L. and J. D. Shore (1971) Biochemistry 10, 3855-3858 41. <2. 21. 22. 23. 24. 25. 26. 27. 28. 29. 31. 30. 32. 33. 34. 35. 36. 37. 38. 38a. 39. 40. 41. 42. 73 Bush, K., Shiner, V. J ., and H. R. Mahler (1973) Biochemistry 12, 4802-4805 Beijer, N. A., Buck, H. M., Sluyterman, L. A. AE., and Meijer, E. M. (1990) Biochimica Biophysica Acta 1039, 227-233 Dunn, M. F. and Hutchison, J. S. (1973) Biochemistry 12:4882-4892 Jacobs, J. W., McFarland, Wainer, I., Jeanmaier, D., Ham, C., Hamm, K., Wnuk, M., and Lam, M. (1974) Biochemistry 13, 60-64 McFarland, Chu, Y.-H., and J. W. Jacobs (1974) Biochemistry 13, 65-69 Eklund, H., Nordstrbm, B., Zeppezauer, E., Soderlund, G., Ohlsson, I., Boiwe, T. and Branden, C.-I. (1974) FEBS Lett. 44, 200-204 Rossmann, M. G. Moras, D., and K. W. Olsen (1974) Nature 250, 194-199 Wierenga, R. K. and H01, G. J. (1983) Nature 302, 842-844 Prelog, V. (1964) Pure Appl. Chem 9, 119-130 Bradshaw, C. W., Hummel, W. and Wong, C.-H. (1992) J. Organic Chem. 57, 1533-1536 Bradshaw, C. W., Shen, J.-G. and Wong, C.-H. (1992) J. Organic Chem. 57, 1526-1532 Jendrossek, D., Steinbuechel, A. and Schlegel, H. G. (1988) J. Bacteriol. 170, 5248-5256 Peretz, M. and Y. Burstein (1989) Biochemistry 28, 6549-6555 Kullmann, W. (1987) CRC Press, Inc., Boca Raton, FL Cheetham, P. S. J. (1993) Trends Biotechnol. 11, 478-488 Dordick, J. S. (1992) Trends Biotechnol. 10, 287-293 Faber, K. and Franssen, M.C.R. (1993) Trends Biotechnol. 11, 461-470 Jones, J. B. (1986) Tetrahedron 42, 3351-3403 Sarney, D. B. and Vulfson, E. N. (1995) Trends Biotechnol. 13, 164-172 Zong, M.—H., Fukui, T., Kawamoto, T. and Tanaka, A. (1991) Appl. Microbiol. Biotechnol. 36, 40-43 Fuganti, C. and Grasselli, P. (1985) Tetrahedron Let. 26, 101-104 217%? K., Inuoe, K., Ushio, K., Oka, s. and Ohno, A. (1987) Chem. let. 4, Trincone, A., Lama, L., Lanzotti, V., Nicolaus, B., De Rosa, M., Rossi, M. and Gambacorta, A. (1990) Biotech. Bioeng. 35, 559-564 LII irl . (I 61 63. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 74 Hummel W. (1990)App1. Microbiol. Biotechnol. 34, 15-19 Whitesides, G. M. and Wong, C.-H. (1983) Aldrichimica Acta 16, 27-34 Rosazza, J. P. N. (1995) Amer. Soc. Microbiol. News 61, 241-245 Meyer, H.-P. (1991) Biol. Ferm. Eng. 8, 602-606 Wong, C.-H. (1989) Science 244, 1145-1152 Irwin, J. B., Lok, K. P., Huang K. W. C. and Jones, J. B. (1978) J. Chem. Soc. Perkin I 12, 1636-1641 Bradshaw, C. W., Hummel, W. and Wong, C.-H. (1992) J. Org. Chem. 57, 1532-1536 Bradshaw, C. W., Fu, H., Shen, G.-J. and Wong, C.-H. (1992) J. Org. Chem. 57, 1526-1532 Hummel, W. and Kukla, M. R. (1989) Eur. J. Biochem. 184, 1-13 May, S. W. and Padgette, S. R. (1983) Bio/Technology 1, 677-686 Persson, M., MAnsson, M.-O., Biilow, L. and Mosbach, K (1991) Biotechnology 9, 280-284 Klibanov, M. (1983) Proc. Biochem. 18, 13-23 Eeg-Rghlgcgbk, Fauve, A., Renard, M. F. and Veschambre, H. ( 1992) Biocatalysis Zaks, A. and Klibanov, A. M. (1988) J. Biol. Chem. 263, 8017-8021 Klibanov, A. M. (1989) Trends Biol. Sci. 14, 141-144 231911213, J. R, Lee, K. M. and Biellmann, J. F. (1987) Eur. J. Biochem. 163, Cowan, DA. (1992) Trends Biotechnol. 10, 315-323 Lee, Y.-E., Lowe, S. E. and Zeikus, J. G. (1993) Appl. Environ. Microbiol. 59, 3134—3137 Lehmacher, A. and Hensel, R. (1994) Mol. Gen. Genet. 242, 163-168 Lee, C., Bhatnagar, L., Saha, B. C., Lee, Y.-E., Takagi, M., Imanaka, T., Eggasarian, M. and Zeikus, J. G. (1990) Appl. Environ. Microbiol. 56, 2638- Zeikus, J. G. and Brock, T. D. (1971) Biochim. Biophys. Acta 228, 736-745 Adams, M. W. W. (1993) Ann. Rev. Microbiol. 47, 627-658 so. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 75 Coolbear, T., Whittaker, J .M. and Daniel, RM. (1992) Biochem. J. 287, 367- 374 Daniel, R. M. (1992) Origins Life Evol. Biosphere 22, 33-42 Lowe, S. E., Jain, M. K., and Zeikus, J. G. (1993) Microbial. Rev. 57, 451-509 Stetter, K. 0., Fiala, G., Huber, G., Huber, R., and Segerer, A. ( 1990) FEMS Microbiol. Rev. 75, 117-124 Wiegel, J. and Ljungdahl, L. G. (1984) CRC Crit. Rev. Biotechnol. 3, 39-108 Coolbear, T., Daniel, R. M. and Morgan, H. W. (1992) Adv. Biochem. Eng.- Biatechnal. 45, 57-98 Vieille, C., Burdctte, D. S. and Zeikus, J. G. (1996) in Biotech. Ann. Rev., vol. 2, (M. R. E1 Geweley, ed.) pp. 1-83. Elsevier, Amsterdam, Neth. Berquist, P. L. and Morgan, H. W. (1992) in Molecular biology and Biotechnology of extremaphiles, (Herbert, R. A. and Sharp, R. J., eds.), pp. 44- 75. Chapman and Hall, NY Winter, J ., Lerp, C., Zabel, H.-P., Wildenauer, F. X., Kanig, W. H., and Schindler, F. (1984) System. Appl. Microbial. 5, 457-466 Schmitz, R.A., Richter, M., Linder, D. and Thauer, R.K. (1992) Eur. J. Biochem. 207, 559-565 O'Brien, W. E., Brewer, J. M. and Ljungdahl, L. G. (1973) J. Biol. Chem. 248, 403-408 Hatchikian, E. C., and Zeikus, J. G. (1983) J. Bacterial. 153, 1211-1220 Benachenhau-Lahfa, N., Labedan, B. and Farterre, P. (1994) Gene 140, 17-24 Lacher, K. and Schafer, G. (1993) Arch. Biochem. Biophys. 302, 391-397 35mg £16, Schmid, R., and Schafer, G. (1993) Arch. Biochem. Biophys. 307, Lawyer, F. C., Staffel, S., Saiki, R. K., Myamba, K., Drummand, R. and Gelfand, D. H. (1989) J. Biol. Chem. 264, 6427-6437 ggajses, G. J., Littlechild, J. A., Watson, H. C., and Hall, L. (1991) Gene 109, Fabry, S., and Hensel, R. (1987) Eur. J. Biochem. 165, 147-155 Sakai, K., Narihara, M., Kasama, Y., Wakayama, M., and Mariguchi, M. (1994) Appl. Environ. Microbial. 60, 2911-2915 Hartag, A. T. and Daniel, R. M. (1992) Int. J. Biochem. 24, 1657-1660 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 76 Bealin-Kelly, R, Kelly, C. T. and Fogarty, W. M. (1991) Appl. Microbial. Biotechnol. 36, 332-336 Brosnan, M. R, Kelly, C. T., and Fagarty, W. M. (1992) Eur. J. Biochem. 203, 225-231 Gray, G. L., Mainzer, S. E., Rey, M. W., Lamsa, M. H., Kindle, K. L., Carmana, C. and Requadt, C. (1986) J. Bacterial. 166, 635-643 Fukusumi, S., Kamizana, A., Harinauchi, S., and Beppu, T. (1988) Eur. J. Biochem. 174, 15-21 Harinauchi, S., Fukusumi, S., Ohshirna, T. and Beppu, T. (1988) Eur. J. Biochem. 176, 243-253 ‘ Hyun, H. H. and Zeikus, J. G. (1985) Appl. Environ. Microbiol. 49, 1162-1167 Melasniemi, H. (1987) Biochem. J. 246, 193-197 Mathupala, S. P., and Zeikus, J. G. (1993) Appl. Microbial. Biotechnol. 39, 487- 493 Plant, A. R., Clemens, R. M., Morgan, H. W. and Daniel, R. M. (1987) Biochem. J. 246, 537-541 Saha, B. C., Lamed, R., Lee, C.-Y., Mathupala, S. P. and Zeikus, J. G. (1990) Appl. Environ. Microbial. 56, 881-886 Madi, E., Antranikian, G., Ohmiya, K., and Gottschalk, G. (1987) Appl. Environ. Microbiol. 53, 1661-1667 Guy, G. R. and Daniel, R. M. (1982) Biochem. J. 203, 787-790 Schafer, G. and Meyering-Vos, M. (1992) Ann. N.Y. Acad. Sci. 671, 293-309 Saha, B. C. and Zeikus, I. G. (1990) Appl. Environ. Microbial. 56, 2941-2943 Wind, R. D., Liebl, W., Buitelaar, R. M., Penninga, D., Spreinat, A., Dijkhuizen, L., and Bahl, H. ( 1995) Appl. Environ. Microbial. 61, 1257-1265 gaggaritis, A., and Merchant, R. F. J. (1986) CRC Crit. Rev. Biotechnol. 4, 327- Schafield, L. R. and Daniel, R. M. (1993) Int. J. Biochem. 25, 609-617 Liithi, E., Jasmat, N. B. and Bergquist, P. L. (1990) Appl. Environ. Microbial. 56, 2677-2683 Lee, Y.-E., Lowe, S.E., Henrissat, B., and Zeikus, LG. 1993. J. Bacterial. 175:5890-5898. Nakao, M., Nakayama, T., Harada, M., Kakuda, A., Ikemato, H., Kobayashi, S. and Shibano, Y. (1994) Appl. Microbial. Biotechnol. 41, 337-343 l‘. H ‘14 (‘fi' 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 77 Saha, B. C., and Zeikus, J. G. (1991) Appl. Microbial. Biotechnol. 35, 568-571 Shaa, W., Obi, S. K. C., Puls, J. and Wiegel, J. (1995) Appl. Environ. Microbial. 61, 1077-1081 Lee, S.-G., Lee, D.-C., Hang, S.-P., Sung, M.-H., and Kim, H.-S. (1995) Appl. Microbial. Biotechnol. 43, 270-276 Gibbs, M. D., Saul, D. J., Liithi, E. and Bergquist, P. L. (1992) Appl. Environ. Microbiol. 58, 3864-3867 Kuriki, T., Okada, S. and Imanaka, T. (1988) J. Bacteriol. 170, 1554-1559 Freeman, 8. -A., Peek, K., Prescott, M., and Daniel, R. (1993) Biochem. J. 295, 463-469 Peek, K., Daniel, R. M., Monk, G, Parker, L., and Coolbear, T. (1992) Eur. J. Biochem. 207, 1035-1044 Peek, K., Veitch, D. P., Prescott, M., Daniel, R. M., MacIver, B., and Bergquist, P. L. (1993) Appl. Environ. Microbial. 59, 1168-1175 Kuriki, T. Park, J. -H., Okada, S. and Imanaka, T. (1988)App1. Environ. Microbial. 54, 2881-2883 Shen, G.-J., Srivastava, K. C., Saha, B. C. and Zeikus, J. G. (1990) Appl. Microbial. Biotechnol. 33, 340-344 Aubert, J. -P., Béguin, P. and Millet, J. (1993)Genetics and Molecular Biology of Anaerobic Bacteria 1993, (Sebald, M., ed.). Springer-Verlag, NY Nashihara, N., Nakamura, N., Nagayama, H. and Horikashi, K. (1988) FEMS Microbiol. Lett. 49, 385-388 Plant, A. R., Morgan, H. W. and Daniel, R. M. (1986) Enz. Microbial. Technol. 8, 668-678 Richter, O.-M. H. and Schiifer, G. (1992) Eur. J. Biochem. 208, 343-349 Richter, O.-M. H. and Schafer, G. (1992.) Eur. J. Biochem. 209, 351-355 3230, W., DeBlais, S., and Wiegel, J. (1995) Appl. Environ. Microbiol. 61, 937- Lee, Y.-E., and Zeikus, J. G. (1993) J. Gen. Microbial. 139, 1235-1243 Liithi, E., and Bergquist, P. L. (1990) FEMS Microbial. Lett. 67, 291-294 Barnes,E. M. Akagi,J. M. and Himes, R. H. (1971) Biochim. Biophys. Acta 227,199- 220 Russell, R. J. M., Haugh, D. W., Dansan, M. J. and Taylor, G. L. (1994) Structure2, 1157- 1167 125. 126. 127. 128. 129. 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 1 44. 78 Kenealy, W. R. and Zeikus, J. G. (1982) FEMS Microbial. Lett. 14, 7-10 Shing, Y. W., Akagi, J. M. and Himes, R. H. (1975) J. Bacteriol. 122, 177-184 Lee, C. and Zeikus, J. G. (1991) Biochem. J. 273, 565-571 Lee, Y.-E., Ramesh, M. V. and Zeikus, J. G. (1993) J. Gen. Microbiol. 139, 1227-1234 Lehmacher, A. and Bisswanger, H. (1990) J. Gen. Microbiol. 136, 679-686 Dekker, K., Sugiura, A., Yamagata, H., Sakaguchi, K. and Udaka, S. (1992) Appl. Microbial. Biotechnol. 36, 727-732 Dekker, K., Yamagata, H., Sakaguchi, K. and Udaka, S. (1991) J. Bacterial. 173, 3078-3083 Chan, M. K., Mukund, S., Kletzin, A., Adams, M. W. W. and Rees, D. C. (1995) Science 267, 1463-1469 Mukund, S. and Adams, M. W. W. (1991) J. Biol. Chem. 266, 14208-14216 Ahem, T. and Klibanov, A. M. (1985) Science 228, 1280-1284 DiRuggiera, J., Robb, F. T., Jagus, R., Klump, H. H., Borges, K. M., Kessel, M., Mai, X. and Adams, M. W. W. (1993) J. Biol. Chem. 268, 17767-17774 Britton, K. L., Baker, P. J., Barges, K. M. M., Engel, P. C., Pasqua, A., Rice, D. W., Robb, F. T., Scandurra, R., Stillman, T. J. and Yip, K. S. P. (1995) Eur. J. Biochem. 229, 688-695 Cansalvi, V., Chiaraluce, R., Paliti, L., Vaccaro, R., De Rosa, M., and Scandurra, R. (1991) Eur. J. Biochem. 202, 1189-1196 Robb, F. T., Park, J .-B. and Adams, M. W. W. (1992) Biochim. Biophys. Acta 1120, 267-272 Cansalvi, V., Chiaraluce, R., Paliti, L., Gambacorta, A., De Rosa, M., and Scandurra, R. (1991) Eur. J. Biochem. 196, 459-467 Wrba, A., Schweiger, A., Schultes, V. and Jaenicke, R. (1990) Biochemistry 29, 7584-7592 Igrggchy, A., Glockshuber, R. and Jaenicke, R. (1993) Eur. J. Biochem. 214, gehultes, V., Deutzrnann, R. and Jaenicke, R. (1990) Eur. J. Biochem. 192, 25- Zwickl, P., Fabry, S., Bogedain, C., Haas, A. and Hensel, R. (1990) J. Bacterial. 172, 4329-4338 Hensel, R., Jakob, 1., Scheer, H. and Lottspeich, F. (1992) Biochem. Soc. Symp. 58, 127-133 145. 146. 147. 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. 160. 161. 162. 163. 79 Hensel, R., Laumann, S., Lang, J ., Heumann, H. and Lattspeich, F. (1987) Eur. J. Biochem. 170, 325-333 Shah, N. N. and Clark, D. S. (1990) Appl. Environ. Microbial. 56, 858-863 Bryant, F. O. and Adams, M. W. W. (1989) J. Biol. Chem. 264, 5070-5079 Pihl, T. D., and Maier, R. J. (1991) J. Bacterial. 173, 1839-1844 Juszczak, A., Aana, S. and Adams, M. W. W. (1991) J. Biol. Chem. 266, 13834-13841 Wrba, A., Jaenicke, R., Huber, R. and Stetter, K. O. (1990) Eur. J. Biochem. 188. 195-201 Ostendorp, R., Liebl, W., Schurig, H. and Jaenicke, R. (1993) Eur. J. Biochem. 216, 709-715 . Hanka, E., Fabry, S., Niermann, T., Palm, P. and Hensel, R. (1990) Eur. J. Biochem. 188, 623-632 Breitung, J., Bdrner, G., Scholz, S., Linder, D., Stetter, K. O. and Thauer, R. K. (1992) Eur. J. Biochem. 210, 971-981 Ma, K., Zirngibl, C., Linder, D., Stetter, K. O. and Thauer, R. K. (1991) Arch. Microbial. 156, 43-48 Ma, K., Linder, D., Stetter, K. O. and Thauer, R. K. (1991) Arch. Microbial. 155, 593-600 Kunaw, J., Schwarer, B., Stetter, K. O. and Thauer, R. K. (1993) Arch. Microbiol. 160, 199-205 Kunaw, J., Linder, D. and Thauer, R. K. ( 1995) Arch. Microbial. 163, 21-28 glamey, J. M. and Adams, M. W. W. (1993) Biochim. Biophys. Acta 1161, 19- 7 Blamey, J. M. and Adams, M. W. W. (1994) Biochemistry 33, 1000-1007 Blake, P. R., Park, J.-B., Bryant, F. 0., Aana, S., Magnusan, J. K, Ecclestan, E., Howard, J. B., Summers, M. F. and Adams, M. W. W. (1991) Biochemistry 30, 10885-10895 Perler, F. B., Comb, D. 6., Jack, W. E., Moran, L. S., Qiang, B., Kucera, R. B., Benner, J., Slatka, B. E., Nwankwa, D. 0., Hempstead, S. K., Carlaw, C. K. S. and Jannasch, H. (1992) Proc. Natl. Acad. Sci. USA89, 5577-5581 Uemori, T., Ishina, Y., Tah, H., Asada, K and Kata, I. (1993) Nucleic Acids Res. 21, 259-265 Hooper, A. B., Hansen, J. and Bell, R. ( 1967) J. Biol Chem. 242, 288-296 164. 165. 166. 167. 168. 169. 170. 171. 172. 173. 174. 175. 176. 177. 178. 179. 180. 181. 182. 80 Klein, A. R., Breitung, J., Linder, D., Stetter, K. O. and Thauer, R. K. (1993) Arch. Microbial. 159, 213-219 Koch, R., Zablawski, P., Spreinat, A. and Antranikian, G. (1990) FEMS Microbial. Lett. 71, 21-26 Laderman, K A., Davis, B. R., Krutzsch, H. C., Lewis, M. S., Grika, Y. V., Privalov, P. L., and Anfmsen, C. B. (1993) J. Biol. Chem. 268, 24394-24401 Laderman, K.A., Asada, K., Uemori, T., Mukai, H., Taguchi, Y., Kata, I. and Anfinsen, CB. (1993) J. Biol. Chem. 268, 24402-24407 Koch, R., Spreinat, A., Lemke, K. and Antranildan, G. (1991) Arch. Microbial. 155, 572-578 Chung, Y. C., Kobayashi, T., Kanai, H., Alciba, T. and Kuda, T. (1995) Appl. Environ. Microbial. 61, 1502-1506 Schumann, J ., Wrba, A., Jaenicke, R., and Stetter, K. O. (1991) FEBS Lett. 282, 122-126 Brown, S.H., and Kelly, RM. (1993) Appl. Environ. Microbial. 59, 2614-2621 Schuliger, J. W., Brown, S. H., Barass, J. A. and Kelly, R. M. (1993) Mol. Marine Biol. Biotechnol. 2, 76-87 Phipps, B. M., Hoffman, A., Stetter, K. O. and Baumeister, W. (1991) EMBO J. 10, 1711-1722 Heyne, R. I. R., de Vrij, W., Crielaard, W. and Kanings, W. N. (1991) J. Bacterial. 173, 791-800 Villa, A., Zecca, L., Fusi, P., Calumba, S., Tedeschi, G. and Tartara, P. (1993) Biochem. J. 295, 827-831 Colombo, S., D'Auria, S., Fusi, P., Zecca, L., Raia, C. A. and Tortora, P. (1992) Eur. J. Biochem. 206, 349-357 Brannenmeier, K., Kern, A., Liebl, W., and Staudenbauer, W. L. (1995) Appl. Environ. Microbial. 61, 1399-1407 Simpson, H. D., Haufler, U. R. and Daniel, R. M. (1991) Biochem. J. 277, 413- 417 Ruttcrsmith, L. D., and Daniel, R. M. (1991) Biochem. J. 277, 887-890 Rugtersmith, L. D., and Daniel, R. M. ( 1993) Biochim. Biophys. Acta 1156, 167- 17 Castantina, H. R., Brown, S. H., and Kelly, R. M. (1990) J. Bacterial. 172, 3654-3660 Bowen, D., Littlechild, J. A., Fathergill, J. B., Watson, H. C. and Hall, L. (1988) Biochem. J. 254, 509-517 183. 184. 185. 186. 187. 188. 189. 190. 191. 192. 193. 194. 195. 196. 197. 198. 199. 200. 201 . 202. 81 Breitung, J., Schmitz, R. A., Stetter, K. O. and Thauer, R. K. (1991) Arch. Microbial. 156, 517-524 Eggen, R., Geerling, A., Watts, J. and deVos, W. M. (1990) FEMS Microbial. Lett. 71,17-20 Marikawa, M., Izawa, Y., Rashid, N., Haaki, T. and Imanaka, T. (1994) Appl. Environ. Microbial. 60, 4559-4566 Ruttersmith, L. D., Daniel, R. M. and Simpson, H. D. (1992) Ann. N.Y. Acad. Sci. 672, 137-141 Slesarev, A. 1., Lake, J. A., Stetter, K. 0., Gellert, M. and Kazyavkin, S. A. (1994) J. Biol. Chem. 269, 3295-3303 Robb, F. T., Masuchi, Y., Park, J .-B. and Adams, M. W. W. (1992) in Biocatalysis at extreme temperatures, (Adams, M. W. W. and Kelly, R.M., eds.), pp. 74-85. American Chemical Society, Washington, DC Tibani, O., Cammarano, P. and Sanangelantani, A. M. (1993) J. Bacterial. 175, 2961-2969 Vieille, C., Hess, J. M., Kelly, R. M. and Zeikus, JG. (1995) Appl. Environ. Microbial. 61, 1867-1875 grtsrvgré,6 S. H., Sj¢halm, C. and Kelly, R. M. (1993) Biotechnol. Bioeng. 41, 7 - Menéndez-Arias, L. and Argos, P. (1989) J. Mal. Biol. 206, 397-406 Watanabe, K., Chishira, K, Kitamura, K. and Suzuki, Y. (1991) J. Biol. Chem. 266, 24287-24294 ngell, C.R., Przybyla, A. and Ljungdahl, L. G. (1990) Biochemistry 29, 5687- 5 4 Bahm, G. and Jaenicke, R. (1994) Int. J. Peptide Protein Res. 43, 97-106 Burdette, D. S., Vieille, C. and Zeikus, J. G. (1996) Biochem. J. 313 (in press) Tomazic, S. J. and Klibanov, A. M. (1988) J. Biol. Chem. 263, 3086-3091 Fantana, A. (1990) in Life under extreme conditions: Biochemical adaptation, (di Prisca, G., ed.), pp. 89-113. Springer-Verlag, Heidelberg Meldgaard, M. and Svendsen, I. (1994) Microbial. 140, 159-166 Olsen, 0. and Thamsen, K K (1991) J. Gen. Microbial. 137, 579-585 Mirsky, A. E. and Pauling, L. (1936) Proc. Natl. Acad. Sci. USA 22, 439-447 Pfeil, W. (1986) in Thermodynamic data for biochemistry and biotechnology, (Hinz, H.-J., ed.), pp. 349-376. Springer-Verlag, Heidelberg 1) A.) 203. 204. 205. 206. 207. 208. 209. 210. 211. 212. 213. 214. 215. 216. 217. 218. 219. 220. 221. 222. 223. 224. 82 Jaenicke, R. (1991) Eur. J. Biochem. 202, 715-728 Ahem, T. J. and Klibanov, A. M. (1988) Meth. Biochem. Anal. 33, 91-127 Steigerwald, V. J., Beckler, G. S., and Reeve, J. N. (1990) J. Bacterial. 172, 4715-4718 ‘ Tchemajenka, V., Vieille, C., Burdette, D. S. and Zeikus, J. G. (1996) manuscript in preparation Londesborough, J. (1980) Eur. J. Biochem. 105, 211-215 Segel, I. H. (1975) Enzyme Kinetics. John Wiley & Sans, NY Smith, C. A., Rangarajan, M. and Hartley, B. S. (1991) Biochem. J. 277, 255- 261 Sakamata, N ., Katre, A. M. and Savageau, M. A. (1975) J. Bacterial. 124, 775- 783 Caultan, J. W. and Kapoor, M. (1973) Can. J. Microbiol. 19, 439-450 Hooper, A. B., Hansen, J. and Bell, R. (1967) J. Biol. Chem. 242, 288-296 LéJahn, H. B., Suzuki, 1. and Wright, J. A. (1968) J. Biol. Chem. 243, 118-128 Hudson, R. C., Ruttersmith, L. D. and Daniel, R. M. (1993) Biochim. Biophys. Acta 1202, 244-250 Glaser, P., Presecan, B., Delepierre, M., Surewicz, W. K., Mantsch, H. H., Bar-Lu, O. and Gilles, A.-M. (1992) Biochemistry 31, 3038-3043 Jencks, W. P. (1975) in Advances in Enzymalagy and related areas of molecular biology vol. 43, (Meister, A., ed), pp. 219-410. John Wiley & Sans, NY Rellas, P. and Scopes, R. K. (1994) Prat. Express. Purif. 5, 270-277 Nagara, T., Makira, Y., Yamamata, K., Urabe, I. and Okada , H. (1989) FEBS lett. 253, 113-116 Kauzmann, W. (1959) Adv. Protein Chem. 14, 1-47 FrleiireésE” Haynie, D. T. and Xie, D. (1993) Proteins: Struct. Funct. Genet. 17, 1 - Haynie, D. T. and Freire, E. (1993) Proteins: Struct. Funct. Genet. 16, 115-140 Haynie, D. T. and Freire, E. (1994) Biapalymers 34, 261-271 Privalov, P. L. and Gill, S. J. (1988) Adv. Protein Chem. 39, 191-234 Dill, K.A. ( 1990) Biochemistry 29, 7133-7155 225. 226. 227. 228. 229. 230. 231. 232. 233. 234. 235. 236. 237. 238. 239. 240. 241 . 242. 243. 244. 83 Dill, K. A., Alonso, D. O. V. and Hutchinson, K. (1989) Biochemistry 28, 5439- 5449 Creighton, T. E. (1990) Biochem. J. 270, 1-16 Creighton, TE. (1993) Proteins: structures and molecular properties 2nd edition. W. H. Freeman and Company, NY Tsubai, M., Ohta, S. and Nakanishi, M. (1978) in Biochemistry of thermaphily (Friedman, M., ed.), pp. 251-266. Academic Press, NY Wiitrich, K., Rader, H. and Wagner, G. (1980) in Protein folding, (Jaenicke, R., ed.), pp. 549-564. Elsevier/North Holland Biomedical Press, Amsterdam Argos, P., Rossmann, M. G., Grau, U. M., Zuber, H., Frank, G. and Tratschin, J. D. (1979) Biochemistry 18, 5698-5703 Matthews, B.W. (1993) Ann. Rev. Biochem. 62, 139-160 Zhang, X.-J., Baase, W. A. and Matthews, B. W. (1992) Protein Sci. 1, 761-776 Heming, T., Yutani, K., Inaka, K., Kuraki, R., Matsushima, M. and Kikuchi, M. ( 1992) Biochemistry 31, 7077-7085 Hurley, J. H., Baase, W. A. and Matthews, B. W. (1992) J. Mal. Biol. 224, 1143-1159 Watanabe, K., Masuda, T., Ohashi, H., Mihara, H. and Suzuki, Y. (1994) Eur. J. Biochem. 226, 277-283 Erikssan, A. B., Baase, W. A., Zhang, X.-J., Heinz, D. W., Blaber, M., Baldwin, E. P. and Matthews, B. W. (1992) Science255, 178-183 Biro, J., Fabry, S., Dietrnaier, W., Bogedain, C. and Hensel, R. (1990) FEBS Lett. 275, 130-134 Kaizumi, J.-I., Zhang, M., Imanaka, T. and Aiba, S. (1990) Appl. Environ. Microbial. 56, 3612-3614 Matsumura, M., Yasumura, S. and Aiba, S. (1986) Nature 323, 356-3358 Shih, P. and J. F. Kirsch (1995) Prat. Sci. 4, 2050-2062 Shih, P. and J. F. Kirsch (1995) Prat. Sci. 4, 2063-2072 Kuraki, R., Inaka, K, Taniyama, Y., Kidakara, S.-I., Matsushima, M., Kikuchi, M. and Yutani, K (1992) Biochemistry 31, 8323-8328 Heming, T., Yutani, K, Taniyama, Y. and Kikuchi, M. ( 1991) Biochemistry 30, 9882-9891 Bell, J. A., Becktel, W. J., Sauer, U., Baase, W. A. and Matthews, B. W. (1992) Biochemistry 31, 3590-3596 ,.p f-_) "J 245. 246. 247. 248. 249. 250. 251. 252. 253. 254. 255. 25 6. 257. 258. 259. 260. 261. 262. 263. 84 Pjura, P. and Matthews, B. W. (1993) Protein Sci. 2, 2226-2232 Matsumura, M., Signor, G. and Matthews, B. W. (1989) Nature 342, 291-293 Erikssan, A. B., Baase, W. A., and Matthews, B. W. (1993) J. Mol. Biol. 229, 747-769 Matthews, B. W., Nicholson, H. and Becktel, W. (1987) Proc. Natl. Acad. Sci. USA 84. 6663-6667 Pjura, P., McIntosh, L. P., Wozniak, J. A. and Matthews, B. W. (1993) Proteins: Struct. Funct. Genet. 15, 401-412 Kuraki, R., Taniyama, Y., Seka, C., Nakamura, H., Kikuchi, M. and Ikehara, M. (1989) Proc. Natl. Acad. Sci. USA 86, 6903-6907 Blaber, M., Lindstram, J. D., Gassner, N., Xu, J ., Heinz, D. W. and Matthews, B. W. (1993) Biochemistry 32, 11363-11373 Anderson, D. E., Hurley, J. H., Nicholson, H., Baase, W. A. and Matthews, B. W. (1993) Protein Sci. 2, 1285-1290 Erikssan, A. E., Baase, W. A., Wozniak, J. A., and Matthews, B. W. (1992) Nature 355, 371-373 Hardy, F., Vriend, G., Veltrnan, O. R., van der Vinne, B., Venema, G. and Eijsink, V. G. H. (1993) FEBS Lett. 317, 89-92 Vriend, G., Berendsen, H. J. C., van der Zee, J. R., van den Burg, B., Venema, G. and Eijsink, V. G. H. (1991) Protein Eng. 4, 941-945 Van den Burg, B., Dijkstra, B. W., Vriend, G., van der Vinne, B., Venema, G. and Eijsink, V. G. H. (1994) Eur. J. Biochem. 220, 981-985 Ishikawa, K., Nakamura, H., Marikawa, K. and Kanaya, S. (1993) Biochemistry 32, 6171-6178 Kimura, S., Nakamura, H., Hashimato, T., Oabatake, M. and Kanaya, S. (1992) J. Biol. Chem. 267, 21535-21542 éfizgmra, S., Kanaya, S. and Nakamura, H. (1992) J. Biol. Chem. 267, 22014- 17 Fabian, H., Schultz, C., Backmann, J., Hahn, U., Saenger, W., Mantsch, H. H. and Naumann, D. (1994) Biochemistry 33, 10725-10730 Mitchinsan, C. and Wells, J. A. (1989) Biochemistry 28, 4807-4815 Wells, J. A. and Powers, D. B. (1986) J. Biol. Chem. 261, 6564-6570 Takagi, H., Takahashi, T., Mamase, H., Inauye, M., Maeda, Y., Matsuzawa, H. and Ohta, T. (1990) J. Biol. Chem. 265, 6874-6878 7'11 “I 718 *1 279 264. 265. 266. 267. 268. 269. 270. 271. 272. 273. 274. 275. 276. 277. 278. 279. 280. 281. 282. 85 Quax, W. J ., Mrabet, N. T., Luiten, R. G. M., Schuurhuizen, P. W., Stanssens, P. and Lasters, I. (1991) Bio/Technology 9, 738-742 Mrabet, N. T., Van den Braeck, A., Van den brande, I., Stanssens, P., Larache, Y., Lambeir, A.-M., Matthijssens, G., Jenkins, J ., Chiadmi, M., van Tilbeurgh, H., Rey, F., Janin, J., Quax, W. J., Lasters, 1., De Maeyer, M. and Wadak, S. J. (1992) Biochemistry 31, 2239-2253 Meng, M., Bagdasarian, M. and Zeikus, J. G. (1993) Bio/Technology 11, 1157- 1161 Valkin, D. B. and Klibanov, A. M. (1987) J. Biol. Chem. 262, 2945-2950 Alber, T., Daa-pin, S., Nye, J. A., Muchmare, D. C. and Matthews, B. W. (1987) Biochemistry 26, 3754-3758 MacArthur, M. W. and Thornton, J. M. (1991) J. Mal. Biol. 218, 3397-3412 Dixon, M. M., Nicholson, H, Shewchuk, L., Baase, W. A. and Matthews, B. W. (1992) J. Mal. Biol. 227, 917-933 Perutz, M. F., Kendrew, J. C. and Watson, H. C. (1965) J. Mol. Biol. 13, 669- 678 El Hawrani, A. S., Moretan, K. M., Sessions, R. B., Clarke, A. R. and Halbroak, J. J. (1994) Trends Biotechnol. 12, 207-211 Creighton, TE (1993) Proteins: structures and molecular properties 2nd ed., W.H. Freeman and Company, NY Hensel, R., Fabry, S., Biro, J., Bogedain, C., Jakob, I. and Siebers, B. (1994) Biocatalysis 11, 151-164 Ishikawa, K., Okumura, M., Katayanagi, K., Kimura, S., Kanaya, S., Nakamura, H. and Marikawa, K. (1993) J. Mal. Biol. 230, 529-542 Davies, G. J., Gamblin, S. J., Littlechild, J. A. and Watson, H. C. (1993) Proteins: Struct. Funct. Genet. 15, 283-289 Machius, M., Weingand, G. and Huber, R. (1995) J. Mol. Biol. 246, 545-559 Baker, E. N., and Hubbard, R. E. (1984) Prag. Biophys. Mol. Biol. 44, 97-179 Cleland, W. W. and Kreevoy, M. M. (1994) Science 264, 1887-1890 Hibberr, F. and Emsley, J. (1990) in Advances in Physical Organic Chemistry, (Bethell, D., ed.), pp. 255-379. Academic Press Ltd., London flagsumura, M., Becktel, W. J. and Matthews, B. W. ( 1988) Nature 334, 406- Yutani, K., Ogasahara, K, Tsujita, T. and Sugina, Y. (1987) Proc. Natl. Acad. Sci. USA 84, 4441-4444 283. 284. 285. 286. 287. 288. 289. 290. 291. 292. 293. 294. 295. 296. 297. 298. 298a. 299. 300. 86 Rashin, A. A., Iafin, M. and Hanig, B. (1986) Biochemistry 25, 3619-3625 McRee, D. E., Redford, S. M., Getzaff, E. D., Lepack, J. R., Hallewell, R. A. and Tainer, J. A. (1990) J. Biol. Chem. 265, 14234-14241 Wilson, K P., Malcolm, B. A. and Matthews, B. W. (1992) J. Biol. Chem. 267 , 10842-10849 Yamada, H., Kanaya, B., Inaka, K., Ueno, Y., Ikehara, M. and Kikuchi, M. (1994) Biol. Pharm. Bull. 17, 192-196 Kanaya, S. and Itaya, M. (1992.) J. Biol. Chem. 267, 10184-10192 Sandberg, W.S. and Terwilliger, T. C. (1989) Science 245, 54-57 Baldwin, E. P., Hajiscyedjavadi, 0., Base, W. A., and Matthews, B. W. (1993) Science 262, 1715-1718 Kwan, K.-S., Kim, J., Shin, H. S. and Yu, M.-H. (1994) J. Biol. Chem. 269, 9627-9631 Heinz, D. W., Baase, W. A. and Matthews, B. W. (1992) Proc. Natl. Acad. Sci. USA 89, 3751-3755 Sicard, P. J., Leleu, J.-B., Duflat, P., Drocaurt, D., Martin, F., Tiraby, G., Petsko, G. and Glasfeld, A. (1990) Ann. N.Y. Acad. Sci. 613, 371-375 Hallewell, R. A., Imlay, K. C., Lee, R, Fang, N. M., Gallegos, C., Getzaff, E. D., Tainer, J. A., Cabelli, D. E., Tekamp-Olsan, P., Mullenbach, G. T. and Couscns, L. S. (1991) Biochem. Biophys. Res. Comm. 181, 474-480 Lepack, J. R., Frey, H. E. and Hallewell, R. A. (1990) J. Biol. Chem. 265, 21612-21618 Zale, S. E. and Klibanov, A. M. (1986) Biochemistry 25, 5432-5444 Valkin, D. B. and Middaugh, C. R. (1992) in Stability of protein pharmaceuticals, Part A: Chemical and physical pathways of protein degradation, (Ahem, TJ. and Manning, M.C., eds.), pp. 215-247. Plenum Press, NY Tomazic, S. J. and Klibanov, A. M. (1988) J. Biol. Chem. 263, 3092-3096 Kuraki, R., Kawakita, S., Nakamura, H. and Yutani, K. (1992) Proc. Natl. Acad. Sci. USA 89, 6803-6807 Gerwig, G. J., Kamerling, J. P., Vliegenthart, J. F. G., Marag, E., Lamed, R. and Bayer, E. A. (1991) Eur. J. Biochem. 196, 115-122 17,25,7215" Lowe, S. E. and Zeikus, J. G. (1993) Appl. Environ. Microbial. 59, Ng, T. K. and Zeikus, J. G. (1981) Biochem. J. 199, 341-35 301. 302. 303. 304. 305. 306. 308. 309. 310. 311. 312. 313. 313a. 313b. 306. 307. 314. 315. 316. 87 Chen, B.-l and King, J. (1991) in Protein Refalding, (Georgian, G. and De Bemardez-Clark, B., eds.), pp. 119-132. American Chemical Society, Washington, DC Bael, E., Brady, L., Brzazawksi, A. M., Derewenda, Z., Dodson, G. G., Jensen, V. J ., Petersen, S. B., Swift, H., Thim, L. and Waldike, H. F. (1990) Biochemistry 29, 6244-6249 Marg, G. A. and Clark, D. S. (1990) Enzyme Microbial. Technol. 12, 367-373 Whitlow, M., Howard, A. J ., Finzel, B. C., Paulas, T. L., Winborne, E. and Gillialand, G. L. (1991) Proteins Struct. Funct. Genet. 9, 153-173 Kasumi, T., Hayashai, K. and Tsumura, N. (1982) Agric. Biol. Chem. 46, 21-30 Stigter, D., Alonso, D. O. V., and Dill, K. A. (1991) Proc. Natl. Acad. Sci. USA 88, 4176-4180 Hei, D. J .and Clark, D. S. (1994) Appl. Environ. Microbial. 60, 932-939 Miller, J. R, Nelson, C. M., Ludlaw, J. M., Shah, N. N. and Clark, D. S. (1989) Biotechnol. Bioeng. 34, 1015-1021 Ltithi, E., Reif, K., Jasmat, N. B. and Bergquist, P. L. (1992) Appl. Microbial. Biotechnol. 36, 503-506 Wilks, H. M., Hart, K. W., Feeney, R., Dunn, C. R., Muirhead, H., Chia, W. N., Barstaw, D. A., Atkinson, T., Clarke, A. R. and Halbroak, J. J. (1988) Science 242, 1541-1544 Wilks, H. M., Halsall, D. J., Atkinson, T., Chia, W. N., Clarke, A. R. and Halbroak, J. J. (1990) Biochemistry 29, 8587-8591 Meng, M., Lee, C., Bagdasarian, M. and Zeikus, J. G. (1991) Proc. Natl. Acad. Sci. USA 88, 4015-4019 Arnold, EH. (1993) FASEB J. 7, 744-749 Chen, K. and Arnold, F. H. (1991) Bio/Technology 9, 1073-1077 De Cardt, S., Hendrickx, M., Maesmans, G. and Tabback, P. (1994) Biotech. Bioeng. 43, 107-114 Oshima, T. (1986) in Thermophiles: General, molecular, and applied microbiology (Brock, T. D., ed.) pp. 137-158. John Wiley & Sans, NY Liza, £10, McKenzie, T. and Hageman, R. (1986) Proc. Natl. Acad. Sci. USA 83, 57 5 Bai, Y., Milne, J. S., Mayne, L. and Englander, S. W. (1994) Proteins: Struct. Funct. Genet. 20, 4-14 Muheim, A., Todd, R. J., Casimira, D. R., Gray, H. B. and Arnold, EH. (1993) J. Am. Chem. Soc. 115, 5312-5313 317. 318. 319. 320. 321. 322. 323. 324. 325. 88 Woese, C. R. (1994) Microbial. Rev. 58, 1-9 Keinan, E., Hafeli, E. K., Seth, K. K. and Lamed, R. L. (1986) J. Am. Chem. Soc. 108, 162-169 Pham, V. T., Phillips, R. S. and Ljungdahl, L. G. (1989) J. Amer. Chem Soc. 111, 1935-1936 Periera, D. A., Gerson, F. P. and Oestreicher, E. G. (1994) J. Biotechnol. 34, 43- 50 Keinan, E., Seth, K. K. and Lamed, R. (1986) J. Am. Chem. Soc. 108, 3474- 3480 Keinan, E., Seth, K. K., Lamed, R., Ghirlanda, R. and Singh, S. P. (1990) Biocatalysis 4, 1-15 Lamed, R. J., Keinan, E. and Zeikus, J. G. (1981) Enz. Microbial. Technol. 3, 144-148 Aaltanen, O. and Rantakyla, M. (1991) Chemtech 1, 240-248 Lee, K. M. and Biellmann, J.-F. (1987) Nuav. J. Chim. 11, 1-4 Chapter II Cloning and Expression of the Gene Encoding the Thermoanaerobacter ethanolicus 39E Secondary-Alcohol Dehydrogenase and Enzyme Biochemical Characterization Accepted for publication in The Biochemical Jamnal 89 90 ABSTRACT The ath gene encoding Thermoanaerobacter ethanolicus 39E secondary-alcohol dehydrogenase (2° ADH) was cloned, sequenced, and expressed in Escherichia coli. The 1056 bp gene encodes a hamatetrameric recombinant enzyme consisting of 37.7 kDa subunits. The purified recombinant enzyme is optimally active above 90°C with a half-life of approximately 1.7 h at 90°C. An NADP(H)-dependent enzyme, the recombinant 2° ADH has 170-fold greater catalytic efficiency in propan-2-ol versus ethanol oxidation. The enzyme was inactivated by chemical modification using dithionitrobenzoate (DTNB) and diethylpyrocarbonate, indicating that Cys and His residues are involved in catalysis. Zinc was the only metal enhancing 2° ADH reactivation after DTNB modification, implicating the involvement of a bound zinc in catalysis. Arrhenius plots for the oxidation of propan-2- 01 by the native and recombinant 2° ADHs were linear from 25°C to 90°C when the enzymes were incubated at 55°C prior to assay. Discontinuities in the Arrhenius plots for propan-2-ol and ethanol oxidations were observed, however, when the enzymes were preincubated at 0°C or 25 °C. The observed Arrhenius discontinuity, therefore, resulwd from a temperature dependent, catalytically significant 2° ADH structural change. Hydrophobic Cluster Analysis comparisons of both mesophilic versus thermophilic 2° ADH and 1° versus 2° ADH amino acid sequences were performed. These comparisons predicted that specific proline residues may contribute to 2° ADH thermostability and thermophilicity, and that the catalytic Zn ligands are different in 1° ADI-Is (two Cys and a His) and 2° ADHs (Cys, His, and Asp). 91 INTRODUCTION Alcohol dehydrogenases (ADHs) (EC 1.1.1.1 [NADH] or EC 1.1.1.2 [NADPHD are well studied as a structurally conserved class of enzyme [1]. The X-ray structure of the horse liver primary-alcohol dehydrogenase (1° ADH) is known, and the properties of this enzyme have been extensively detailed [1,2]. ADHs are typically dimeric or tetrameric pyridine dinucleotide-dependent metalloenzymes with a zinc atom involved in catalysis. ADHs have been classified as 1° or 2° ADHs based on their relative activities toward 1° and 2° alcohols. It generally has been assumed that 1° and 2° ADHs are structm‘ally similar, and that their substrate differences are due to relatively small changes in their active site architecture. Tetrameric 2° ADI-Is have been repormd from a number of microorganisms [3-7]. The Thermoanaerobacter ethanolicus 39E 2° ADH is a bifunctional ADH/acetleoA reductive thioesterase [7]. It has been proposed to function physiologically by oxidizing nicotinamide cofactor during ethanol formation, indirectly preventing glycolytic inhibition at the glyceraldehyde dehydrogenase step [8]. 2° ADHs are attractive subjects for research into chiral chemical production because of their broad specificities and their highly enantiospecific conversion of prachiral ketones to alcohols [9-12]. Two issues facing the commercial scale application of 2° ADHs in chiral syntheses are the difficulty of regenerating and retaining expensive nicotinamide cofactor and the lack of inexpensive, highly stable enzymes. Cofactor regeneration and retention have been overcome by numerous strategies [13,14]. While thermophilic enzymes are generally much more stable than their mesophilic counterparts, the organisms that produce thermophilic 2° ADHs grow slowly and to low cell densities, making an alternative expression system for these enzymes crucial for both their commercial application and detailed protein structure-function studies. The discovery of thermophilic bacteria has provided the opportunity to isolate therrrrostable enzymes directly. The thermophilic anaerobes T. ethanolicus and 92 Thennoanaerobacter brocla'i [15] express extremely enantiospecific 2° ADI-ls that are stable above 70°C. These two 2° ADHs have been proposed to be extremely structrn'ally similar based on similar molecular weight and kinetic characteristics [16]. While the Clostridiwn beijen‘nckii gene (ath) encoding a mesophilic NADP(H)-dependent 2° ADH has been cloned and sequenced (Genbank, Acc. no. M84723), and while the amino acid sequence of the T. brockii 2° ADH has been determined by Edman degradation [17], no cloned thermophilic 2° ADH is available for detailed biochemical studies. The work reported here describes the cloning, sequencing, and expression of the gene encoding the thermophilic T. ethanolicus 2° ADH. The kinetic and thermal properties of the recombinant enzyme are determined. The biochemical basis for enzyme discontinuous Arrhenius plots is investigawd Finally, the protein sequence information is used to examine the structural similarities between thermophilic and mesophilic 1° ADHs and 2° ADHs. These comparisons are used to predict the 2° ADH catalytic Zn liganding residues and the potential involvement of specific prolines in 2° ADH thermal stability. MATERIALS AND METHODS Chemicals and reagents All chemicals were of at least reagent/molecular biology grade. Gases were provided by AGA specialty gases (Cleveland, OH), and made anaerobic by passage through heated copper filings. Oligonucleatide syntheses and amino acid sequence analyses were performed by the Macromolecular Structure Facility (Department of Biochemistry, Michigan State University). The kanamycin resistance GenBlock (EcoRI) DNA cartridge used in expression vector construction was purchased from Pharmacia (U PPsala, Sweden) [1 8] . 93 Media and strains T. ethanolicus 39E (ATCC #33223) was grown in TYE medium as previously described [19]. All batch cultures were grown anaerobically under an N2 headspace. Escherichia coli DH50t containing the recombinant (1th gene was grown in rich complex medium (20 g 1'1 tryptone, 10 g 1'1 yeast extract, 5 g 1'1 NaCl) at 37°C in the presence of 25 pg ml-l kanamycin and 100 ug ml'l ampicillin. DNA manipulations and library construction Plasmid DNA purification, restriction analysis, PCR, and colony and DNA hybridizations were performed using conventional techniques [20,21]. T. ethanolicus chromosomal DNA was purified as previously described [22]. Partially digested 2—5 kb Sau3A I fragments were isolated by size fractionation from a 10-40% sucrose gradient [20] and ligated into pUC18 BamH I/BAP (Pharmacia, Uppsala, Sweden). The ligation mixture was transformed into E. coli DHSa by electroporation [21]. Degenerate primers (1) (5'- ATGAA(R)GG(N)T1‘(H)GC(N) ATG(Y)T) and (2) (5'- G(W)(N)GTCATCAT(R)TC(N)G(K)(D)ATCAT) were used to synthesize the homologous probe for colony hybridization [23]. The DNA fragment containing the ath gene was sequenced using the method of Sanger et al. [24]. Enzyme purification The native 2° ADH was prnified by established techniques [7]. The recombinant enzyme was pmified from E. coli DHSa aerobically. The pelleted cells from batch cultures were resuspended (0.5 g wet wt. ml°1) in 50 mM Tris:HCl [pH 8.0] (buffer A) containing 5 mM DTT, and 10 mM ZnClz, and lysed by passage through a French pressure cell. The clarified lysate was incubated at 65°C for 25 min and then centrifuged for 30 min at 15000g. The supernatant was applied to a DEAE-sephacryl column (2.5 cm x 15 cm) that was equilibrated with buffer A and eluted using a 250 ml NaCl gradient (0-300 mM). 94 Active fractions were diluted 4—fold in buffer A and applied to a Q-sepharose column (2.5 cm x 10 cm) equilibrated with buffer A. Purified enzyme was eluted using a 250 ml NaCl gradient (0-300 mM). Molecular mass determination Recombinant holoenzyme molecular mass was determined by comparison with protein standards (Sigma; St. Louis, MO) using gel filtration chromatography (0.5 ml min'l) with a Pharmacia S300 column (110 cm x 1.2 cm) equilibrated with buffer A containing 200 mM NaCl. Subunit molecular mass values were determined by matrix associated laser desorption ionization mass spectrometry (MALDI) at the Michigan State University Mass Spectrometry Facility. Kinetics and thermal stability The standard 2° ADH activity assay was defined as NADP" reduction coupled to prapan-2- ol oxidation at 60°C as previously described [7]. The enzyme was incubated at 55°C for 15 min prior to activity determination unless otherwise indicated. Assays to determine Kmllpp and Vmaxapp were conducted at 60°C with substrate concentrations between 20meapp and 02me,“. Kinetic parameters were calculated from nonlinear best fits of the data to the Michaelis-Menten equation using Kinzyme software [26]. Protein concentrations were measured using the bicinchoninic acid procedure (Pierce; Rockford, IL) [27]. 2° ADH thermostability was measured as the residual activity after timed incubation at the desired temperatures. Thermal inactivation was stopped by incubation for 30 min at 25°C, and samples were prepared for activity assays by preincubation at 55°C for 15 min. Incubations were performed in 100 [.11 PCR tubes (cat. #72.733.050, Sarstedt; Newton, NC) using 200 pg ml'l protein in 100 111 buffer A. Activity was determined using the unfractionated samples. The temperature effect on enzyme activity was studied using the substrates propan-2-ol or ethanol at 10meapp concentrations. 95 To study the effect of enzyme preincubation temperature on enzyme reaction rates, the 2° ADH was incubated at 0°C, 25°C, or 55 °C for 15 min prior to activity determinations in the temperature range 30—90°C. TriszHCl buffer pH values were adjusted at 25 °C to be pH 8.0 at the temperature they were used (thermal correction factor = -0.031 ApH °C'1). The statistical significance of the differences between the Arrhenius plot slapes above and below the discontinuity temperatures was determined by covariance analysis [25]. Chemical modification Cysteine residues were reversibly modified using dithionitrobenzoate (DTNB) at 25 °C and 60°C [28]. DTNB-inactivated 2° ADH was reactivated using DTT in the presence of 0.01 mM to 1.0 mM metal salts or 0.5 mM to 3.0 mM EDTA. Histidine residues were chemically modified with diethylpyrocarbonate (DEPC) by incubation with 20 mM or 40 mM DEPC in 50 mM phosphate buffer (pH 6.0) at 25°C for 1.0 h and the reaction was quenched by addition of 0.5 x volume of 0.5 M imidazole (pH 6.5) [29]. Protein sequence comparisons The peptide sequences of the Bacillus stearorhermophilus (acc. #D90421), Sulfolobus solfataricus (acc. #851211), and Zymomonas mobilis (acc. #M32100) 1° ADHs and of the Alcaligenes eutrophus (acc. #J03362) and Clostridium beijerinckii (acc. #M84723) 2° ADHs were obtained from GenBank. The horse liver 1° ADH [1] and the T. brocla'i 2° ADH [17] peptide sequences were obtained from the literature. Access to GenBank, standard sequence alignments, and percentages of amino acid similarity/identity were performed using the Program Manual for the Wisconsin Package, Version 8, Sept. 1994 (Genetics Computer Group; Madison, WI). Protein sequence alignments were performed using Hydrophobic cluster analysis (HCA) [30]. HCA plots of individual protein sequences were generated using HCA-Plot V2 computer software (Doriane; Le Chesnay, France). 96 Nucleotide sequence accession number The GenBank accession number for the sequence published in this article is U49975. RESULTS Cloning and sequencing of the T. ethanolicus ath gene The T. ethanolicus 2° ADH was cloned from a T. ethanolicus chromosomal DNA library by homologous hybridization. The N-terminal sequence of the native T. ethanolicus 2° ADH (MKGFAML) was identical to those of C. beijerinckii and T. brockii 2° ADHs. This sequence was used to generate primer (1). Alignment of the C. beijerinckii and T. brockii 2° ADH peptide sequences indicated another conserved region (residues 147-153) that would reverse translate into a low-degeneracy Oligonucleatide (primer [2]). The PCR product obtained using primers (1) and (2), and using T. ethanolicus chromosomal DNA as the template was 470 bp, as expected from the position of primers (1) and (2) in the C. Beijerinclcii and T. brockii 2° ADHs. This PCR product was used as a homologous probe to screen the T. ethanolicus genomic library. The positive clone showing the highest 2° ADH activity at 60°C was selected for frn'ther studies. Plasmid pADHBZS-C contained a 1.6 kb Sau3A I insert and was shown by subsequent sequencing and peptide analyses to carry the complete ath gene. The 1.6 kb insert was subcloned into the pBluescriptIlKS(+) Xba I site to construct the expression plasmid pADHB25. The physical map of pADHB25 is shown in Figure 1. Plasmid pADI-IB25 was stabilized by insertion of a kanamycin resistance cartridge into the vector EcoR I site, allowing dual selection on kanamycin and ampicillin. This final construct (pADHB25-kan) was used for all subsequent work. 97 The nucleotide sequence of the pADI-IB25 insert is shown in Figure 2. A unique open reading frame (ORF) was identified that encoded a polypeptide which was highly homologous to C. beijerinckii 2° ADH, and which started with the N-terminal sequence of the native T. ethanolicus 2° ADH. Two consecutive ATG codons were identified as potential translation initiation codons. The N-terminus of the native 2° ADH starts with a single Met, suggesting that the first ATG codon is not translated or is post-translationally removed. A potential ribosome binding site (RBS) (positions 223-228) is located 10 bp upstream of the start codon. A potential promoter was identified approximately 70—100 bp upstream of the RBS. The "-35" and "-10" regions are highly similar to the E. coli consensus promoter sequences [31], and are separated by 16 bp. The "-10" region is duplicated with the second copy overlapping the first. Because of its improper distance from the "-35" region, this second copy may only provide an A+T rich sequence to aid in strand separation. N a transcriptional stop site was identified downstream of ath. In this region, instead, a truncated ORF, preceded by an RBS, was identified. It translated into a 45 amino acid peptide fragment 46% identical and 68% similar to the product of a similar truncated ORF located downstream of C. beijerinckii ath gene. Little similarity was found with other sequences in GenBank, so the function of this ORF remains unknown. Sequence comparison of 1° and 2° ADHs Standard alignments of 1° and 2° ADH amino acid sequences indicated a high level of similarity among the three 2° ADI-Is from obligate anaerobes (Table 1). The T. ethanolicus 2° ADH differed from the T. brocla'i enzyme by only three residues, and was 75% identical to the enzyme from the mesophile C. beijerinckii. The similarity was lower for comparisons with the 2° ADH from the obligate aerobe A. cutrophus. The 1° ADHs showed less sequence conservation (25 to 54% identity and 49 to 71% similarity) than the 2° ADHs, and showed only 20 to 27% identity (and 48 to 51% similarity) to the 2° ADHs. Based on these standard alignments, the conservation of important core domain residues, 98 Figure 1. Restriction map of the T. ethanolicus 39E 2° ADH clone (pADHB25). The flanking restriction sites are from the plasmid polylinker. 0 kb 0.5 kb 1.0 kb 1.5 kb EcoR I} ‘Hind 111 Pst I EcoR V KpnI / ‘ Pstl Ss I \ Del 14ch P \Ach . Xbal EcoRV 100 Figure 2. Nucleotide sequence and deduced amino acid sequence of T. ethanolicus 39E ath and of the downstream open reading frame. The putative promoter -35 and -10 regions and RBSs are underlined. The ath stop cation is indicated by three asterisks. of T. WM 101 I I 90 TGAACAATAGACAACCCCTTTCTGTGATCTTGTTTTTTGCAAATGCTATTTTATCACAAGAGATTTCTCTAGTTCTTTTTTACTTAAAAA I I I I I I I I 180 AACCCTACGAAATTTTAAACTATGTCGAATAAATTATIGAIAATTTTTAACTATGTGCIAIIATATTATTGCAAAAAATTTAACAATCAT _ -1 I I I 3? I I o I I 270 CGCGTAAGCTAGTTTTCACATTAATGACTTACCCAGTATTTTAGGAGGTGTTTAATGATGAAAGGTTTTGCAATGCTCAGTATCGGTAAA RBS M K G F A M L S I G K I I I I I I I I 360 GTTGGCTGGATTGAGAAGGAAAAGCCTGCTCCTGGCCCATTTGATGCTATTGTAAGACCTCTAGCTGTGGCCCCTTGCACTTCGGACATT V G W I B K E K P A P G P F D A I V R P L A V A P C T S D I 450 CATACCGTTTTTGAAGGAGCCATTGGCGAAAGAATAACATGATACTCGGTCACCGAAGCTGTAGGTGAAGTAGTTGAAGTAGGTAGTGAG H T V P E G A I G E R H N M I L G H E A V G E V V E V G S E I I I I I I I I 540 GTAAAAGATTTTAAACCTGGTGATCGCGTTGTTGTGCCAGCTATTACCCCTGATTGGTGGACCTCTGAAGTACAAAGAGGATATCACCAG V K D P K P G D R V V V P A I T P D N W T S E V O R G Y R Q I I I I I I I I 630 CACTCCGGTGGAATGCTGGCAGGCTGGAAATTTTCGAATGTAAAAGATGGTGTTTTTGGTGAATTTTTTCATGTGAATGATGCTGATATG H S G G H L A G N K F S N V K D G V F G E F F H V N D A D M I I I I I I I . I 720 AATTTAGCACATCTGCCTAAAGAAATTCCATTGGAAGCTGCAGTTATGATTCCCGATATGATGACCACTGGTTTTCACGGAGCTGAACTG N L A H L P K E I P L E A A V M I P D M M T T G P H G A B L. I I I I I I I I 810 GCAGATATAGAATTAGGTGCGACGGTAGCAGTTTTGGGTATTGGCCCAGTAGGTCTTATGGCAGTCGCTGGTGCCAAATTGCGTGGAGCC A D I E L G A T V A V L G I G P V G L M A V A G A K L R G A I I I I I I I I 900 GGAAGAATTATTGCOGTAGGCAGTAGACCAGTTTGTGTAGATGCTGCAAAATACTATGGAGCTACTGATATTGTAAACTATAAAGATGGT G R I I A V G S R P V C V D _A A K Y Y G A T D I V N Y K D G I I I I I I I I 990 CCTATCGAAAGTCAGATTATGAATCTAACTGAAGGCAAAGGTGTCGATGCTGCCATCATCGCTGGAGGAAATGATGACATTATGGCTACA P I E S Q I H N L T E G K G V D A A I I A G G N A D I M A T I I I I I I I I 1080 GCAGTTAAGATTGTTAAACCTGGTGGCACCATCGCTAATGTAAATTATTTTGGCGAAGGAGAGGTTTTGCCTGTTCCTCGTCTTGAATGG A V K I V R P G G T I A N V N Y F G E G B V L P V P R L E I I I I I I I I I 1170 TGGCTCATAAAACTATAAAAGGCGGGCTATGTTCCGGTGGACGTCTAAGAATGGAAAGACTGATTGACCTTGTTTTTTAT G C G H A R K T I K G G L C P G G R L R M B R L I D L V F Y I I I I I I I I 1260 AAGCCTGTCGATCCTTCTAAGCTCGTCACTCACGTTTTCCAGGGATTTGACAATATTGAAAAAGCCTTTATGTTGATGAAAGACAAACCA K P V D P S K L V T H V P Q G F D N I B R A P M L M R D K P I I I I I I I I 1350 AAAGACCTTATCAAACCTGTTGTAATATTAGCAIAAAAATGGGGACTTAGTCCATTTTTATGCTAATAAGGCTAAATACACTGGTTTTTT K D L I K P V V I L A *** I I I I I I I I 14‘0 TATATGACACATCGGCCAGTAAACTCTTGGTAAAAAAATAACAAAAAATAGTTATTTTCTTAACATTTTTACGCCATTAACACTTGATAA I I I I I I I I 1530 CATCATOGAAGAAGTAAATAAACAACTATTAAATAAAAGAAGAAGGAGGATTATCATGTTCAAAATTTTAGAAAAAAGAGAATTGGCACC M F K I L E K R E L A P t r r r t I r t 1620 TTCCATCAAGTTGTTTGTAATAGAGGCACCACTAGTAGCCAAAAAAGCAAGGCCAGGCCAATTCGTTATGCTAAGGATAAAAGAAGGAGG S I K L P V I B A P L V A K K A R P G N F V M L R I K B G G I AGAAAGAATT 1630 E R I —fl#" “um" I I 102 as. emeoofiefi u... .55 Seen unz— e5 3235 .N £9: .N goreaboe. .m fie. .m ”QEQQEbSESa .m Home. .m Sacha» .< S.» .< "agreemed "N3 .0 "383.563 H "wan esoaeeefie H ”5» H G .8 commeofiomea $5235 9% e8 novice; .23 888E new UUG 05 .5 moan—3:8 203 «outage v5 83:82 agate 8e 35 2m son 33 see ”.2 came in one in $.49 3a 8.3 new oi 2: one on... 3:33 3.5 93 ewe new 933 “.3 83:3 oosN 2: 8.3 one 8.3 «.3 see in 8.5 eon 6.5 SN 3... .e... 8“ one in Goo ea 83 SN Goo SN 8:. .m... ESQ. 8e Goo «.8 one «on 58 eon .3 .< 8“ see one see 4.: E .o 8“ $59 go one 6. 2: so .5... g be: 82 .N 3. .m one .m 3» .< E .o . one .5 so .5 fig. "me—€338 05 Set 805.58 55 85.68 we Swain—mm .5 3:83 am 8353633 3:83 3:803 93 3.»: 88.. e833 bean—:8 3583» SEE we canton—=00 A o—efl. 103 Figure 3. HCA comparison of thermophilic and mesophilic 1° and 2° ADHs centered on the putative nicotinamide binding motifs. The proposed catalytic Zn liganding (O), structural Zn liganding Cl), and Rossmann fold binding motif 0) residues are indicated. Symbols:*: proline”; glycine;l: serine; n: threonine. 104 UJD" I 8W5. A o N KAf t I l D K N Ev A 0'85 O N. D K LK I. ethanolicus 39E 2° Adh Q. beijerinckii 2° Adh _B. stearothermophilus 1° Adh horse liver 1° Adh 105 amino acids lining the active site pocket, and structurally conserved glycines identified for the horse liver enzyme [1] was higher in all the peptides. HCA comparisons allow the alignment of potentially similar 2° or 3° structural regions between enzymes with dissimilar amino acid sequences. The Rossmann fold consensus sequences {Gly-Xaa-Gly-Xaa-Xaa-Gly-(Xaa)13.20[negatively charged amino acid for NAD(H) dependent or neutral amino acid for NAD(P)(H)]} [32] identified in each of the ADH peptides were used to direct HCA alignments of 1° and 2° ADHs. Figure 3 shows a representative HCA multiple comparison between a thermophilic (T. ethanolicus ) and a mesophilic (C. beijerinckit) 2° ADH with a thermophilic (B. stearothermophilus) and a mesophilic (horse liver) 1° ADH based on the identified Rossmann fold consensus sequences. The consensus motifs of the N ADP(H) linked T. ethanolicus and C. beiierinckii 2° ADHs were identified as Gly174 -Xaa-Gly175-Xaa-Xaa-Gly179, Gly193, and those of the NAD(H) dependent B. stearothermophilus and horse liver 1° ADH were identified as Gly172 - Xaa - Gly174-Xaa-Xaa-Gly177, Asp195 and Gly199-Xaa-Gly201-Xaa- Xaa-Gly“, Asp223, respectively. This alignment also predicts 71 to 94% overall similarity between 1° and 2° ADI-Is for the critical hydrophobic core residues (71 to 94%), active site pocket residues (71 to 100%), and structurally critical glycines (100% in all cases) identified in the horse liver enzyme. The HCA comparison allowed the identification of corresponding cluster regions and catalytically important residues in all fatn' dehydrogenases based on the corresponding residues identified for the horse liver enzyme [1,2]. The horse liver 1° ADH contains two Zn atoms per subunit, one structural and one catalytic [1]. The thermophilic B. Stearothermophilus 1° ADH contained a region involving cysteine residues 97, 100, 103, and 111 that is analogous to the structural Zn binding loop in the horse liver ADH (involving Cys92, Cys95, Cys93, and Cy8105). However, no analogous structural Zn binding loop regions were identified in the T. ethanolicus and C. beijerinckii 2° ADHs (Fig. 3). The catalytic Zn ligands in the horse liver enzyme have been established (Cys45, His53, 106 and Cysl74). The N-terminal cysteine and histidine ligands to the catalytic Zn atom appear to be conserved in both the 1° and 2° ADH sequences (Fig. 3). However, the second cysteine ligand in the horse liver 1° ADH (Cys174), conserved in the B. stearothermophilus 1 ° ADH(Cys143), is substituted with an aspartate residue (Asp150) in both the T. ethanolicus and C. beijerinckii proteins. The A. eutrophus and T. brockii 2° ADHs also conserved corresponding aspartate residues and appeared to lack 1° ADH-like structural Zn binding loops (data not shown). Sequence comparison of 2° ADH enzymes from a mesophile and th ermophiles The structln'al constraints introduced by proline residues have been proposed as a Incchanism involved in protein thermostabilization [33]. Among the twelve nonconservative sequence substitutions between the mesophilic C. beijerinckii 2° ADH and the thermophilic T. ethanolicus 2° ADH, nine correspond to the introduction of prolines (22, 24, 149, 177, 222, 275, 313, 316, and 347) in the T. ethanolicus protein. All but P103 13 are also present in the thermophilic T. brockii 2° ADH. The mesophilic 2° ADH subunit contained a total of 13 prolines (3.7%) while the T. ethanolicus and T. brockii subunits contained 22 prolines (6.2%) and 21 prolines (6.0%), respectively. Prolines 20, 22. and 24 are in a 9 residue stretch of hydrophilic amino acids near the putative catalytic 2“ 1igand Cys37. Pro”9 interrupts a hydrophobic cluster and is next to another putative Zn ligand, Asp150. The nicotinamide cofactor binding motif includes Pro177 which also interrupts a hydrophobic cluster region. Prolines 222, 275, 313, 316, and 347 are located In Shon (two to four residue) hydrophilic stretches that form putative turn regions. PI“‘if‘ication and characterization of the recombinant 2° ADH The recombinant r. ethanolicus 2° ADH was highly expressed in E. coli in the absence of 1ndfiction and no significant increase was seen upon induction with 5.0 mM isapropylthio- 107 b-D—galactoside. The enzyme expression level was similar in E. coli and in the native organism (1% to 5% of total protein). The recombinant enzyme was purified 36-fold to homogeneity (as determined by the presence of a single band on SDS-PAGE). The subunit molecular masses were calculated from the average of 3 (native enzyme). and 5 (recombinant enzyme) determinations using MALDI, yielding masses for the native and recombinant enzyme subunits of 37707 Da and 37854 Da, respectively. These values are within the generally accepted error of the technique (~1%), and are in agreement with the theoretical molecular mass for the native enzyme based on the gene sequence (37 644 Da). N —terminal amino acid analysis of the recombinant enzyme indicated that in 72 i 5% of the recombinant protein the two N-terminal ATG codons had been translated, while 28 i 2% of the protein contained a single N-terminal Met residue, like the native enzyme. The increased mass of 147 Da determined for the recombinant enzyme is consistent with the mass of one Met residue (149.2 Da), although this difference is small compared to the measurement error associated with MALDI. The recombinant holoprotein molecular mass was determined to be 160 kDa by gel filtration chromatography demonstrating that the recombinant T. ethanolicus 2° ADH, as the native enzyme, is a homotetramer. The 2° ADH was completely inactivawd at both 25°C and 60°C by cysteine-specific DTNB modification. Inactivation was partially reversed at 60°C by the addition of DTT allowing the recovery of 34% initial activity (initial activity = 54 1 3.0 units mgl). The addition of CdSO4 , FeClz, MnClz, CaClz, MgClz, NaCl (0.01 mM to 1.0 mM), or EDTA (0:5 mM to 3.0 mM) did not affect enzyme reactivation by DTT, and the addition of CoC12 01’ Nicr; reduced the recovered activity to only 10% or 15% of the initial activity, respectively. Zinc was the only metal to enhance the reactivation (up to 48% activity rt"unwary in the presence of 100 M ZnClz). DEPC modification of histidine residues also completely inactivated the enzyme. 108 Characterization of thermal and kinetic properties of the recombinant T. ethanolicus 2° ADH The activity of the recombinant 2° ADH toward ADH substrates was characterized. The Vmaxapp for propan-2-ol (68 units mg‘l) was 3.6-fold higher than for ethanol (19 units mg‘l), and the Kmapp toward the 1° alcohol (53 mM) was almost 50-fold higher than for the 2° alcohol (1.1 mM). The catalytic efficiency of the recombinant enzyme was determined to be approximately 170-fold greater toward the 2° alcohol (0.062 ml min'1 mg“ 1) than toward the 1° alcohol (0.00036 ml min'1 mg'l). The Kmapp for NADP+ was 0.011 mM, and no NAD(H) dependent activity was detected. The temperature dependence of native and recombinant enzyme activities were determined to be similar. Thus, only the data obtained with the recombinant enzyme are reported here. T. ethanolicus 2° ADH activity was detected below 25 °C and increased to beyond 90°C (Fig. 4A). The 2° ADH half-life at 90°C was 1.7 h. The Arrhenius plot for the oxidation of propan-2ool was linear from 30°C to 90°C when the enzyme was incubated at 55°C prior to assay. Under the same conditions, however, a distinct discontinuity was seen in the Arrhenius plot for ethanol oxidation (Fig. 4B). Discontinuities were also observed at ~55°C and ~46°C for propan-2- 01 and ethanol oxidations, respectively, when the enzyme was preincubated at 0°C or 25°C (Table 2). The slopes of the best fit regression lines above and below the discontinuity were significantly different beyond the 95% confidence level except for propan-2-ol oxidation by enzyme preincubated at 55°C, where the regression line slopes were similar at the 95% confidence level. The activation energies for propan-2-ol and ethanol oxidations were similar at assay temperatures above the discontinuities but were 15 to 20 kJ mol'1 higher for ethanol oxidation than for propan-Z—ol oxidation at temperatures below the discontinuity temperatures (Table 2). Furthermore, the differences between activation energies above and below the discontinuity temperatures (D) decreased with increasing preincubation temperatures. and the differences for ethanol oxidation were at least 3-fold higher than for propan-2-ol oxidation. 109 Figure 4. Arrhenius plots for the recombinant T. ethanolicus 39E 2° ADH between 25°C and 90°C. (A) Temperature-activity data for propan-2-ol oxidation with the enzyme preincubated at 55°C. Linear regression best fits to the data was determined to be y = 13.647 - 3.3694x (R2 = 0.993). (B) Temperature-activity data for ethanol oxidation with the enzyme preincubated at 55°C. Points above and below the discontinuity are indicated by. and., respectively. Linear regression best fits to the data were determined to be y = 9.7929 - 2.5036x (R2 = 0.946) for points above and y = 15.730 - 4.3899x (R2 = 0.969) for points below the discontinuity. 110 [(Bw/suun) Amuse ouioadslm n 0. v: C. v: 0. MMNN—iv—t 3.3 3.1 2.9 2.7 A 3.3 3.1 2.9 Lir Ifijfififii' I - a Q '0 9. V? q W. Q in V V M M N N [win/SIN“) KJIAIIOB 01.1109dSJUI 3 10 x 1 [Temperature (K) 111 .293 858:8 e3 2.. a anemones 382.com a; 53:25:. as 23 osaosoo .oz Haze Arm—55586 05 26% 3.88 533.03 - 33658:. 05 323 3.85 ecu—268v u an .83 05 S 8:: comic." E “no; 05 .«o abuse 85580 .3 ~32 853:8 $3 05 a vogue 83 “ca 3205 2335 a .«e consumes ofiu S on _N no nn cm 3. mm 3 mm R we 2 me o 3:an fl S on 96 mm o em on em mm o 8 S R c l «receded 325.88% 33558:. Q. 323 26% Gee Gov Between 88% 5% 9.38 we 3.85 eeuguo< 335.585— :cufleoemoun l 43. em mom gouge g .3 Scene .0550 e5 ~e§q€9~ e8 3.85 52:68 05 no gages 5393.35 go we Scum .N 3%.”. 112 DISCUSSION This first report of the cloning and expression of a thermophilic 2° ADH also provides evidence on the molecular bases for enzyme activity, thermophilicity, and discontinuous Arrhenius plots. Protein sequence comparisons predict that 1° and 2° ADHs have some structure-function similarities related to catalytic Zn and nicotinamide cofactor binding. However, differences in the hydrophobic clusters present in the overall 1° and 2° ADH sequences predict significant differences in the overall structures of these enzymes. The recombinant T. ethanolicus protein, with kinetic properties similar to those of the native enzyme [6], is a thermophilic and thermostable NADP(H) dependent enzyme that exhibits significantly greater catalytic efficiency toward 2° alcohols than 1° alcohols due to both lower Kmapp and higher Vmaxapp values. Chemical modification experiments suggest that enzyme activity requires at least Cys and His residues and a tightly bound Zn atom. Finally, the magnitude of the bend in the Arrhenius plots varied inversely with enzyme preincubation temperature, indicating the existence of a catalytically significant, temperature dependent structural change in the enzyme. The ath gene encoding the thermophilic T. ethanolicus 2° ADH was cloned by hybridiution and expressed in E. coli. Comparisons of the native and recombinant enzyme N-terminal sequences and molecular masses showed that the recombinant enzyme differed from the native protein only by the addition of an N-terminal methionine. While the first ATG is the main translation start codon in E. coli, it is unknown if it is the sole initiation codon (and the first methionine is processed later), or if the second ATG is also used as an initiation codon. The presence of a truncated ORF (preceded by an RES) 180 nucleotides downstream of ath and the absence of any potential transcriptional stop signal suggested that ath may be the first gene of an operon. Although confirmation of this hypothesis will require additional experimentation, the similarity between, the truncated ORF reported 1 13 here and the one downstream of C. beijerinckii ath lends further support to this hypothesis. HCA alignments of the 1° and 2° ADHs directed by their pyridine dinucleotide binding motifs indicated some structural similarity between these classes of enzymes. The negatively charged Asp(223r 195» 215) residues are consistent with the NAD(H) dependence of the horse liver, B. stearothennophilus , and A. eutrophus enzymes, respectively, while the presence of an uncharged residue (Gly193) at the analogous position in the T. ethanolicus, T. brockii and C. beijerinckii 2° ADHs is consistent with the NADP(H) dependence of these enzymes [3 ,6,7]. Furthermore, residues identified in the enzyme hydrophobic core, residues lining the active site cavity, and structurally important glycines previously identified in the horse liver 1° ADH [l], were better conserved between proteins in this alignment than the overall peptide sequences. Similarity of putative structurally critical regions has been reported for other ADH comparisons [34]. The correlation between enzyme structure-function properties and the predictions of this alignment support its use in further ADH structure-function studies. However, the apparent lack of a structural Zn binding loop in the 2° ADHs, the reports of only tetrameric 2° ADHs but both tetrameric and dimeric 1° ADHs, and the significantly greater similarity between the 2° ADHs than between the 1° and 2° ADHs predicted by HCA comparison also suggests that, while functionally similar, 1° and 2° ADHs may be less str'uctm'ally similar than was it previously believed. The thermophilic T. ethanolicus 39E 2° ADH incubated at 90°C retained detectable activity for more than one how while the mesophilic C. beijerinckii 2° ADH was completely inactivated within 10 min at 70°C [6]. Still, these two enzymes shared more than 85% sequence similarity. Nine of the nonconservative substitutions between the subunits of these proteins correspond to prolines in the T. ethanolicus enzyme. The similarly thermophilic T. brockii 2° ADH contains 8 of these additional prolines. This difference in proline content of the thermophilic versus mesophilic 2° ADHs is consistent 1 14 with the hypothesis of Matthews er al. regarding the role of prolines in protein stabilization [33] and with the observations by numerous investigators of protein stabilization due to proline insertion [35 ,36], constrained loop regions [37], and proline substitution into loops [38—40]. The additional prolines in the T. ethanolicus 2° ADH were either in short putative loop regions or in longer putative loop regions containing multiple prolines, suggesting that the specific placement as well as the number of prolines may be critical to their stabilizing effect. The difference in proline content between the T. ethanolicus and C. beijerinckii enzymes and their overall high sequence similarity makes these 2° ADHs an excellent system for testing the effect of proline insertion on protein structtne stabilization. Comparative sequence analysis predicted the involvement of specific Cys, His, and Asp residues in 2° ADH catalysis. Chemical modification of the recombinant and native T. ethanoll’cus 2° ADHs with DTNB established the importance of cysteine residues in catalysis. The observation that Zn was the only metal enhancing the enzyme reactivation after DTNB modification suggested that the enzyme is Zn dependent. This finding is in agreement with the previous report that, once inactivated by the sulfhydryl modifying reagent p-chloromercuribenzoate, the T. brockl'i 2° ADH recovered activity only in the presence of ZnClz [3]. Here, the recovery of 34% initial activity upon reversal of DTNB modification in the absence of added metal and the inability of EDTA to reduce the rate of reactivation, even at 60°C, suggested that the catalytic metal remained tightly bound to the protein. Histidine specific modification of the T. ethanolicus enzyme by DEPC was also accompanied by complete enzyme inactivation, implicating a histidine residue in catalysis. The apparent lack of a structural Zn binding region in the 2° ADH subunits argues that the DTNB linked inactivation and Zn dependent reactivation are not due to the loss and recovery of a cysteine-liganded structural Zn. Therefore, the catalytically important cysteine and histidine residues may act as ligands to the catalytic (Zn as described for other ADHs [1]. Determination of the importance of Asp150 to 2° ADH activity awaits mutagenic, crystallographic, or physical biochemical analysis. 1 15 The temperature dependence of catalytic activities for the native and recombinant enzymes were similar. The Arrhenius plot for propan-2-ol oxidation by T. ethanolicus 2° ADH preincubated at 55°C was linear, unlike that previously reported for the T. brockii 2° ADH [3]. However, statistically significant discontinuities in the Arrhenius plots for both ethanol and propan-2-ol oxidation by the T. ethanolicus enzyme were seen when the enzyme was preincubated at lower temperatures. A change in the rate determining step of the overall reaction and other explanations that do not invoke alterations in catalyst structure predict Arrhenius plot discontinuities [41,42]. A shift in the reaction slow step not related to an alteration in enzyme structure would be independent of the initial temperature of the enzyme, and is inconsistent with the inverse relationship observed here between enzyme preincubation temperature and the difference in activation energies above and below the discontinuity (D). At assay temperatures above the discontinuity, the reaction activation energies were similar for enzymes preincubated at 0°C, 25°C and 55°C, unlike those at lower assay temperatures. This observation is consistent with the enzyme attaining its optimally active conformation rapidly enough at higher assay temperatures not to affect the measured rate. The discontinuity temperatures, being below the lowest reported temperatures for T. ethanolicus growth, are physiologically irrelevant, but they underscore the importance of treating thermophilic enzymes differently from mesophilic enzymes when conducting kinetic analyses. Furthermore, the differences between the low temperatme activation energies for ethanol and propan-2-ol argue that substrate/product-protein interactions are important to the rate determining step in catalysis. The linear Arrhenius plot for propan-2-ol oxidation by T. ethanolicus 2° ADH preincubated at 55°C indicates that the low mesophilic temperature activity of this enzyme may be explained by the substrate energy alone. Arrhenius theory predicts increasing Kmapp and Vmaxapp with increasing temperature [41], and this effect has been confirmed for the Thermotoga neopolitana D-xylose isomerase [43]. The thermophilic 2° ADH has lower Vmaxapp values than would be predicted from extension of the mesophilic enzyme 1 16 activity at 25°C. In fact, at 60°C the thermophilic 2° ADH maintains Kmapp and Vmaxapp values toward substrates similar to those reported for the mesophilic enzyme at 25°C. These similar kinetic values suggest that the thermophilic enzyme channels more total reaction energy into substrate binding and less into turnover [44]. Therefore, an alternative explanation for the comparatively poor T. ethanolicus 2° ADH low temperature activity which argues that substrate affinity at high temperatures is maintained by sacrificing high turnover number must also be considered. ACKNOWLEDGMENTS I gratefully acknowledge Dr. Claire Vieille for her aid in performing the molecular biology and in writing this manuscript. 1 also gratefully acknowledge Maris Laivenieks for his assistance in protein purification and molecular biology, as well as Drs. Vladimir Tchemajenko and David Giegel for their helpful discussions. The MSU Mass Spectrometry Facility is supported, in part, by a grant (DRR-00480) from the Biotechnology Resource Technology Program, National Center for Research Resources, National Institutes of Health. This research was supported by a grant from the Cooperative State Research Service, US. Department of Agriculture, under the agreement 90-34189- 5014. 10. ll. 12. 13. 14. 15. l6. 17. 18. 19. 117 REFERENCES Branden, C.-I., Jornvall, H., Eklund, H. and Furugren, B. (1975) in The enzymes, vol. XI part A (Boyer, P. D., ed.), pp. 103-190, Academic Press, NY Eklund, H., Nordstrom, B., Zeppezauer, B., Soderlund, G., Ohlsson, I., Boiwe, T. and Branden, C.-I. (1974) FEBS Lett. 44, 200-204 Lamed, R. J. and Zeikus, J. G. (1981) Biochem. J. 195, 183-190 Steinbl'ichel, A. and Schlegel, H. G. (1984) Eur. J. Biochem. 141, 555-564 Bryant, F. 0., Wiegel, J. and Ljungdahl, L. (1988) Appl. Env. Microbiol. 54, 460- 465 Ismaiel, A. A., Zhu, C.-X., Colby,G. D. and Chen, J.-S. (1993) J. Bacteriol. 175, 5097-5105 Burdette, D. S. and Zeikus, J. G. (1994) Biochem. J. 302, 163-170 Lovitt, R. W., Shen, G.-S. and Zeikus, J. G. (1988) J. Bacteriol. 170, 2809-2815 Keinan, B., Hafeli, E. K., Seth, K. K. and Lamed, R. L. (1986) J. Am. Chem. Soc. 108, 162-169 Keinan, B., Hafeli, E. K., Seth, K. K. and Lamed, R. L. (1986) Ann. N. Y. Acad. ’ Sci. 501, 130-149 Hummel, W. (1990) Appl. Microbiol. Biotech. 34, 15-19 Keinan, B., Seth, K. K., Lamed, R. L., Ghirlando, R. and Singh, S. P. (1990) Biocatalysis 4, 1-15 Hummel, W. and Kula, M.-R. (1989) Eur. J. Biochem. 184, 1-13 Persson, M., Mansson, M.-O., Biilow, L. and Mosbach, K. (1991) Biotechnology 9, 280-284 Lee, Y.-E., Jain, M. K., Lee, C., Lowe, S. B., and Zeikus, J. G. (1993) Int. J. Syst. Bacteriol. 43, 41-51 Nagata, N., Maeda, K. and Scopes, R. K. (1992) Bioseparation 2. 353-362 Peretz, M. and Burstein, Y. (1989) Biochemistry 28, 6549-6555 Oka, A., Sugisaki, H. and Takanami., M. (1981) J. Mol. Biol. 147, 217-226 4Zfil4ctsls, J. G., Hegge, P. W. and Anderson, M. A. (1979) Arch. Microbiol. 122, 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. l 1.8 Sambrook, J., Fritsch, E. F. and Maniatis, T. (1989) Molecular cloning: A laboratory manual 2nd edition. (Nolan, C., ed.), Cold spring Harbor Press, NY Ausubel, F. M., Brent, R., Kingston, R. B., Moore, D. D., Seidman, J. G., Smith, J. A. and Struhl, K. (1993) Current Protocols in Molecular Biology (Janssen, K., ed.), Current Protocols, NY Doi, R. H. ( 1983) in Recombinant DNA Techniques: An Introduction (Rodriguez, R. L, and Tait, R. C., eds.), pp. 162-163, Benjamin-Cummings, Menlo Park, CA Lee, C. C., Wu, X., Gibbs, R. A., Cook, R. G., Muzny, D. M. and Caskey, C. T. (1988) Science 239, 1288 - 1291 Sanger, F., Nicklen, S. and Coulson, A. R. (1977) Proc. Natl. Acad. Sci. USA 74, 5463-5467 Remington, R. D. and Schork, M. A. (1985) Statistics with applications to the biological and health sciences, pp. 296-309, Prentice-Hall, NJ Brooks, S. (1992) Biotechniques 13, 906-911 Smith, P. K., Krohn, R. I., Hermanson, G. T., Mallia, A. K., Gartner, F. H., Provenzano, M. D., Fugimoto, E. K., Goele, N. M., Olson, B. J. and Klenk, D. C. (1985) Anal. Biochem. 150. 76-85 Habeeb, A. F. S. A. (1972) in Methods in enzymology vol. 25: Enzyme structure, part B (Hirs, C. H. W. and Timasheff, S. N ., ed.), pp. 457-459, Academic press, NY Miles, E. W. (1977) in Methods in enzymology vol. 47 (Hirs, C. H. W. and Timasheff, S. N., ed.), pp. 431-442, Academic press, NY Lemesle-Verloot, L., Henrissat, B., Gaboriaud, C., Bissery, V., Morgat, A. and Mornon, J. P. (1990) Biochimie 72, 555-574 Hawley, D. K. and McClure, W. R. (1983) Nucl. Acids Res. 11, 2237-2255 Wierenga, R. K. and H01, G. J. (1983) Nature 302, 842-844 Matthews, B. W., Nicholson, H. and Becktel, W. ( 1987) Proc. Natl. Acad. Sci. USA 84, 6663-6667 Jendrossek, D., Steinbuechel, A. and Schlegel, H. G. (1988) J. Bacteriol. 170, 5248-5256 Heming, T., Katsuhide, Y., Taniyama, Y. and Kikuchi, M. (1991) Biochemistry 30, 9882-9891 Heming, T., Katsuhide, Y., Inaka, K., Matsushima, M. and Kikuchi, M. (1992) Biochemistry 31, 7077-7085 Eijsink, V. G. H., Vriend, G., van de Burg, B., van de Zee, J. R., Veltman, O. R., Stulp, B. K. and Venema, G. (1993) Protein Engineer. 5, 157-163 38. 39. 40. 41. 42. 43. 119 Watanabe, K., Masuda, T., Ohashi, H., Mihara, H. and Suzuki, Y. (1994) Eur. J. Biochem. 226, 277-283 Watanabe, K., Chishiro, K., Kitamura, K. and Suzuki, Y. (1991) J. Biol. Chem. 266, 24287-24294 Hardy, F., Vriend, G., Veltman, O. R., van de Vinne, B., Venema, G. and Eijsink, V. G. H. (1993) FEBS Lett. 317, 89-92 Segel, I. H. (1975) in Enzyme kinetics: behavior and analysis of rapid equilibrium and steady-state enzyme systems, pp. 929-934, John Wiley & Sons, NY Londesborough, J. (1980) Eur. J. Biochem. 105, 211-215 Vieille, C., Hess, J. M., Kelly, R. M., and Zeikus, J. G. (1995) Appl. and Environ. Microbiol. 61, 1867-1875 Jencks, W. P. (1975) in Advances in enzymology and related areas of molecular biology, vol. 43, (Meister, A., ed.), pp. 219-411, John Wiley & sons, NY Chapter 111 Effect of thermal and chemical denaturants on Thermoanaerobacter ethanolicus 39E secondary-alcohol dehydrogenase stability Prepared for submission to Biotechnology and Applied Biochemistry 120 121 ABSTRACT Thermoanaerobacter ethanolicus secondary-alcohol dehydrogenase (2° ADH) was optimally active near 90°C with calculated thermostability half-lives of 1.2 days, 1.7 h, 19 min, 9.0 min, and 1.3 min at 80°C, 90°C, 92°C, 95°C, and 99°C, respectively. Enzyme activity loss upon heating (90-100°C) was accompanied by precipitation, but the soluble enzyme remaining after partial sample inactivation retained complete activity, mechanistically linking enzyme inactivation and precipitation. Enzyme thermoinactivation was modeled by a pseudo-first order rate equation suggesting that the rate determining step was unimolecular with respect to protein, thus, also implying that thermoinactivation preceded aggregation. Structural unfolding of the 2° ADH occurred with a Tm (50% unfolding) at ~115°C. This Tm was much higher than the temperature for maximal activity, indicating that the 2° ADH was rigid and tightly folded throughout its active temperature range. Thermodynamic calculations indicated that the active folded structure of the 2° ADH, like other proteins, is stabilized by a relatively small Gibbs energy (AGismb, = 110 kl mol'l). 2° ADH was also active in > 1.0 M GuHCl, demonstrating its high resistance to chemical denaturation. Enzyme activities were increased 2-fold in low guanidine hydrochloride (GuHCl) concentrations (150-230 mM). The additive denaturing effects of GuHCl and temperature on T. ethanolicus 2° ADH TmlGu] predict that at higher temperatures lower GuHCl concentrations should be required to reduce enzyme rigidity, thus lower GuHCl concentrations should result in optimal enzyme flexibility and activity. The increasing Gui-1C1 concentration that provided maximal 2° ADH catalytic rate enhancement at higher temperatures was also seen with weakly chaouopic (KN 03) or neutral (KCl) inorganic salts, suggesting that this 2° ADH catalytic rate enhancement results from specific ionic and not hydrophobic interactions. 122 INTRODUCTION As industrial catalysts, enzymes must have robust activities that are stable under process chemical and thermal conditions. The saccharidases used in starch processing and the proteases used in detergents typify current industrial enzyme standards. The potential biotechnological application of enzymes as chiral chemical catalysts has long been recognized [1-10]. The chiral synthesis of high value bioactive compounds such as pharmaceuticals is increasingly important due to the recent FDA ruling mandating either . enantiomeric purity or exhaustive toxicity testing to demonstrate the lack of any detrimental effects due to enantiomeric impurities [11]. Traditional chemical synthetic methods can often only ensure chiral purity using chiral feedstock chemicals because racemate resolution is often too difficult and expensive. Thus, the economic biocatalytic synthesis of either chiral feedstock chemicals or of the final chiral products can significantly advance current technologies. Alcohol dehydrogenases (ADHs) are almost exclusively nicotinamide cofactor dependent, enantioselective enzymes with broad specificity [12]. While NAD(P)(H) is costly, the demonstrated effectiveness of numerous NAD(P)(H) retention and redox recycling strategies has removed this economic barrier to the scale-up of even moderate value adding processes [9,10,13]. Analytical scale chiral alcohol formation by 1° ADHs has been demonstrated for a diverse set of substrates [14.5.6]. These 1° ADHs are far less active on 2° alcohols or ketones than on 1° alcohols or aldehydes. The emerging class of 2° ADHs have greater catalytic efficiency for 2° alcohol oxidation and ketone reduction than for 1° alcohol oxidation and aldehyde reduction and they display broad substrate specificities [15-17]. The potential for 2° ADH catalyzed chiral chemical syntheses has been demonstrated using the Thermoanaerobacten'wn brockii and Thermoanaerobacter ethanolicus enzymes [18-22]. 123 The kinetic similarity of the 2° ADHs isolated from a number of similar thermoanaerobes led Nagata et al. to conclude that the enzymes were also structurally similar [15]. A comparison of the T. brockl'l' 2° ADH peptide sequence, determined by Edman degradation, to that of liver 1° ADH predicted some potential structure function similarities [21], however, without a 2° ADH clone these predictions could not be tested. Recently, the ath gene encoding the T. ethanolicus 2° ADH was cloned, characterized, and overexpressed, demonstrating that the T. erhanolicus and T. brockii 2° ADHs differed by only 3 amino acids [23]. Therefore, research using the T. ethanolicus 2° ADH may provide knowledge representative of other 2° ADHs. Thermozymes [24], with their greater resistance to both thermal and chemical denaturation than mesophilic enzymes, may alleviate some of the limiting process conditions for applications of enzymes and enhance enzyme lifespans. Irreversible mesophilic and thermophilic protein thermoinactivation has been shown to result from the loss of folded protein structure and from covalent peptide or amino acid destruction [see 24], suggesting that the higher stability of thermozymes is due to interactions mechanistically similar but stronger than those stabilizing mesophilic enzymes [see 24]. Thermophilic enzymes recombinantly expressed in mesophilic hosts retain their thermal characteristics, indicating that the peptide sequences encode these properties. Comparative analyses of thermophilic and mesophilic enzymes have typically concluded, from the lack of exotic amino acids or 1° structural differences, that the added stability of one enzyme compared to another results from numerous subtle differences in the folded protein structure [24,25]. Enzyme thermostabilizing mechanism theories including increased core hydrophobicity and atomic density, additional salt bridges, 2° structural element stabilization, and constrained surface loop architecture have been reported [see 24]. Detailed studies on thermal and chemical denaturation of a thermophilic 2° ADH have not been reported The data presented here measure T. ethanolicus 2° ADH thermophilicity, thermostability, and stability to denaturation by GuHCl. The relationship 124 between T. ethanolicus 2° ADH rigidity and thermostability is determined through kinetic and biophysical characterization of enzyme activity, precipitation, and unfolding. Also, the effect of the chemical denaturant guanidine hydrochloride (GuHCl) on the activity of this robust biocatalyst is examined. MATERIAL AND METHODS Chemicals and reagents All chemicals were of at least reagent/molecular biology grade. Gases were provided by AGA specialty gases (Cleveland, OH) and made anaerobic by passage though heated copper filings. Spectrophotometrically pure GuHCl was obtained as an 8.0 M aqueous solution (Cat. #24115: Pierce; Rockford, IL). Tris buffer pH values were measured at 25°C and formulated for pH 8.0 at the temperature they were used (thermal correction factor = -0.031 ApH °C'1). Media and strains Escherichia coli (DHSa) containing the 2° ADH recombinant plasmid [23] was grown in rich, complex medium (20 g 1'1 tryptone, 10 g l-1 yeast extract, 5 g 1'1 NaCl) at 37°C with 25 tlg ml'1 kanamycin and 100 ug ml-1 ampicillin. Enzyme purification The recombinant 2° ADH was purified from E. coli (DHSa) aerobically. The pelleted cells from batch cultures were resuspended (0.5 g wet wt. ml'l) in buffer A [50 mM Tris:HC1 (pH 8.0), 5 mM DTT, and 10 11M ZnClz} and lysed by passage though a French pressure cell. The clarified lysate was incubated at 65°C for 25 min then centrifuged for 30 min at 15000g. The clarified supernatant was applied to a DEAE-sephacryl column (2.5 cm x 15 125 cm) equilibrated with buffer B {50 mM TriszHCI (pH 8.0)}. Active fractions were diluted 4-fold in buffer B and applied to a Q-sepharose column (2.5 cm x 10 cm) equilibrated with buffer B. Purified recombinant enzyme was eluted using a 250 ml NaCl gradient (0-300 mM). Enzyme activity The standard assay for 2° ADH activity was defined as the reduction of NADP+ using propan-2-ol at 60°C as previously described [17]. Kinetic parameters were calculated from nonlinear best fits of the data to the Michaelis-Menten equation using Kinzyme software [26] on an IBM PC. Protein concentrations were measured in all cases by the bicinchoninic acid (BCA) procedure (Pierce; Rockford, IL). The effect of GuHCl on 2° ADH activity was determined at 37°C, 50°C, 60°C, 70°C, and 75°C in the absence or presence of 100 mM, 200 mM, 400 mM, 800 mM, or 1.6 M GuHCl. ‘ Ionic activities Guanidinium ionic activities were determined using the Debye-Hiickel limiting law [see 27]. The mean ionic activity (at) is defined as 3:1: = 711111: where rm; is the mean electrolyte molality and 71 is the mean activity coefficient. The mean activity coefficient is determined as a function of temperature (T), ionic charge (2), dielectric constant of water (a = 78.54), and ionic strength (I) using the equation: log y: = {-1.824x106/[(er)(3/2)] } (lz.,.z.l)[(I)(1/2)] where 126 I = (1/2)Z(mi2i2). The GuHCl concentration dependent enzyme activity experiments were performed in 45 mM to 50 mM Tris (pH 8.0) at 37°C, 50°C, 60°C, 70°C, and 75°C. Tris buffer was assumed to be 50% dissociation (pKa = 8.1). Because of GuHCl's pKa (> 10) it was assumed to be completely ionized. All species ionic charges were 11. Thermostability Purified recombinant 2° ADH thermostability was evaluated by timed incubation at the desired temperatures, followed by incubation for 30 min at 25 °C. Samples were then equilibrated at 55°C for at least 15 min prior to activity determination. Activity was determined using the unfractionated samples. Incubations were performed in 100 til PCR tubes (cat. #72.733.050, Sarstedt; Newton, NC) using 0.2 mg ml'1 protein in 100 pl of buffer B. Protein precipitation due to heating for 2 min at 99°C was quantified in similar experiments by fractionation of the soluble and aggregated protein. First, the 70 pl of sample remaining after thermostability activity determination was centrifuged for 5 min at 12000g. The supernatant was then removed from the pelleted protein aggregate, the pellet was resuspended in 100 [ll of buffer B, centrifuged (12000g, 5 min), the wash buffer removed, and the repelleted protein resuspended in 70 ul of buffer B. Enzyme activities and protein concentrations were determined for the initial sample, supernatant, pellet wash, and the resuspended pellet. Specific acivities and tom] protein values from the pellet wash and resuspended pellet fractions were combined and reported as the precipitate specific activity and total protein, respectively. 127 Heat induced enzyme precipitation Heat-induced enzyme precipitation was monitored from 25 °C to 100°C by light scattering (it = 5 80 nm), using protein solutions (0.2 mg ml'l) in buffer B. Absorbance measurements were conducted in 0.3 ml quartz cuvettes (pathlength = 1.0 cm), using a Gilford Response spectrophotometer (Coming; Oberlin, OH) equipped with a Peltier cuvette heating system. The apparent OD530 increase is due to light scattering from protein precipitation. Heat-induced enzyme unfolding 2° ADH unfolding experiments were conducted in buffer B in the absence of GuHCl and in the presence of 2.0—6.0 M GuHCl. Unfolding was monitored from 25°C to 100°C by absorbance at 280 nm [28-30] in 0.3 ml quartz cuvettes (pathlength = 1.0 cm), using a Gilford Response spectrophotometer equipped with a Peltier cuvette heating system. Melting temperature (Tm) is defined as the temperature of the absorbance transition midpoint during an increasing thermal gradient (1.0°C min'l). The 95% confidence interval error boundaries for the predicted 2° ADH Tm were determined by standard statistical treatment of the data [31]. Fluorescence spectrometry Fluorimetric unfolding measurements were conducted with protein solutions (0.2 units O.D.230) in buffer C [10 mM Tris:HCl (pH 8.0)] at 25°C. Fluorescence spectra “excitation = 295 nm, Aemission = 315-450 nm) were recorded in 2.25 ml quartz cuvettes (pathlengths: excitation = 0.5 cm, emission = 1.0 cm), using an SLM spectrofluorimeter (Champain-Urbana, Il) equipped with a thermal, cell holder. Decreasing flomescence intensity indicates exposure of the aromatic amino acid residues to the polar solvent environment. 128 RESULTS Thermophilicity and thermostability T. ethanolicus 2° ADH activity increased with temperature to at least 90°C with no activity observed at 100°C (Fig. 1). Accrnately quantitating enzyme activity at temperatures above 90°C was difficult due to NADP(H) instability and equipment constraints, however, no product formation was detected during reactions at 100°C. Recombinant 2° ADH thermostability was determined at 65°C, 80°C, 90°C, 92°C, 95°C, and 99°C (Fig. 2). The enzyme was stable at 65°C, and, it had calculated half-lives of 1.2 days, 1.7 h, 19 min, 9.0 min, and 1.3 min at 80°C, 90°C, 92°C, 95°C, and 99°C, respectively. The rate of thermoinactivation was described by equation 1 for all temperature measured (k=rate constant and Ftime (min)). ln(residual activity) = 1n(irlitial activity) - kt (1) This pseudo first order relationship indicated that the slow step in T. ethanolicus 2° ADH thermoinactivation did not depend on enzyme concentration. Arrhenus analysis of the inactivation rate constants (6.70x10‘6 SCC’l, 1.34le seC'l, 6.66x10'4 seC'l, 1.60x10‘3 seC'l, and 1.87x10'2 sec1 at 80°C, 90°C, 92°C, 95°C, and 99°C, respectively) yielded a transition state Gibbs energy barrier of 110 kJ mol'1 resulting from the difference between AH? = 440 kJ mol'1 and A81 = 330 kJ mol'1 (temperature = 363 K). Thermostability and unfolding Enzyme inactivation was accompanied by precipitation and the effect of protein aggregation on enzyme activity was investigated. Spectrophotometric determination of heat-induced protein precipitation using a 10°C per minute thermal gradient indicated that the enzyme 129 Figure l Recombinant T. ethanolicus 39E 2° ADH thermophilicity. 2° ADH propan-2-ol oxidation activity was measured at temperatures spanning 30°C to 100°C. 130 100 80- 60" 40- 20- A539:— uEEEE a=>39< 958.5 120 100 80 60 40 20 Temperature (°C) 131 Figure 2 Recombinant T. ethanolicus 2° ADH thermostability. 2° ADH residual activity was determined after timed (1 min to 100 min) incubations at 65°C (0), 80°C (0), 90°C (A), 92°C (0), 95°C m, and 99°C a3). 132 A539:— 9532:: 3:235 #:90an us— 100 80 60 4O 20 (min) Time 133 began to precipitate at temperatures above 85°C (Fig. 3). Precipitation was a rapid, cooperative process that was essentially complete at 100°C with the transition midpoint at ~95°C. The potential correlation between enzyme precipitation and loss of catalytic activity was then investigated by direct measurement (Table 1). After heat treatment for 2 min at 99°C (approximately the enzyme half-life at this temperature), the specific activity of the unfractionated heat treated sample was half that of the unheated enzyme. Of the 11 i 1.1 [lg of protein originally added, 10 i 1.0 pg were recovered after heat treatment. Subsequent fractionation of the remaining sample (70 pl of the original 100 111) by centrifugation indicated that the supernatant contained half of the initial enzyme. This soluble enzyme had similar specific activity to the unheated sample. The resuspended pellet, which did not redissolve, also contained half of the protein but displayed no detectible enzyme activity. Once precipitate had formed, enzyme continued to inactivate and aggregate slowly upon standing at room temperature (data not shown). The role of protein unfolding in 2° ADH thermoinactivation was investigated by measuring the effect of GuHCl on temperature dependent enzyme unfolding. Spectrophotomeuically determined transition midpoints for protein unfolding (using a 1.0°C per minute thermal gradient) with 2.0 M, 2.25 M, 2.5 M, 2.75 M, and 3.0 M added GuHCl yielded melting temperature (TmlGul) values of 68°C, 59°C, 58°C, 53°C, and 48°C, respectively. Tm is defined as the midpoint of the unfolding transition and is interpreted as the point where the protein is 50% unfolded. The inset in figme 4 of a representative 2° ADH melting curve used to obtain the Tm[Gu] values indicates that unfolding was a rapid, c00perative process. At GuHCl concentrations below 2.0 M enzyme precipitated prior to completing the unfolding transition so no midpoint temperature could be identified. Also, at concentrations above 3.0 M the absorbance change registered upon 2° ADH denaturation became to small for reliable Tm[Gu] determination. Enzyme Tm[Gu] values decreased linearly with increasing GuHCl concentrations (Fig. 4). The GuHCl dependence of 2° 134 Figure 3 Temperature dependence of 2° ADH precipitation. 2° ADH precipitation during an incresaing thermal gradient (1.0°C min-1) was determined by sample absorbance at 580 nm. OD. (580 nm) 135 Temperature (°C) 100 136 6088833 803 888.888 5208 E .388 088:» 0388 880382. 05 :ounamom “gage e8 888695 3 hem: .ASegoeSh 8 BE 0388 20388 05 we mum—:8 :8 3588 293 3 cc 83.9 eoeoee a 8: was as 9 so... menace: .3 B .8838 one as? seen eosaoa Bob .8on RE 05 e5 :33 05 8 805 .«o 88 05 8a Seen—E Egon 05 you concav— moea> 2F .:eeaw_£b:oo .3 30:25: on: 8.95 five .23 3:33 83 88880.:— eoeo=om 2F 4:288: :8; Bee Summer—:50 .3 ages—09$ ageE 0528 05 88a 3883 «:3 aux—Re 3538 33860.8 25. .58 N to: come a 3385 2o.» m 88.. a moans soon 38:. need a M: £3 a Ed 3 a. 2 1385 Se v 283 v serene.“ see one a m Sod a w; some a cw assuage nae? 2 H mm 8.3 a ”8 82. a : :8: one sec 9.83 Ame been ocean area Boa £2883 $383.85 38.—2: Ina. ou guagefie H :m 53:338.:— 538: he 23— n 935—. 137 Figure 4 Effect of GuHCl on T. ethanolicus 2° ADH melting temperature. The Tmmu] values from absorbance were measured at 280 nm during an increasing thermal gradient (1.0°C min'l) with 2.0 M, 2.25 M, 2.5 M, 2.75 M, and 3.0 M added GuHCl. The inset shows percent protein unfolding versus temperature for 2° ADH in 2.5 M GuHCl. 138 wwwwaw 33.-=2.— T2 120 100- 80q 60- 40' Gov 9.59—2.83 ant—o: 20" GuHCl concentration (M) l 39 ADH denaturation determined by fluorescence intensity at 25 °C indicated a melting transition midpoint at 3.7 M (Fig. 5). Extrapolating to zero GuHCl from both the absorbance and fluorimetric Tm[Gu] data predicts a 2° ADH Tm of 115 i 5.8°C (at the 95% confidence level). Chemical stability and 2 °ADH activity The generally poor low temperature activity of thermozymes has been attributed to their high rigidity at these temperatures [32]. Therefore, reducing enzyme rigidity by adding a chemical denaturant would be expected to enhance thermozyme low temperature activity. The chemical denatm'ant GuHCl and thermal energy were also used to study the effect of enzyme rigidity on catalysis. Figure 6 shows the effect of GuHCl concentrations up to 1.6 M on enzyme activity at temperatures between 37°C and 75°C. Maximal enzyme specific activity was seen in the presence of 150 mM to 230 mM GuHCl at all temperatures tested and it was approximately doubled that in the absence of GuHCl. The GuHCl concentration corresponding to maximal enzyme activity increased with increasing temperature (147 mM, 184 mM, 206 mM, 237 mM, and 231 mM at 37°C, 50°C, 60°C, 70°C, and 75°C, respectively). The mean ionic activity coefficients (7:) calculated from the Debye-Hiickel limiting law for these maximal activity conditions averaged to 0.808 1!: 0.0051. The calculated GuHCl activities also increased with increasing temperature (120 mM, 149 mM, 165 mM, 183 mM, and 187 mM at 37°C, 50°C, 60°C, 70°C, and 75°C, respectively). Similar but less dramatic low concentration 2° ADH rate enhancements were seen for added KCl and KN03 but not for K2804 (data not shown). 140 Figure 5 Effect of GuHCl on T. ethanolicus 2° ADH fluorescence. The flourescenoe measurements (lacim=295 nm, Missim=340 nm) at 25 °C were conducted using protein in GuHCl concentrations from 0.0 to 6.0 M. Relative Florescence 141 GuHCl concentration (M) 142 Figure 6 GuHCl and temperature dependence of 2° ADH activity. 2° ADH catalyzed propan-2-ol oxidation rates were determined at 37°C (A), 50°C (0), 60°C (0), 70°C a) and 75°C (0) in GuHCl concentrations from 0.0 to 1.6 M. 143 100 q 0 8 A589:— wEEEB .3333. 95025 GuHCl concentration (M) 144 DISCUSSION These data provide quantitative evidence that the T. ethanolicus 2° ADH is a highly thermophilic and thermostable enzyme with catalytic activity that is also robust to high levels of chemical denaturants. The enzyme was rapidly thermoinactivated only at temperatures above 90°C. Enzyme thermoinactivation was modeled by a pseudo-first order rate equation suggesting that the rate determining step was unimolecular with respect to protein. The large thermoinactivation transition state enthalpy and entropy changes are consistent with a significant protein structural change during this rate limiting step. Enzyme activity loss upon heating was accompanied by precipitation, but the soluble enzyme remaining after partial sample inactivation retained complete activity, mechanistically linking enzyme inactivation and precipitation. Finally, the increased GuHCl concentration necessary for enzyme rate enhancement at higher temperatures was also seen with certain inorganic salts arguing that this 2° ADH catalytic property results from specific ionic and not general hydrophobic interactions. The highly thermophilic T. ethanolicus 2° ADH has a projected half-life of 2 months at 70°C which compares to the complete inactivation of the structurally similar mesophilic Closm'diwn beijen'nckii enzyme [23,33] is also more stable than the T. brocla'i 2° ADH [33a] that differs by only 3 amino acids. Tomazic and Klibanov proposed that enzyme molecules reversibly enter structural states that can lead to essentially irreversible protein aggregation and inactivation [34]. With respect to protein, this process involves an initial intramolecular structural change followed by an intermolecular aggregation step. Thermoinactivation of the T. ethanolicus enzyme follows a pseudo-first order rate law, suggesting that the rate is independent of the protein concentration as reported for a- amylase [34] and D-xylose isomerase [35]. Thus, the process appears to be unimolecular with respect to 2° ADH protein. The soluble enzyme recovered after incomplete enzyme thermoinactivation retained specific activity comparable to the unheated enzyme, implying 145 complete inactivation of part of the enzyme population and not partial inactivation of the entire enzyme population. These data suggest that catalytically compromised 2° ADH intermediate structures did not persist, consistent with a rapid second step in irreversible thermoinactivation (precipitation). The absence of detectible enzyme activity in the precipitated protein demonstrated that loss of catalytic activity either accompanied or preceded protein precipitation. Therefore, the observed unimolecular T. ethanolicus 2° ADH thermoinactivation rate limiting step preceding precipitation of structurally scrambled protein molecules fits the two step Tomazic and Klibanov model [34] for protein thermoinactivation with a slow native to non-native enzyme structural transition followed by rapid aggregation of thermoinactivated enzyme. Identifying the exact molecular mechanism of irreversible 2° ADH thermoinactivation will require precise measurement of inactivated protein molecular mass or detection of covalent destruction products like ammonia to determine whether this rate limiting step involves covalent protein modification (eg., deamidation) or partial loss of native enzyme structure. The inactivation slow step transition state thermodynamics were evaluated to determine the stabilizing energy for the 2° ADH. The transition state enthalpy, almost 20- fold that reported for enzymatic catalysis [23], is lOO-fold greater than the expected molecular kinetic energy (~RT; where R = 8.341 J mol'1 K4, and T is the system temperature in Kelvin). Calculating the astmbmty from absolute rate theory requires the assumption that the inactivated, partly unfolded protein is an intermediate following the slow step transition state which is itself in rapid equilibrium with the native protein structure. Assuming this mechanism, the 2° ADH is stabilized by a relatively small energy difference (110 kJ mol'l) between large enthalpic (440 kJ mol'l) and entropic (330 1:] mol' 1 at 393 K) contributions as described for other proteins [36]. The AI-Ii therefore, appears to be an effective barrier to 2° ADH thermoinactivation opposing a large destabilizing transition state entropy change. 146 Unfolding of T. ethanolicus 2° ADH was measured to examine the nature of this slow, inactivating 2° ADH structural transition. The Tmmu] data from absorbance measurements predicted that a GuHCl concentration of 3.75 M would yield an enzyme Tm[Gu] value of 25 °C. This predicted value was corroborated by Trp fluorescence measurements. 2° ADH melting measured by both absorbance and fluorescence indicated an essentially linear Tm[Gu] dependence on GuHCl concentration. Extrapolatin g this linear dependence of enzyme Tm[Gu] on GuHCl concentration to zero GuHCl predicts a Tm value of 115 3‘- 5.8°C. This temperature is 20°C higher than the midpoint temperature for 2° ADH precipitation (~95°C) arguing that the precipitating protein is not significantly unfolded. Consequently, the high thermal stability of T. ethanolicus 2° ADH is associated with its high resistance (i.e., rigidity) to chemically and thermally induced unfolding. Biophysical measurements have indicated that enzymes are more rigid at low temperatmes than at high temperatures within their active temperature ranges [25.36.37]. However, the hypothesis that poor low temperature enzyme activity results from excessive structural rigidity [32] has been challenged based on the general Arrhenius theory adherence of enzyme catalytic rate temperature dependence [24,38] and specifically based on the temperature activity behavior of the T. ethanolicus 2° ADH [25]. Unlike other enzymes, the 2° ADH was totally folded at 90°C where it was optimally active. Here we attempted to alter 2° ADH rigidity by the addition of GuHCl. The additive denaturing effects of GuHCl and temperature on T. ethanolicus 2° ADH Tm[Gu] predict that at higher temperatures lower GuHCl concentrations should be required to reduce enzyme rigidity, thus lower (3qu concentrations should result in optimal enzyme flexibility and activity. While added GuHCl did significantly enhance enzyme activity, the effect was equally pronounced at low and at high temperatures. The slightly increased denaturant concentration corresponding to maximal 2° ADH activity at higher temperatures is inconsistant with the lower denaturant concentration predicted to yield optimal enzyme rigidity at higher temperatures. Similar 2° ADH activity enhancement was seen with low 147 concentrations of inorganic salts which are not strong protein denaturants, supporting the conclusion that ionic interactions and not optimal rigidity were responsible for the observed increase in the catalytic rate. Industrial biocatalysts must have activities robust to a broad range reaction conditions. Both chemical and thermal stability are therefore important properties of a potential industrial chiral catalyst. The T. ethanolicus 2° ADH is active from 30°C to over 90°C with active half-lives greater than 1 hr to temperatures exceeding 90°C. GuHCl concentrations exceeding 1.0 M were necessary to eliminate enzyme activity and this thermophilic 2° ADH was also highly active in ethanol concentrations greater than 2.0 M (data not shown). Therefore, the extremely high thermal and chemical stabilty of T. ethanolicus 2° ADH activity, demonstrates that this is an extremely robust enzyme catalyst. As a chiral chemical catalyst, the effect of temperature on T. ethanolicus 2° ADH chirality has been shown to be reaction specific [39]. The effect of chemical denaturants on 2° ADH product chirality however, remains to be determined. ACKNOWLEDGMENTS I gratefully acknowledge Dr. Vladimir Tchemajencko for conducting the structural melting experiments. I also gratefully acknowledge Maris Laivenieks for his assistance in protein purification. This research was supported by a grant from the Cooperative State Research Service, US Department of Agriculture, under the agreement 90-34189-5014. 10 11 12 13 14 15 148 REFERENCES Whitesides, G. M. and Wong, C.-H. [1983] Enzymes as catalysts in organic synthesis. Aldrichimica Acta 16, 27-34 Rosazza, J. P. N. [1995] Biocatalysis, microbiology, and chemistry: The power of positive linking. Amer. Soc. Microbial. News 61, 241-245 Meyer, H.-P. [1991] Microbiology in the contemporary organic chemical industry. Biol. Ferm. Eng. 8, 602-606 Wong, C.-H. [1989] Enzymatic catalysts in organic synthesis. Science 244, 1 145-1 152 Jones, J. B. [1986] Enzymes in organic synthesis. Tetrahedron 42, 3351-3403 Irwin, J. B., Lok, K. P., Huang K. W. C. and Jones, J. B. [1978] Enzymes in organic synthesis. Influence of substrate structure on rates of horse liver alcohol dehydrogenase-catalyzed oxidoreductions. J. Chem. Soc. Perkin I 12, 1636-1641 Bradshaw, C. W., Hummel, W. and Wong, C.-H. [1992] Lactobacillus kefir alcohol dehydrogenase: a useful catalyst for synthesis. J. Org. Chem. 57, 1532- 1536 Bradshaw, C. W., Fu, H., Shen, G.-J. and Wong, C.-H. [1992] A Pseudomonas sp. alcohol dehydrogenase with broad substrate specificity and unusual stereospecificity for organic synthesis. J. Org. Chem. 57, 1526-1532 Hummel, W. and Kukla, M. R. [1989] Dehydrogenases for the synthesis of chiral compounds. Eur. J. Biochem. 184, l-13 May, SW. and Padgette, SR. [1983] Oxidoreductase enzymes in biotechnology: current status and future potential. Bio/Technology 1, 677-686 Faber, K. and Franssen, M. C. R. [1993] Prospects for the increased application of biocatalysts in organic transformations. Trends Biotechnol. 11, 461-470 Brandén, C.-I., Jornvall, H., Eklund, H. and Fururgren, B. [1975] Alcohol dehydrogenases, in The enzymes, 331 ed., vol. XI part A (P. D. Boyer, ed.) pp. 105-106. Academic Press, New York Klibanov, M. [1983] Biotechnological potential of the enzyme hydrogenase. Proc. Biochem. 18, 13-23 Zong, M.-H., Fukui, T., Kawamoto, T. and Tanaka, A. [1991] Bioconversion of organosilicon compounds by horse liver alcohol dehydrogenase: the role of the silicon atom in enzymatic reactions. Appl. Microbiol. Biotechnol. 36, 40-43 Nagata, Y., Maeda, K. and Scopes, R. K. [1992] NADP-linked alcohol dehydrogenases from extreme thermophiles: Simple affinity purification schemes, and gomparative properties of the enzymes from different strains. Bioseparation 2, 353- 62 16 17 18 19 20 21 22 23 24 25 26 27 28 149 Keinan, B., Hafeli, E. K., Seth, K. K. and Lamed, R. L. [1986] Thermostable enzymes in organic synthesis. 2. Asymmetric reduction of ketones with alcohol dehydrogenase from Thermoanaerobium brockii. J. Am. Chem. Soc. 108, 162-169 Burdette, D. S. and Zeikus, J.G. [1994] Purification of acetaldehyde dehydrogenase and alcohol dehydrogenases from Thermoanaerobacter ethanolicus 39B and characterization of the secondary-alcohol dehydrogenase (2° ADH) as a bifunctional alcohol dehydrogenase-acetyl-CoA reductive thioesterase. Biochem. J. 302, 163-170 Zheng, C., Pham, V. T. and Phillips, R. S. [1992] Asymmetric reduction of ketoesters with alcohol dehydrogenase from Thermoanaerobacter ethanolicus. Bioorg. Med. Chem. Let. 2, 619-622 Keinan, B., Seth, K.K. and Lamed, R. [1986] Organic synthesis with enzymes. 3. TBADH-catalyzed reduction of chloro ketones. Total synthesis of (+)- (S ,S)-(cis-6-methyltetrahydropyran-2-yl)acetic acid: a civet constituent. J. Am. Chem. Soc. 108, 3474-3480 Keinan, B., Seth, K.K., Lamed, R., Ghirlando, R. and Singh, SP. [1990] thermostable enzyme in organic synthesis, 4. TBADH-catalyzed preparation of bifunctional chirons. Total synthesis of S-(+)-Z=tetradec-5-en-13-olide. Biocatalysis 4, 1-15 Peretz, M. and Burstein, Y. (1989) Amino acid sequence of alcohol dehydrogenase from the thermophilic bacterium Thermoanaerobium brockii. Biochemistry 28, 6549-6555 Lamed, R.J., Keinan, E. and Zeikus, J.G. [1981] Potential applications of an alcohol-aldehyde/ketone oxidoreductase from thermophilic bacteria. Enz. Microbiol. T echnol. 3, 144-148 Burdette, D.S., Vieille, C. and Zeikus, J .G. [1996] Cloning, expression, and biochemical characterization of the thermophilic secondary-alcohol dehydrogenase from Thermoanaerobacter ethanolicus 39E. Biochem. J ., in the press Vieille, C., Burdette, D. S. and Zeikus, J. G. [1996] Thermozymes, in Biotechnology Annual Reviews, vol. 2, (M. R. El Geweley, ed.) pp. 1-83. Elsevier Press, Amsterdam, Neth. Bbhm, G. and Jaenicke, R. [1994] Relevance of sequence statistics for the properties of extremephilic proteins. Int. J. Peptide Protein Res. 43, 97-106 Brooks, S. [1992] A simple computer program with statistical tests for the analysis of enzyme kinetics. Biotechniques 13, 906 - 911 Chang, R. [1977] Physical chemistry with applications to biological systems. MacMillan Pub. Co., NY Hermans, J. and Scheraga, H. A. [1961] Structural studies of ribonuclease V. Reversible changes of configuration. J. Am. Chem. Soc. 83, 3283-3292 29 30 31 32 33 33a 34 35 36 37 38 39 150 Tiktopulo, E. I. and Privalov, P. L. [1974] Heat denaturation of ribonuctease. Biophys. Chem. 1, 349-357 Klump, H., Di Ruggiero, J., Kessl, M., Park, J.-B., Adams, M. W. W. and Robb, F. T. [1992] Glutamate dehydrogenase from the hyperthermophile Pyrococcus furiasis J. Biol. Chem. 267, 22681-22685 Remington, R. D. and Schork, M. A. [1985] Statistics with applications to the biological and health sciences, Prentice-Hall, Engelwood Cliffs, NJ Wrba, A., Schweiger, A., Schultes, V., Zévodszky, P. and Jaenicke, R. [1990] Extremely thermostable D—glyceraldehyde-3-phosphate dehydrogenase from the eubacterium Thermatoga matin'ma. Biochemistry 29, 7584-7592 Ismaiel, A. A., Zhu, C.-X., Colby, G. D. and Chen, J .-S. [1993] Purification and characterization of a primary-secondaryalcohol dehydrogenase from two strains of Clastridium beijerinckii. J. Bacteriol. 175, 5097-5105 Lamed, R. J. and Zeikus, J. G. [1981] Novel NADP-linked alcohol aldehyde/ketone oxidoreductase in thermophilic ethanologenic bacteria. Biochem J. 195, 183-190 Tomazic, SJ, and Klibanov, AM. [1988] Mechanisms of irreversible thermal inactivation of Bacillus a-amylases. J. Biol. Chem. 263, 3086-3091 Meng, M., Bagdasarian, M. and Zeikus, J. G. [1993] The role of active-site aromatic and polar residues in catalysis and substrate discrimination by xylose isomerase. Biotechnology 11, 1157-1161 J aenicke, R. [1991] Protein stability and molecular adaptation to extreme conditions. Eur. J. Biochem. 202, 715-728 Fontana, A. [1990] How natrne engineers protein (thermo) stability, in Life under extreme conditions: Biochemical adaptation (G. di Prisco, ed.) pp. 89-113. Springer-Verlag, Heidelberg. Tchemajenko, V., Vieille, C., Burdette, D.S., and Zeikus, J .G. [1996] Physicobiochemical studies of xylose isomerases from thermophilic and mesophilic microorganisms. Manuscript in preparation Phillips, R. S., Zheng, C., Pham, V. T., Andrade, F. A. C. and Andrade, M. A. C. (1994) Effects of temperature on the ctereochemistry of enzyme reactions. Biocatalysis 10, 77-86 Chapter IV Mutagenic and Biophysical Analysis of Thermoanaerobacter ethanolicus Secondary-Alcohol Dehydrogenase Activity Prepared for submission to the Journal of Biological Chemistry 151 152 ABSTRACT The Thermoanaerobacter ethanolicus 398 ath gene encoding the secondary- alcohol dehydrobenase (2° Adh) was overexpressed in Escherichia cali (DH5a) at ~15% of total protein. Recombinant enzyme was purified in high yield (67%) by simple heat treatment at 85°C and ammonium sulfate precipitation. Purified site directed 2° Adh mutants Cys37 to Ser, His59 to Asn, Asp150 to Asn, Asp150 to Glu, and Asp150 to Cys were analyzed to test the peptide sequence comparison based predictions of amino acids responsible for putative catalytic Zn binding. X-ray absorption spectrometry confirmed the presence of a protein bound Zn atom (spectral transition at ~9662 eV) with a ZnSl(N-O)3.4 coordination sphere. Induction coupled plasma emmission spectrometry measured 0.48 i 0.021 Zn atoms per wild type 2° Adh subunit. The Cys37 to Ser, HisS9 to Asn, and Asp150 to Asn mutant enzymes bound only 0.11, 0.13, and 0.33 Zn per subunit, respectively, suggesting that these residues were involved in Zn liganding. The Asp150 to Glu and Asp150 to Cys mutants retained 0.47 and 1.2 Zn atoms per subunit, suggesting that an electron rich sidechain moiety at this position preserves the bound Zn. All five mutant enzymes had 5 3% of wild type catalytic activity, indicating that the T. ethanolicus 2° Adh requires a catalytic Zn atom. Also, the Hi559 and Asp150 mutations altered 2° Adh affinity for propan-2-ol over a 140-fold range while the overall change in affinity for ethanol spanned a range of only 7-fold, supporting the importance of the metal in 2° ADH substrate binding. The lack of significant changes in cofactor affinity due to these catalytic Zn ligand mutations suggested that 2° ADH substrate and cofactor binding are structurally distinct. Altering Gly198 to Asp reduced the enzyme specific activity 2.7-fold, increased the Kmapp for NADP+ 225-fold, and decreased the Kmapp for NAD+ 3-fold, supporting the prediction that this 2° Adh binds cofactor in a common nicotinamide cofactor binding motif, a Rossmann fold. Therefore, these mutant enzyme data indicate that, unlike the liver 1° ADH, the NADP(H) linked T. ethanolicus 2° Adh binds its catalytic Zn atom using a 153 novel Cys-His-Asp motif, it does not bind a structural Zn atom, and it uses a Rossmann fold to bind nicotinarrride cofactor with G1y198 significantly responsible for its NADP(H) specificity. 154 INTRODUCTION Alcohol dehydrogenases (ADHs), central to prokaryotic and eukaryotic metabolism, are also potential biocatalysts for chiral chemical production [Keinan et al., 1986a; Keinan et al., 1986b; Hummel, 1990; Keinan et al., 1990, Bradshaw et al.; 1992a, Bradshaw et al.; 1992b]. The functionally and presumably structurally similar ADHs are classified as primary or secondary based on their higher catalytic efficiencies toward primary or secondary alcohols. These typically homodimeric or homotetrameric enzymes are almost exclusively Zn containing metalloenzymes [Branden et al., 1975] that use NAD(H)(EC1.1.1.1), NADP(I-l) (EC 1.1.1.2), or both (EC 1.1.1.71) as cofactor. Fe linked [Scopes, 1983] and ferredoxin Ego-dependent [Bleicher and Winter, 1991] 1° ADHs have also been reported The catalytic Zn's role in liver 1° ADH activity has been proposed to involve binding directly to the substrate oxygen atom, facilitating transfer of a hydride between the adjacent substrate carbon atom and the nicotinamide cofactor [Branden et al., 1975]. While 1° ADHs from diverse sources have been extensively characterized [Branden et al., 1975], little is known about the molecular basis for 1° ADH substrate specificity. Far less is is known about 2° ADH structme—function, with relatively few reported studies on enzyme purification and characterization [Lamed and Zeikus, 1981; Steinbiichel et al., 1984; Bryant et al., 1988; Ismaiel et al., 1993; Burdette and Zeikus, 1994] and a complete lack of exact structrn'al or site directed mutagenic information. Because of the 2° ADI-Is' central metabolic roles [Levitt et al., 1984; Burdette and Zeikus 1994] and their potential biotechnological value [Lamed et al., 1981], structure-function analysis of catalysis and substrate specificity would contribute significantly to the basic understanding of enzyme function and to the design of biocatalyst specificity. The exact 3-dimensional structure of the 1° ADH from horse liver has been determined [Eklund et al., 1974], indicating that the catalytic Zn atom was bound by two Cys and a His residue and that the nicotinamide cofactor was bound in a Rossman fold 155 [Branden et al., 1975; Rossmann and Argos, 1976]. Research to date has failed to identify ADH active site amino acid residues which are specifically responsible for substrate binding. Instead, substrate appears to bind the catalytic metal directly [Dunn and Hutchinson, 1973; Jacobs et al., 1974; McFarland et al., 1974]. The presumed similarity between 1° and 2° ADH structures based on shared catalytic properties provided the rationale for sequence based comparisons between these enzymes [Jendrossek et al., 1988; Peretz and Burstein, 1989; Burdette et al., 1996]. Peptide sequence alignments of thermophilic and mesophilic 1° and 2° ADHs were used to hypothesize that 2° ADHs similarly bind nicotinamide cofactor in a Rossman fold but that 2° ADHs lack a structural Zn binding loop and they use a unique Cys-His-Asp motif for catalytic metal liganding [Burdette et al., 1996]. Chemical modification experiments have implicated Zn, Cys, and His in Thermaanaerobium brackii and Thermoanaerobacter ethanolicus 39E 2° ADHs catalysis [Lamed and Zeikus, 1981; Burdette et al., 1996] but validation of the predicted Zn liganding motif has awaited assessment by biophysical analysis and site directed mutagenesis. Furthermore, the strong similarities among 2° ADH peptide sequences, kinetic parameters, subunit compositions, and molecular masses [Nagata et al., 1992; Burdette et al., 1996] suggest that the results of T. ethanolicus 2° ADH catalytic structme- function analysis will be generally applicable to other 2° ADHs. The research reported here describes the construction of a recombinant overexpression and purification system providing high yields of homogeneous enzyme. The results of site directed mutagenesis to test sequence comparison based predictions of amino acid residues important for T. ethanolicus 39E 2° ADH catalysis are also described. Predictions of catalytic Zn binding ligands and the amino acid responsible for NADP(H) cofactor specificity are assessed from site directed mutation effects on enzyme Kmapp, Vmaxapp, and Zn binding. Finally, the mutant enzyme kinetic data on 2° ADH mutants and exact structural information on 1° ADH are used to create a hypothetical working model for 156 the 2° ADH active site and to distinguish key structure-function differences between 2° and 1° ADHs. EXPERIMENTAL PROCEDURES Chemicals and reagents - All chemicals were of at least reagent/molecular biology grade. Oligonucleotide synthesis and amino acid sequence analysis were performed by the Macromolecular Structure Facility (Department of Biochemistry, Michigan State University). The kanamycin resistance GenBlock (EcoRI) DNA cartridge used in expression vector construction was purchased from Pharmacia (Uppsala, Sweden) [Oka et al., 1981]. DNA for sequencing was isolated using the Wizard Miniprep kit (Promega; Madison, WI). Media and strains - Escherichia coli (DI-15a) containing the 2° ADH recombinant plasmids were grown in rich complex medium (20 g l'1 tryptone, 10 g 1'1 yeast extract, 5 g 1-1 NaCl) at 37°C in the presence of 25 ug ml:1 kanamycin and 100 ug ml-l ampicillin. Mutagenesis - All DNA manipulations were performed using established protocols [Sambrook et al., 1989; Ausubel et al., 1993]. Point mutations were introduced into the ath gene by PCR [Ausubel et al., 1993] using the T. ethanolicus 39E 2° ADH gene clonal plasmid pADI-IB25-kan [Burdette et al., 1996] as template DNA. An oligonucleotide primer (KA4 N-end) was synthesized to bind the noncoding strand and included a Kpnl restriction enzyme site, the native ath gene ribosome assembly site, and the initiation codon for the ath gene (Table I). An oligonucleotide primer (KA4 C—end) was synthesized to bind the coding strand, it included the compliment of the ath termination codon and an Apal restriction enzyme site. Complimentary 3045 base oligonucleotide primers that contained the mutated bases were used in conjunction with two KA4 end primers to amplify the N-terminal and C-terrninal segments of the ath gene. PCR 157 Table I Oligonucleatide primers for PCR amplification of mutant ath gene DNA List of PCR primers used in the construction of 2° Adh overexpression plasmids and genes encoding mutant enzymes. The primers for the 5' and 3' end of the ath gene are listed with the mutated bases underlined and the restriction sites in italics. Altered bases in the mutation encoding primers are underlined. Mutation Primer sequence 10.4 ' 5'-end 5.CGGGGTACCCCGTA’I’I'ITAGGAGGTG’ITI‘AATGATGAAAGG3' 3'-end 5CAGTCCGGGCCCITATGC’I‘AATA'ITACAACAGG'I'I'I‘G3 lMetl . ' 5'-end 5'CGGGGTACCCCGTA'ITI'I‘AGGAGGTGTI’I‘AITAATGAAAGG3 3'-end 5CAGTCCGGGCCC'I‘I‘ATGCI‘AA’I‘A’ITACAACAGG'I'I'l‘G3 Cys37 to Ser ' . coding 5'GCI‘GTGGCCCCT'I‘QCAC'ITCGGACATI‘CATACC3 noncoding 5GGTATGAATG’I‘CCGAAGTGQAAGGGGC3 HisS9 to Asn ' ' coding S'CAroArAcrcoorAAcoAAocrorM. noncoding 5C'ITCACCI'ACAGC'I'I‘CGTIACCGAG3 Asp150 to Asn ' ' coding 5.GGAAGCI‘GCACTI‘ATGATTCCCAATATGATGACCACI‘GG3' noncoding 5GCTCCGTGAAAACCAGTGGTCATCA’I‘A'IIGGGAA’I‘CATA3 Asp150 to Cys . ' coding 5'GGAAGCTGCACITATGATI‘CCCIQTATGATGACCACI‘GG3 ' noncoding 5GCTCCGTGAAAACCAGTGGTCATCATACAGGGAATCATAAG3 Asp150 to Glu . . coding 5.GGAAGCI‘GCACI‘I‘ATGA’ITCCCGAAATGATGACCACI‘GG? noncoding 5GC'ICCGTGAAAACCAGTGGTCATCA’III‘CGGGAATCATA3 Gly 198 to Asp coding noncoding 5'CACAAACI‘GGTCI‘AC'I‘GICI‘ACGGCAATAA’ITC3' 5'GAA'ITATI‘GCCGTAGACAGTAGACCAG1T1‘GTG3' 158 syntheses of partial and complete mutated genes were performed using the Taqplus, exonuclease containing polymerase (Strategene; La Jolla, CA). All clones were expressed in pBluescriptII KS(+) with a kanamycin resistance cartridge introduced into the polylinker EcoRI site. Mutations were verified by DNA sequencing using the method of Sanger et al. ( 1977). Enzyme purification - The recombinant enzyme was purified from E. coli (DHSa) aerobically. The pelleted cells from batch cultures were resuspended (0.5 g wet wt. ml'l) in buffer A (50 mM TriszHCl [pH 8.0], 5 mM DTT, and 10 LLM ZnClz) containing 3 [lg ml'1 lysozyme. The resuspended cells were incubated on ice for 30 min, frozen in N2 (liq), thawed and centrifuged for 30 min at 15 000g. The clarified lysate was incubated at 85°C for 15 min, cooled on ice for 30 min, and then centrifuged for 30 min at 15 000g. We have recently shown that T. ethanolicus 2° Adh displayed a half-er of 1.2 days at 80°C (see chapter 3). Ammonium sulfate {50% (wt/vol)} was added to the supernatant and stirred at 4°C for 30 min. The 2° ADH was recovered from the supernatant after centrifugation for 30 min at 15 000g and precipitated with stirring at 4°C for 30 min in 70% (wt/vol) ammonium sulfate. After centrifugation for 30 min at 15 000g, the purified 2° ADH was resuspended in buffer A and stored at 4°C. Zn binding and molecular mass determination - EXAFS determination of 2° ADH Zn coordination was performed by Dr. Bob Scott (Univ. of Georgia; Athens, GA). Enzyme samples were concentrated to 1-10 mg ml'1 in 50 mM Tris:HCl (pH 8.0) and stored in N2 (liq) prior to analysis. 2° ADH samples were dialyzed against 50 mM NH4HCO3 (pH 8.0) containing 100 g 1'1 Chelex resin (BioRad; Hercules, CA) for 12 hr to remove unbound or advantitiously bound metals. The metal cleared samples were concentrated to between 1.2 and 12 mg ml'1 using centricon 30 ultrafiltration units (Amicon; Beverly, MA) for induction coupled plasma emission spectrometry (ICP) analysis (Chemical Analysis Laboratory at the University of Georgia, Athens). Subunit 159 molecular mass values were determined by comparison with standards (BioRad; Hercules, CA) using SDS-PAGE (12% polyacrylamide). Enzyme kinetics - The standard 2° ADH activity assay was defined as NADP“ reduction coupled to propan-2-ol oxidation at 60°C as previously described [Burdette and Zeikus, 1994]. The enzyme was incubated at 55°C for 15 min prior to activity determination unless otherwise indicated. Tris buffer pH was adjusted at 25°C to be pH 8.0 at 60°C (thermal correction factor = -0.031 ApH °C‘1). Assays to determine Km.pp and Vmaxm, were conducted at 60°C with substrate concentrations between 20me.pp and 0.2mew. Kinetic parameters were calculated from nonlinear best fits of the data to the Michaelis-Menten equation using Kinzyme software [Brooks, 1992] on an IBM PC. Protein concentrations were measured using the bicinchoninic acid (BCA) procedure (Pierce; Rockford, IL). RESULTS Enzyme overexpression - T. ethanolicus 2° ADH was purified from E. cali (DHSa) harboring the pADHBKA4-kan clonal construct (Table II) yielding 67% recovery of recombinant enzyme expressed at ~15% of total soluble protein without induction (Fig. I). This construct contained the native ribosome assembly site and translational initiation codons but removed the native transcriptional promoter regions, placing the gene under transcriptional control of the pBluescriptKS (+) T3 promoter. The recombinant protein population included 56 i 11% enzyme with a duplicated N-terminal methionine residue (MMKGFA...) while 44 i 11% of the protein possessed a single methionine residue at the N-terminus. Alteration of the "ATGATG" translational initiation site to "TTAATG" in the pADHB lMl-kan expression plasmid eliminated the synthesis of 2° ADH with two N- terminal methionine residues but did not alter the level of enzyme expression compared to 160 no we .3. an e: S. 5888 comma. 2 o2 d g 3a SN mm 8888.: 8o: 02 wé hm EN an 8988 :00 3% Sex. was 2.23 18 m8 .35. m8 .338 8088888 mam 2o; Seduced emooem cream memos Boer need new do w 8 588888 08.38 888085 2255 oeoh Q SQEE mean m 88H 8:2. voweuwueeo 85 8 3.838.— 83 8208 REES 80383880 .3 @0qu 88 .88 on 88 03 o. 3382 .88 2 8C 0.. mm 8 38388 83 8.88 :8 8F datum—£88 .3 .8588 e5 .832. .95 N2 5 assoc .o a. e 8e 8 3385 so; Eocene 18 w: m e5 888256 28 n .3 me 688s :8 cm 8 18 SB 83 m 08 £8 Becca—mag 33‘ em mom §ome=§8 .5 2383888 .88 0.8888 5285.85 =3-3¢~th<& 38.83 Eekeemmmxfiu $39 N 8% 3%: 282.5882 85 \e =§8R§m a 23mm. 161 that seen for pADHBKA4-kan (Fi g. I) or the enzyme specific activities toward ethanol and propan-2-ol (Table III). The Kmapp values toward propan-Z-ol, ethanol, and NADP+ were 1.1 i 0.22 mM, 53 i 9.0 mM, and 17 i 2.6 uM, for protein expressed from pADI-IBKA4-kan and 0.87 i 0.32 mM, 37 i 4.2 mM, and 8.5 i 1.4 M for enzyme expressed from pADHB lMl-kan. Catalytic Zn liganding - Chemical modification experiments have implicated Zn, Cys, and His in T. ethanolicus 2° ADH activity. Peptide alignments identified Cys37, Hi859, and Asp150 as potential catalytic Zn ligands [Burdette et al., 1996]. EXAFS analysis of the native and wild type recombinant enzyme suggested the presence of a single type of specifically bound Zn atom in the enzyme (Fig. II). The transition at 9660-9665 eV in Fig. 11a is characteristic of Zn. The Fourier transformed data (Fig. 11b) is best fit by a coordination sphere of ZnSl(N,O)3.4. The sulfur signal is consistent with that of a cysteine residue. The optimal data fit also indicated a single imidazole nitrogen ligand, leaving 2-3 potential oxygen ligands which could be from an amino acid such as Asp, water, or a hydroxide ion. Mutant ath genes encoding amino acid substitutions at residues 37, 59, and 150 were constructed by PCR (Fig. III) to test the involvement of these amino acids in catalytic Zn binding. Mutant proteins were expressed at levels similar to the wild type enzyme and were purified to homogeniety as seen on SDS-PAGE (Fig IV). Sequencing the mutated section of the ath gene confirmed the identity of each mutant. ICP analysis indicated that the enzyme expressed from the wild type gene (pADHBKA4-kan) and the single N- terminal Met mutant enzyme bound 0.47 and 0.50 Zn ions per subunit, respectively. The Cys37 to Ser and His59 to Asn mutant enzymes contained 0.11 and 0.13 Zn per subunit, respectively. Altering Asp150 to Glu preserved binding of 0.47 Zn per subunit while the Asp150 to Asn mutant bound 0.33 Zn per subunit. Changing the putative Asp150 ligand to Cys increased binding to 1.2 Zn per subunit. Finally,the Gly198 to Asp mutant enzyme also bound 0.55 Zn per subunit. Mutant and wild type enzymes treated for Zn analysis by 162 Figure 1. SDS-PAGE of wild type recombinant 2° Adhs. Lanes: 1, clone 1M1 cell extract: 2, clone 1M1 dialized ammunium sulfate precipitate; 3, clone KA4 cell extract; 4, clone KA4 post heat treatment; 5, clone KA4 dialized ammunium sulfate precipitate. The Mr values for the molecular mass markers flanking the sample lanes are indicated on the left. Proteins were stained with Coomassee brilliant blue R-250. l 2 97kDa--=::= 45 fig- 31 ‘13:! 22 s " 163 -. Il'mlll “ .\~\ 3“ 164 5:0 3:88 28 own 8: 28 8K. 8 808: 80.8 88888 88? 088883.? noon 3.. 8 n02 e: as «be: 2 mm 820 «-00: 88d 86 fiend mm Sad W006 ed end 2%: 03.: Red 8.0 woo: cm 3.0.0 088 8.0 8.0 288 no: :8. 2.0 woe: 9: 82. some 8 and :88 3.. 0886 ed 0.36 c: 86 0.00..“ v; QN Dom:— ~-0o.v Rood wed «.08? c2 8 N36: whos m6 $8 3.8 3 20.0 8 room mm a some 2 me E 888 2: 38 .ME 3 ~88 a: 28 7a“: b 18: a: :8 $8: D 38850 38850 38850 2580 ease: 35> crease ease: are; E ea58> 0835 £512 _ofiem Fetmtfiapa §> m: 88.8: 83 38850 083000 .28 8h 8: cmm 8.0383 8088:0080 3:88 8:0 m8? 888.880 803 8:88 mummy 08 .3 88:88 3888 :8 > 2: 808% 0888088 2:. .880 :0 8 088388 05 a: _0.N-:mqe.8 .83 888.88: 0.83 88:.» 8880— 3942 .8888 05 0: +32 88: 888, 88> 8:8: “8.3380 3023 88:8 DwEO 05 «a 00:0 05 8 8000.0 0908.8 tan—<2 .83 8:888“. 083 8: > 8880. .0888 0:: 3.8588 8§~ 8 0:83:80 08 83 8088:0080 888:... 88.. 88 50.8 o. .8588 meme as: .80 as ea 88 a a: 2:8 8528 B 8888 083 8.83 on08> 8: 8.08M .980 ~.28 an.» n 858$ Do 8 8 88888.: LAX—<2 2 080 00:03.83: 8 00.888 08 .3 88588 0:03 808 888 08.38“ mum—8:0 083:0 88:8 8.8 80888:: 0:08M .3880 See... N g :0 30282.35. B00 08:8 08008.. 8&0 80%.8 E 03mm. 165 Figure 2. T. ethanolicus 2° ADH EXAFS analysis. (A) Spectrograph of normalized signal intensity versus energy for native (—) and rewmbinant (- - -)2° Adhs. (B) Plot of Fourier transformed data versus distance for native (—) and recombinant (- - -)2° Adhs. uni—L 166 |_"——'r .SAdH. recombinant B—KG' 'sdbtraotad "———'_T| TADOB ‘ "-73. 8AdH.aslsolated.ADOA(08194) _B(05195 ‘ J Normalized Intensity ] J J I "' 9640 ' 9660 9680 9700 9720 Energy (eV) 167 Figure 3. PCR-based ath gene site directed mutagenesis scheme. The partial gene PCR products from reactions using overlapping coding- and noncoding-strand oligonucleotides encoding the mutation with one complimentaty to the 5' or 3' end were combined. anealed, extended, and amplified in a second PCR reaction which yielded the complete mutant (1th gene. 168 Kpn I \—‘ mut. _‘ IIWéEiPEEIéIQEIEII IIEVEQEYZ'PEEHEIEIEI‘EII vmut Apa I Amplify (26 cycles) 1 K pn I mutant N—end mutant C—end Apa I Aneal (50°C) K pn I A I + P0 KP" I Apa I lExtend (3 cycles) K pn I Complete mutant 7?)?! Amplify (26 cycles) K pn I Complete mutant Apa I 169 Figure 4. SDS-PAGE of 2° ADH mutant proteins. Lanes: 1, cell extract of E. coli DHer containing pBluescriptII-KS (+) plus the kanr cartridge; 2, C378 mutant cell extract; 3, purified C378 mutant; 4, purified D150C mutant; 5, purified DISOE mutant; 6, purified Gl98D mutant; 7, purified HS9N mutant. The Mr values for the molecular mass markers flanking the sample lanes are indicated on the left. Proteins were stained with Coomassee brilliant blue R-250. as E l .5 UI to N y. . ' n C .u 4 '. .._0 ~ e , n - ‘ . . v‘ . . . . I .9: r r r r 170 l7 1 dialysis in buffer containing metal chelating resin were active and indicated no increase in activity upon Zn addition (data not shown). Point mutations altering Cys37, His59, and Asp150 significantly reduced enzyme activity (Table III). The Cys37 to Ser mutation replaced the putative Zn ligand residue with one less than 0.4 A shorter, but the Ser hydroxyl's ionization constant is 7 orders of magnitude smaller than that of the Cys sulfhydryl. This mutant retained less than 1% of the wild type enzyme activity toward both ethanol and propan-Z-ol. Alteration of HisS9 to Asn eliminated the ionizable nitrogen atom and reduced enzyme activity to less than 0.5% of wild type. The Asp150 to Cys mutant which presumably mirrored the Cys-His—Cys motif for 1° ADH catalytic Zn binding, had the highest specific activity of the catalytic Zn mutants, retaining approximately 3% of the wild type enzyme activity. Replacement of the Asp carboxyl moiety with a sulfhydryl group preserved the strong negative ionic character but shortened the amino acid side chain by approximately 0.6 A Mutating Asp150 to Glu lengthened the side chain approximately 1.3 A and reduced the specific activity to 0.2% of wild type. Both the Asp to Cys and Asp to Glu mutations retained the electronic characteristics for metal binding but mutation of Asp150 to Asn significantly reduced the polarity of this residue, maintaining the wild type residue geometry. The Asp150 to Asn mutant retained less than 3% of the wild type activity. Cofactor and substrate specificity - All of these catalytic Zn liganding residue mutants had similarly high affinities for NADP" (Kmapp values less than 40 uM) and, like the wild type enzyme, demonstrated very low activities using NAD‘”. The Asp150 to Asn mutant Kmapp value for NADP+ was 37 “M with an unusually large standard deviation of 50%, making the difference between it and the wild type Kmapp (16 uM) insignificant. The Asp150 to Cys mutant however, had an NADP+ Kmapp value near 1.0 M which is 16-fold lower than that of the wild type enzyme. Weringa and Hoi, 1983, identified a specific amino acid position responsible for confering NAD(H) versus NADP(H) cofactor specificity in Rossman fold containing 172 proteins. Glyl98 was predicted to occupy this position in the T. ethanolicus 2° ADH sequence [Burdette et al., 1996] and is consistent with the enzyme's ability to bind NADP(H) since its small uncharged sidechain would accomodate the additional cofactor ribose phosphate, unlike the Glu or Asp residues present in N AD(H) linked enzymes that repel the negatively charged phosphate moiety. Based on Weringa and Hoi's hypothesis, mutation of Glyl98 to Asp would exclude N ADP(H) from the cofactor binding site but allow NAD(H) binding. The wild type enzyme had a l40-fold lower Kmapp for NADP+ than for NAD+ and a catalytic efficiency for propan-2-ol oxidation 2400 fold greater using NADP+ than using NAD+ (Table N). Mutation of Glyl98 to Asp increased the Kmapp for NADP+ 225-fold and reduced the Vmaxapp using NADP" 4—fold. The mutant enzyme had a 3-fold lower Kmapp toward NAD+ and a 5.5-fold higher propan-2-ol oxidation rate using NAD+ compared to wild type, giving the mutant a 6.7-fold lower catalytic efficiency for NADP“ linked propan—2-ol oxidation than for the NAD+ dependent reaction. Mutational effects on enzyme affinities for 1° versus 2° alcohol substrates and the catalytic efficiencies for their oxidation indicate a role for these putative catalytic Zn ligands in substrate specificity as well as catalytic rate enhancement. Mutation of the Cy837 to Ser increased the Kmapp 6-fold and 20-fold for propan-2-ol and ethanol, respectively (Table III). Determining the Cy837 to Ser mutant enzyme Michaelis-Menten constant for ethanol was complicated by the inability to measure enzyme initial rates using substrate concentrations greater than 700 mM and less than 300 mM. Therefore, this Kmapp is presented for completeness but it is an approximate value and is not included in subsequent data analysis. Alteration of HisS9 to Asn caused a 6-fold increase in Kmapp toward propan-2-ol and no significant change in the Kmapp toward ethanol. The Asp150 to Cys mutant Kmapp for propan-2-ol was similar to the wild type value and that for ethanol was increased only 2-fold. Mutating Asp150 to Glu caused a 3-fold increase in Kmapp for ethanol but a 29-fold increase in Kmapp toward propan-Z-ol. Finally, the Asp150 to Asn mutant enzyme had a Kmapp value toward propan-2-ol S-fold lower than the wild type and 173 25 e3 2 3.. 3 .. 93% 83 Ga v6 086 Nb +92 :«32 :5 we: a 25 was a e... IE and? IE tie - the 35850 queoD +Q85°C. Enzyme unfolding in the absence of chemical denaturant was predicted - using both protein absorbance Tm[(3u] and fluorescence measurements in the presence of GuHCl - to occur at higher temperatures (~115°C) than precipitation. Thermally precipitated protein did not redissolve in buffer nor did it display catalytic activity, indicating that thermoinactivation must precede or accompany precipitation. The pseudo-first order thermoinactivation rate determining step argued that this step was unimolecular with respect to protein and that it preceded multimolecular protein 191 precipitation. Scheme 1 summarizes the proposed thermophilic 2° ADH thermoinactivation process. . slow . . . nanve Q non-nauve _) Precrprtate (1) (active) step (inactive ?) (inactive) Propan-2-ol oxidation by enzyme preincubated at 55°C showed no statistically significant Arrhenius discontinuity. This temperature dependence of enzyme activity, consistent with Arrhenius theory, is completely described by the change in average substrate kinetic energy. Therefore, the discontinuous 2° ADH plots appear to result from a slow enzyme structural change flour a highly active form to a less active form and that thermozyme Arrhenius plots discontinuities are not necessarily evidence of the role of reduced low temperature flexibility in the poor low temperature activity of thermozymes. The observed 2° ADH catalytic rate enhancement at low chemical denaturant concentrations required slightly greater GuHCl concentrations at higher temperatures contrary to the prediction of a catalytically significant effect of GuHCl on enzyme flexibility. The similar low effector concentration, enzyme rate enhancement was seen with added KCl, which is not a strong denaturant, indicated that the rate enhancement resulted from ionic interactions. These data argue that the observation of low temperature enzymatic rate enhancement by added chemical denaturants also may not indicate a role for optimal flexibility in the temperature dependence of thermozyme activity. The possible correlation between activity and flexibility therefore, requires direct measurements of both enzyme properties. 192 Chapter 6 Directions for Future Research 193 SEQUENCE BASED MUTAGENESIS AND KINETICS The HCA alignment between thermophilic and mesophilic 1° and 2° ADHs (see chapter 2) predicted significant structural similarity despite dissimilar peptide sequences. These predicted catalytic structural similarities were supported by the results of mutagenic and biophysical analyses (see chapter 4). The predictive power of comparative sequence analysis was proposed as a tool to identify the molecular determinants of ADH substrate specificity. Furthermore, developing mutagenesis and active recombinant enzyme expression systems provide the means to test these structure function predictions. An extensive critical core residue comparison between the horse liver and B. stearothermophilus 1° ADHs and the T. ethanolicus and C. beijerinckii 2° ADHs indicated strong amino acid characteristic conservation among the peptides (see appendix A; tablel). The conspicuous sequence differences between the 1° and 2° ADHs are listed in Table 1. These residues are mostly clustered around the horse liver enzyme active site based on the crystal structure (Fig. 1). Determining the effects of 1° versus 2° ADH comparison based T. ethanolicus ath gene mutations on the 2° ADH substrate specificity and the effects complimentary mutations have on 1° ADH specificity will advance our understanding of the molecular biochemistry of ADH molecular biochemistry. This information may provide site direcwd mutagenesis based ADH specificity optimization for industrially relevant biotransformations. The higher stability of NAD+ versus NADP+ makes it preferable for indusrtial applications. A single Gly to Asp mutation shifted the T. ethanolicus 2° ADH catalytic efficiency ratio from 2400-fold in favor of NADP+ linked activity to 6.7-fold in favor of the N AD+ dependent reaction. This mutation's effectiveness was strong evidence supporting the predicted 2° ADH Rossmann fold nicotinamide cofactor binding motif. The mutation reduced the maximal NADP+ linked enzyme velocity 3-fold and increased the NAD+ dependent reaction 5-fold creating a faster catalyst using NAD+ but a generally 194 Table 1. Comparison of conserved 2° ADH amino acids with the corresponding 1° ADH residues deduced from HCA based sequence alignments P’ ADI-ls 7° ADI-ls Horse liver B. stearotlierm. C. beijerinckii T. ethanolicus AromanE-nonammafic Va113 Va15 Trp15 Trp14 Val83 Leu77 Phe7 5 Phe75 Leu350 Val310 Phe328 Phe328 Phe93 Phe89 Thr86 lle86 conformational Met40 Ile32 Pro31 Pro31 lle45 Val47 Pro36 Pro36 Charge Thr59 Pr053 Asp51 Glu51 Leul71 Prol45 Met147 Met147 Phel76 Va1151 Met152 Met152 Size Alall Ala3 Leu13 Va112 Ala12 Ala4 Gly14 Gly13 Leu14 Val6 lle16 Ile15 Va136 Va128 Ala27 Ala27 Va158 Va151 Gly50 Gly50 V3163 lle57 Met55 Met55 lle90 Ile85 Val83 Va183 Ala183 Ala155 Gly158 Gly158 Va1186 Va1158 Leu161 Leu161 Va1189 Ala161 Ile164 lle164 Leu342 Ala300 Va1314 Va1314 Ile346 Ile307 Val320 Val320 Glu353 Glu314 Asp328 Asp328 Gly35 8 Va13 19 Ala333 Ala333 Arg368 Arg331 Lys346 Lys346 195 Figure 1. Position of the non-conservative amino acid differences between 1° and 2° ADI-Is relative to the horse liver 1° ADH catalytic site. The liver ADH structural coordinates were obtained from the Cambridge protein structure database (filename "pdbadh6.ent"). Residues identified from the catalytic domain (red) and in the cofactor binding domain (green) are shown in relation to the catalytic Zn atom (light blue), the His ligand to the catalytic Zn (dark blue), and the two catalytic Zn cys residues (yellow). I . ~t I t , t .i _\.‘\J ._l , v r t b - ~— .. t J t.» 3. , V cl} .5“: A _'.l odd 197 lower turnover enzyme. Also mutant enzyme affinity for NAD+ increased only 3-fold while decreasing the 2° ADH affinity for NADP+ 225-fold. This Glyl98 to Asp mutant protein therefore, represents an incomplete transformation of nicotinamide cofactor preferences. The reduction potentials for NAD+ and NADP+ are identical, NAD+ and NADP+ structures differ only by a single phosphate group, the Rossmann fold structures that bind the cofactor forrrrs differ only slightly, and NAD+ linked 2° ADH with high turnover and high cofactor affinity exist in nature. These observations and the partial success already achieved argue that this N ADP+ linked 2° ADH may be successfully converted by further amino acid replacement into an improved functional NAD“ dependent enzyme with high cofactor affinity. Sequence based comparative analysis of NAD+ and NADP+ linked ADHs (Appendix A; table 2) indicates nine further mutations that may complete this conversion (T able 2). These amino acid positions were identified based on their being consistently different between the NAD(H) linked and the NADP(H) dependent enzymes. The residue positions in the liver 3-dimensional structure (Fig. 1) indicate that they are also clustered near the catalyitc Zn atom. Assessing the porential contributions of these mutations on NAD(H) versus NADP(H) specificity requires additional kinetic analysis of the mutated enzymes. The greater than 85% similarity between the T. ethanolicus and C. beijert'nckt‘i 2° ADH peptide sequences plus the vast difference in their reported thermostabilities makes the reciprocal site directed mutation of these proteins based on comparative sequence analysis a practicable and rational approach to understanding enzyme thermostability and thermOphilicity. Of the 89 amino acid differences between these peptide sequences, only 12 were classified as nonconservative. Of these 12, nine were substituted Pro residues in the thermophilic T. ethanolicus enzyme (Table 3). Substituting Pro residues into protein loop regions have been shown to thermostabilize engineered proteins and to be important to natural protein thermostability (see chapter 1). A Proline-zipper hypothesis has been proposed to explain the enthalpic and entropic contributions of Pro residues to folded 198 83: emcee 82m Gene acne 8988c. . 2 £5. ”E: 8e: 93. 229285 83o 8nd 824 use»; use 8&2 8&2 835 Read 8883 3&5 330 Race. 2.8:. 39:. ENE. fined 32d 82:. 82:. 3:3 «fie: IIES Ea: Ewe meefi «8 use? SeagéhF Eggnog U .Euefieeae .m b>m 88m Sufism 8 monument 38 can.“ moewma 3mm “Sex—AMP 855:2? 3:253 e82. <0: :8...— 3933 3:28.. :3. a 285%.. +3.: ”5888.98 2.. 5? 88 2:5: :3 en ace—.53.. tun—<2 eaten-.8 he commune—Sew .u 03.; 199 was: named cow. 28:2. 383% 8835 833 bad 2%: «Nae: 8&8 escapes .eaee sen: REF. 2.3.5 £28 :a .«e EB e: - .32: 38880 a 38 53880“ as“: eN eggs on: 3. a8 83:. 9.3: See £2982 Sauteed: new? See: cone £2852 382.2522 2:9 29: 2me mac: cone peace: eeeeaseeeea «New ac: Aeaaficfiaavemao ace ~92 Nae: 8:2 cease: 2.829554 eaaee aged finerebom .D 3.58:5» H r in? em I cores.“ gene: :5 Sentence .5me $38.9 $555.3 3:253 e83 <0: See e355: :3 en ”em—Enema:— e... 2 8.2.58 8:28.. 8:8.— ::< on 3.28.52: 82:. .n 2.2:. 200 thermozyme [1] stability. Importantly, in this thermophilic-mesophilic 2° ADH system, each Pro residue's contribution to enzyme thermal properties can be assessed by both the destabilizing effects due to thermophilic enzyme Pro removal and the stabilizing effects due to mesophilic 2° ADH Pro addition. This type of extensive reciprocal study using a set of multidomain, multisubunit enzymes has not been reported. Furthermore, the entensive sequence similarity between the mesophilic and thermophilic 2° ADHs reduces the possible number of thermal property determinants, allowing for potentially complete characterization of the structural determinants of these properties. The PCR based mutagenesis system developed here would also allow rapid construction of multiple mutations to assess the independence or cooperativity of these mutational effects. Therefore this system appears to be an excellent model to test the proline zipper hypothesis through biophysical measurement of protein thermal unfolding and thermoinactivation (eg., temperature or chemical denaturant dependent circular dichroism. differential scanning calorimetry, temperature or chemical denaturant dependent absorbance spectrophotometry or fluorimetry) (see chapter 3). i The thermophilic-mesophilic enzyme reciprocal mutagenesis experiments assess the contribution of each Pro to corresponding mesophilic peptide re$idue mutation to the overall folded protein stability. The specific thermophilicity and thermostability contributions of the added Pro residues as opposed to the replacement of the other amino acid must be quantitated by constructing a series of mesophilic and thermophilic protein mutants with a specific neutral (eg., Ser or Ala) residue at all Pro substitution positions and measuring their thermal properties (see chapter 3). The effect of each mutation on the enzyme thermal properties by itself, the effect due to interaction of the substituted residue with the rest of the protein, and the contribution of the individual mutation to the total difference in the thermal properties between the two 2° ADI-1s should be measured to identify potential mechanisms to account for the mutational effect. The difference between the completely neutral residue substituted thermophilic enzyme (thermophilic baseline 201 mutant) and the completely neutral residue substituted mesophilic enzyme (mesophilic baseline mutant) thermal properties reflects the contribution of other peptide factors in determining these thermal stability and activity properties. Next, subtracting the thermophilicity and thermostability values of each baseline mutant from the conesponding completely Pro substituted enzyme will quantitate the total stabilizing effect of these ' prolines on both 2° ADH folds. The difference between the mesophilic and thermophilic total stabilizing effects measures the interaction of these Pro mutations with the rest of the folded enzyme structure. The mutation specific and protein interactive effects of each Pro could be examined in this way. A large difference would indicate that the proline effects are strongly dependent on overall protein structure whereas a small difference would indicate independent thermal property contributions by these residues. The latter result would argue that Pro loop mutations are a predictable general thermostabilizing strategy. The neutral residue used in this study should mimic either general amino acid or Pro characteristics (aside from its side chain structural constraint). I propose that either Ala or Ser be used because they are small, reducing the possibility of disrupted local packing, they are uncharged but are often found at the protein surface, reducing the possibility of altering local 3-dimensional structure or salt bridging, and unlike Gly, both have residue geometric constraints typical of general amino acids. Furthermore, Ala's more Pro-er aliphatic sidechain makes it a closer analog and, since the substitutions are proposed in surface loops, the hypothesized tat-helix stabilizing Ala characteristic should not be relevant. Loop stabilization has also been proposed to enhance enzyme thermostability (see chapter 1). Determining the thermostability and thermophilicity effects of complimentary mutations to swap loop regions between thethermophilic and mesophilic 2° ADHs could specifically test this hypothesis. By identifying putative core regions the HCA representation also proposes the putative loop positions (those stretches of hydrophilic amino acids between the hydrophobic clusters). Like the Pro study, this work would evaluate both the role of this strategy in native enzyme stability and the protein engineering 202 potential for loop stabilization. Furthermore, the relative effects of Pro insertion into 100ps and loop stabilization strategies that do not involve prolines (i.e., added salt bridges) could be examined using this system. Substituting stabilized loops which do not contain Pro into both mesophilic and thermophilic 2° ADH loops that were shown to effect enzyme thermal properties due to the presence or absence of Pro residues would directly compare the relative efficacy Pro versus non-proline loop stabilization strategies. STRUCTURAL ANALYSES The research described in chapters 1 and 2 indicates that the poor low temperature activity of the thermophilic 2° ADH can be explained by the Arrhenius theory. Therefore the temperature dependence of enzyme activity can be described by changes in the average substrate kinetic energy alone. The potential influence of protein flexibility temperature dependence on the temperature dependence of enzyme activity however, has not been explicitly determined. Thermal property altering mutants of both the T. ethanolicus and the C. beijerincla'i enzymes (eg., the Pro residue or loop substituted mutants described in the previous section) would provide an excellent system to examine this relationship using electron paramagnetic resonance spectrometry (EPR). The flexibility of immobilized liver 1° ADH in solvents has been determined from the spin label environment dependent EPR relaxation correlation times using a spin labelled compound coupled to Cys residues [1]. Enzyme flexibility has also been reported as the extent of deuterium exchange between the D20 solvent and the protein (see chapter 1). Correlating the temperature dependence of flexibility and activity for such a system of mutants would test the hypothesis that enzymes require optimal flexibility for optimal activity. Furthermore, direct analysis of the relationship between protein flexibility and activity would examine whether the reported similar mesophilic and thermophilic protein flexibilities at their respective maximal activity temperatm'es are necessary for maximal activity or simply an indication that the folded structures are destabilizing. 203 2° ADH catalytic Zn binding was pr0posed to involve Cys, His, and Asp based on HCA ADH sequence comparisons, and chemical modification experiments (see chapter 2). The loss of catalytic activity and the predicted changes in protein Zn content for the mutant enzymes supported the hypothesis that Cys37, HisS9, and Asp150 provided the catalytic Zn ligands (see chapter 4). Also consistent with this hypothesis, EXAFS data identified a bound Zn atom with a coordination sphere containing 1 Cys sulfur, 1 imidazole nitrogen, and 3-4 other nitrogen or oxygen ligands. Similar EXAFS analysis of the Zn ligand mutant enzymes will provide relative signal intensities for the Zn coordinated atoms. Comparing the expected ligand specific EXAFS signals for each mutant to those observed will determine if the mutated residue directly altered the metal coordination sphere. These data would provide strong evidence to support or to refute the direct role of Cys37, HisS9, and Asp150 in catalytic Zn binding. This work is currently underway in our laboratory. Constructing a recombinant T. ethanolicus 2° ADH overexpression system provided sufficient quantities of purified enzyme for crystallographic analysis. The exact 3- dimensional structure of this thermophilic 2° ADH is currently being determined using x- ray crystallography by our collaborator Dr. Bobby Ami. These data, the first 2° ADH exact 3-dimensional structure, would provide final confirmation of the overall 1° versus 2° ADH structural similarities. Comparing the liver 1° ADH and T. ethanolicus 2° ADH exact structures would also determine the validity of predictions based on the peptide sequence alignments and mutagenic analysis of catalytic Zn and cofactor binding. An exact 2° ADH structure would allow computer modeling to analyze the potential effects of proposed mutations and allow the identification of subtle mutations that may be structurally or functionally significant (ie., those based on the residue's position in the folded protein which do not require significantly altered amino acid properties which cannot be identified from the peptide sequence alone). Specifically, armed with both a 1° and a 2° ADH crystal structme residues influencing enzyme catalytic site geometry and altering substrate selectivity may be identified, allowing detailed analysis of the ADH substrate binding and 204 reaction mechanisms. Furthermore, the regions of subunit interaction would be identified, determining if this tetrameric enzyme is formed by the association of two 1° ADH-like dimers as currently believed. The number of su'ucturally competent active sites in the holoprotein would also be determined. Crystallization of the 2° ADH with acetleoA will also determine whether this physiological substrate is completely contained in the protein during catalysis or whether, as we propose, the long coenzyme A portion of the molecule is outside the protien during thioester reduction. If our hypothesis is correct, then the 2° ADH may catalyze the reduction of a wide range of very large ketone substrates but if the coenzyme is eopletely retained in the protein active site then the 2° ADH substrate range would be expected to have a more limited size range. The crystallographic [3 values, the uncertainties in the atomic positions, for protein regions could be compared in the wild type enzyme x-ray structure and in thermal property altering mutant protein x-ray structures. Significantly altered regional [3 values would point to specific regions potentially responsible for folded enzyme structural integrity. Temperature dependent crystal structural analysis could directly measure differences in regional flexibilities within the protein (Petsco, personal communication). Comparing the measured flexibilities (residue Bt-values) and the corresponding enzyme activities could directly address the role of protein flexibility in enzyme activity. BIOTECHNOLOGICAL UTILITY Changing the cofactor specificity of the T. ethanolicus 2° ADH (eg., conversion from an NADP" to an NAD+ linked enzyme), aside from yielding basic scientific protein structure function information, would construct a more valuable biocatalyst due to more desirable cofactor properties (eg., the greater thermal stability of NAD"). Om' collaborator, Dr. Robert Phillips, is examining the utility of this thermophilic 2° ADH in transforming ketones into chiral alcohols. Expanding the range of identified chiral reactions this 2° ADH performs and the effect of reaction conditions (eg., solvents and temperatm'e) on both _— 2.---..2.-- --- .....-. ..-_.. 205 activity and enantiospecificity also increases the likelyhood that it will be incorporated into a biotechnological production process. For very large substrates, solubility and the need for aggitation may be avoided by reversibly binding the substrate to a solid matrix, exposing it to enzyme in the presence of cofactor (similar to a chromatographic process), and releasing the product. This type of system would interface with the current diversimer organic synthesis technology used in drug development as the immobilized particulate enzyme could be separated from large or small scale batch reactants by centrifugation and it could be retained in a column for flow reacter systems. Furthermore, research identifying the exact molecular determinants of T. ethwwlicus 2° ADH thermal and catalytic properties would allow tailoring the properties of this stable chiral oxidoreductase. The ability to modify this enzyme to perfom specific reactions under specific conditions identified by industry would make this a strong candidate for chiral industrial processes. REFERENCES 1 Vieille, c., Burdette, D. s. and Zeikus,.l. G. (1996) in Biotech. Ann. Rev., vol. 2, (M. R. E1 Geweley, ed.) pp. 1-83. Elsevier, Amsterdam, Neth. 2 Guinn, R. M., Skerker, P. S., Kavanaugh, P. and D. S. Clark (1991) Biotechnol. Bioeng. 37, 303-308 APPENDIX A 206 Appendix A Sequence Comparison Based Predictions of Important Catalytic Domain and Cofactor Binding Domain Amino Acids in 1° and 2° ADH Structures Because of the relatively small number of exact 3-dimensional protein structures determined and the typically well conserved structural folds within each enzyme family, peptide sequence based comparisons are used to predict structurally analogous regions of these proteins. ADHs characteristically share significant 3~dimensional structural architecture but very little amino acid sequence similarity [l]. The horse liver 1° ADH x-ray structure has been determined and its amino acid structure-function characteristics extensively investigated [1]. HCA comparison of the thermophilic T. ethanolicus 39E 2° ADH, the thermophilic B. stearothermophilus 1° ADH, and the mesophilic C. beijerinckii and A. eutrophus 2° ADH peptides sequences to that of liver l° ADH (Fig. I) predicted a Rossmann fold for nicotinamide cofactor binding and specific amino acids associated with cofactor specificity and catalytic Zn liganding. The complete sequence alignments also predict corresponding residues to those that are part of identified liver ADH structural elements. Similarity among these structurally important residues, while not practically testable by point mutations, provides evidence in support of the folded-architectural similarity between the compared peptides. Tables 1 and 2 list the T. ethanolicus 2° ADH, C. biejerinckii 2° ADH, and B. stearothermophilus 1° ADH residues predicted to correspond to important residues identified in the horse liver enzyme. These tables also indicate the percent amino acid similarity and identity between these four peptides for specific sections of the horse liver ADH catalytic and cofactor binding structural domains. Figure 1. HCA comparison of the B. stearothermophilus 1° ADH, the T. ethanolicus 2° ADH, and the C. beiien’nckil 2° ADH amino Acid sequences to that of the horse liver 1° ADH. The proposed catalytic Zn liganding (0). structure Zn liganding CD, and Rossmann fold bindingm) residues are indicated. Symbols :‘k‘, Pr01ine;°.slycinc;'i. scrinc;'n. threonine I. W393 2° Adh e horse liver 1° Adh 209 Table 1. Comparison of aligned amino acids involved in ADH catalytic domain structure and function _Proposed role in Aligned peptide residues _ liver ADH Horse liver B. stearotherm. C. beijerinckii T. ethanolicus structure Active domain Alall A133 Leu13 Va112 core residues Alal2 A134 Glyl4 Gly13 Va113 V315 TrplS Trp14 Leu14 V316 Ilel6 Ile15 Phe21 Leul3 n.d. n.d. Va126 N.D. V3121 Ala21 V3136 V3128 Ala27 A1327 Ile38 V3130 V3129 V3129 Met40 Ile32 Pr031 Pr031 Thr43 Cys45 V3134 Trp34 Ile45 Val47 Pro36 Pro36 Thr59 ProS 3 Asp51 Glu51 Val63 Ile57 Met55 Met55 lle64 Ile58 Ile5 6 Ile56 A1365 ProS 9 Leu57 Leu57 Ala69 Gly63 Ala61 Ala61 A1370 Val64 Val62 Val62 Val73 Ile67 ' V3166 V3166 V3180 V3174 V3172 V3172 V3183 Leu77 Phe75 Phe75 V3189 V3183 Ile82 V3182 Ile90 lle85 Val83 Val83 Pro91 Pr086 Pr084 Pr084 Vall69 Alal43 Alal45 Alal45 Leul7l Prol45 Met147 Met147 Ilel72 Ile146 Ile148 Ile148 Glyl73 Phe147 Thr149 Prol49 Phel76 VallSl Met152 Met152 Ala183 Ala155 Gly158 Gly158 V31186 V31158 Leul6l Leu161 A13187 Thr159 Alal62 Ala162 Va1189 Alal6l Ile164 Ile164 Amino acid similarity (%) Horse liver 100(100) 38(84) 25(80) 25(80) B. stearotherm. 100(100) 2202) 2505) C. beijen’nckii 100( 100) 78(94) T. ethanolicus 100(100) Active site Ser48 Thr40 Ser39 Ser39 pocket residues Leu57 ProSO Leu49 Ile49 Va158 V3151 Gly50 Gly50 Phe93 Phe89 Thr86 Ile86 , Phel 10 LeulOS Leu93 Ser93 Leu116 Tyr114 Phe99 T Thrl78 Thr152 Ser154 Thr154 210 Amino acid similarity (%) Horse liver 100(100) 43(86) 28(78) 28(78) B. stearotherm. 100(100) 14(86) 2801) C. beijerinckii 100(100) 7 8(100) T. ethanolicus 100(100) Catalytic Zn Cys46 Cys38 Cys37 Cys37 ligands His67 His62 HisS9 HisS9 Cysl74 Cysl48 Asp150 Asp150 Amino acid similarity (%) Horse liver 100(100) 100(100) 67(67) 67(67) B. stearotherm. 100(100) 67(67) 67(67) C. beijerinckr’i 100(100) 100(100) T. ethanolicus 100(100) Structural Zn Cys97 Cys92 Cys85 n.d. liganding CyleO Cys95 n.d. n.d. residues Cys103 Cys98 n.d. n.d. Cyslll Cyle6 n.d. n.d. Amino acid similarity (%) Horse liver 100(100) 100(100) 25(25) mm B. stearotherm. 100(100) 25 (25) ------ C. beijerinckii 100(100) ------ T. ethanolicus ------ Structural Gly Gly66 Gly60 Gly60 Gly60 residues Gly77 Gly71 Gly69 Gly69 Gly86 Gly8O Gly78 Gly78 Amino acid similarity (%) Horse liver 100(100) 100(100) 100(100) 100(100) B. stearotherm. 100(100) 100(100) 100(100) C. beijerinckii 100(100) 100(100) T. ethanolicus 100(100) Overall amino acid similarity (%) , Horse lrver 100( 100) 5 1 (88) 32(76) 3004) B. stearotherm. 100(100) 29(25) 30(69) C. beijerinckii 100(100) 79(94) T. ethanolicus 100(100) n.d.: no analogous residue was identified. 211 Table 2. Comaprison of aligned amino acids involved in ADH nicotinamide cofactor binding Proposed role in Aligned peptide residues liver ADH ADP- ribose 4 bindin Horse liver B. stearotherm. C.7§eijerinckii T. Zhanaicus Core residues Aral 83 Thr153 A13159 Ala159 V31186 Tyr154 Leul61 Leul61 A13187 Ala156 Alal62 Ala162 Va1189 Va1159 Ile164 Ile164 Cys195 V31168 Vall70 Vall70 V31197 Ilel70 Vall72 Vall72 Leu200 Ile203 Ilel75 Ilel75 Val203 Leu206 V31178 V31178 A13237 A13209 Ala212 Ala212 Phe264 Val23l aArg23g A13238 Va1268 V31235 Ala242 Ala242 V31288 A13256 Ser263 Ala263 V31290 V31258 Ile265 Va1265 V31292 V3126O Tyr267 Tyr267 Ala317 Ala295 Gly239 Gly293 Alal96 Alal69 Vall7l Ala17l Phel98 Tyrl7l Ile173 Leu173 Ile220 Vall92 Ilel95 Ile195 Va1222 Vall94 V31197 Val97 lle250 I-Ii5217 Ile263 Ile263 Leu254 A13221 V31227 I1e227 V31262 Gly229 V31236 V31236 Ala278 Ser250 Met255 Ile255 Cy5281 Arg252 Gly259 Gly259 CysZ 82 Arg253 Gly260 Gly260 Amino acid similarity (%) Horse liver 100( 100) 36(68) 36(68) 4406) B. stearotherm. 100(100) 20(60) 32(68) C. beijerinckii 100(100) 76(88) T.’ ethanolicus 100(100) Adenine pocket Phel98 Tyrl7l Ile173 Leu173 residues V31222 Vall94 V31197 V31197 Ile224 Leul96 Thrl99 Thrl99 Pro243 Pr0215 Tyr218 Tyr218 Ile250 His217 lle223 Ile223 lle269 V31236 Gly243 Gly243 Thr274 Thr237 Gly244 Gly244 Arg27l A13238 Ser246 A13246 Adenosine Asp223 Aspl95 Glyl98 Glyl98 ribose binding Glyl99 Gly172 Glyl74 Glyl74 residues Ile269 V31236 Gly243 Gly243 212 Asn225 Glyl97 Arg200 Arg200 Ly8228 LysZOO CysZO3 Cy3203 Pyrophosphate Arg47 His39 Thr38 Thr38 interacting Ile269 V31236 Gly243 Gly243 Nicotinamide Gly293 Gly26l Hi3268 Phe268 ribose biding bAmino acid similarity (%) Horse liver 100(100) 5008) 21(28) 21(28) B. stearotherm. 100(100) 14(28) 21(36) C. beijerinckii 100(100) 78(86) T. ethanolicus 100(100) Overall amino acid similarity (%) Horse liver 100(100) 4102) 30(52) 35(57) B. stearotherm. 100(100) 18(48) 28(56) C. beijerinckii 100(100) 77(87) T. ethanolicus 100(100) 3Possibly a sequencing error. bIle269 is only counted once in the calculation. 213 REFERENCES Branden, C.-I., Jornvall, H., Eklund, H., and B. Fururgren (1975) Alcohol dehydrogenases. In The enzymes, 3rd ed., vol. XI part A" p. 105-106, (P. D. Boyer, ed.). Academic Press, New York. APPENDIX B 214 Appendix B Computer programs used to calculate (hyper)thermophilic versus mesophilic total amino acid composition and the theoretical Arrhenius data for thermophilic and mesophilic enzymes Computer program to compile amino acid composition data from peptide sequences and to predict percent enzyme activity values at temperatures below the maximal temperature for mesophilic or thermophilic enzyme activity were written using Turbo Pascal 3.0 for IBM (Borland International; Scotts Valley, CA). Programs were compiled and executed using an IBM PC. Program "TOTALAA" calculates the percent amino acid composition of a peptide sequence input as a suing of letters in an ascii format text file, and it calculates the average amino acid composition of a group of peptide sequence files input in succession (used to calculate the values reported in chapter 1; Table 3). The output is written to the default line printer and to the monitor. Program "Arrheniussimulation" calculates the expected precent catalytic activities within a temperature range below the temperature for optimal activity (Topt) specified by the low temperature (Told) in Kelvin (see chapter 1; Fig. l). The calculation can be performed on enzymes with any reaction activation energy (Ea) in kJ/mol. The output data is written to an ascii format text file named "ARRSIMDA ". 215 program TOTALAA; VAR COMPARE: ARRAY [1..20] OF RECORD LABELC: STRING[25]; NGC: REAL; NPC: REAL; NAC: REAL; NV C: REAL; NLC: REAL; NIC: REAL; NFC: REAL; NWC: REAL; NYC: REAL; NHC: REAL; NCC: REAL; NMC: REAL; NSC: REAL; NTC: REAL; NKC: REAL; NRC: REAL; NDC: REAL; NEC: REAL; NN C: REAL; NQC: REAL; END; FILEIN: TEXT; NAMEIN,CATEGORY: STRING[25]; 216 AA,PAUSE,YNT: STRING[l]; i,J,NG,NP,NA,NV,NL,NI,NF,NW,NY,NH,NC,NM,NS ,NT,NK,NR,ND,NE,NN,NQ ,TOT: INTEGER; N GT,NPT ,NAT,NVT,NLT,NIT,NFT,NWT,N YT,NHT,N CT ,NMT,NST,NT’I’,NKT,N RT,NDT,NET,NNT,N QT: REAL; SQNGT,S QN PT ,8 QNAT,SQNVT,SQNLT,SQNIT,SQNFT ,SQNWT ,SQNYT,SQNHT, SQNCT,SQNMT,SQNST,SQNTT: REAL; SQNKT,SQNRT,SQNDT,SQNET,SQNNT,SQNQT: REAL; MGT,MPTMAT,MVT,MLT,MIT,NH71‘,MWT,MYT,MHT,MC1‘,MMT,MST,MTI‘,MKT ,MRT,MDT,MET,MNT,MQT: REAL; GTSD,PTSD,ATSD,VTSD,LTSD,ITSD,FI‘SD,WTSD,YTSD,HTSD,CTSD,MTSD,STS D,T'I‘SD,KTSD,RTSD,DTSD,ETSD,NTSD,QTSD: REAL; TOTR: REAL; YN,X: BOOLEAN; PROCEDURE YESNO; VAR CK: BOOLEAN; BEGIN REPEAT IF YNT = 'Y' THEN BEGIN YN := TRUE; CK := TRUE; END ELSE IF YNT = 'y' THEN BEGIN 217 YN := TRUE; CK := TRUE; END ELSE IF YNT = 'N' THEN BEGIN YN := FALSE; CK := TRUE; END ELSE IF YNT = 'n' THEN BEGIN YN := FALSE; CK := TRUE; END ELSE BEGIN CK := FALSE; WRITELN; WRITELNCPLEASE ENTER EITHER Y OR N'); READ(YN'I); WRITELN; FUNCTION RD(A:REAL): CHAR; VAR S,SO: STRING[18]; 218 i,M,P: INTEGER; Y,YY: CHAR; BEGIN STR(A,S); FOR i:= 1 TO 18 DO BEGIN IF S[i] = '.' THEN BEGIN M := i + 3; P := i; END; IF S[i] = 'E' THEN BEGIN IF S[i+3] = '1' THEN BEGIN Y := S[P+1]; S[P] := Y; S[P+1] :='.'; END; IF S[i+3] = '3’ THEN BEGIN Y := S[P-l]; S[P-l] := ‘.'; S[P] := Y; END; IF S[i+3] = '4' THEN 2 l9 BEGIN FORi:=1T018 DO BEGIN SO[i] := S[i]; END; Y := SO[3]; S[3] := '.'; S[4] := '0'; S[S] := SO[3]; S[6] := SO[S]; S[7] == SO[6]; S[8] := SO[7]; M := 8; END; END; END; FOR i := 1 TO M DO BEGDI WRITE(S[i]); END; RD := ' '; END; FUNCTION PRD(A:REAL): CHAR; VAR S: STRING[18]; i.M.P: INTEGER; ll" 220 Y: CHAR; BEGIN STR(A,S); FOR i:= 1 TO 18 DO BEGIN IF S[i] = '.' THEN BEGIN M := i + 3; P := i; END; IF S[i] = 'E' THEN BEGIN IF S[i+3] = '1' THEN BEGIN Y := S[P+l]; S[P] := Y; S[P+1] := '.'; END; END; END; FOR i := 1 TO M DO BEGIN WRITE(LST,S[i]); END; PRD := ' '; END; 221 BEGIN WRITELNCINPUT THE LABEL FOR THIS SET OF FILES (25 CHAR OR LESS)'); READ(CA'IEGORY); WRITELN; REPEAT WRITELNCENTER INPUT FILE N AME'); READLNCNAMEIN); ASSIGN(FILEIN,NAMEIN); RESET(FILEIN); CLRSCR; i??? $29.2»; ll 9 érégiéai gee; II S? 222 II II 99 2%??? C? 9 NO := 0; TOT := O; WHILE NOT EOF(FILEIN) DO BEGIN READ(F1LEIN.AA); IF UPCASE(AA)='A' THEN BEGIN NA := NA + 1; END ELSE IF UPCASE(AA) = 'G' THEN BEGIN NG := NG + 1; END ELSE IF UPCASE(AA) = 'P' THEN BEGIN NP := NP + 1; END ELSE IF UPCASE(AA) = 'V' THEN BEGIN NV := NV + 1; END ELSE IF UPCASE(AA) = 'L' THEN 223 BEGIN NL := NL + 1; END ELSE IF UPCASE(AA) = '1' THEN BEGIN N1 := N1 + 1; END ELSE IF UPCASE(AA) = 'F' THEN BEGIN NF := NF + 1; END ELSE IF UPCASE(AA) = 'W' THEN BEGIN NW := NW + 1; END ELSE IF UPCASE(AA) = 'Y' THEN BEGIN NY := NY + 1; END ELSE IF UPCASE(AA) = 'H' THEN BEGIN NH := NH + 1; END ELSE IF UPCASE(AA) = 'C' THEN BEGIN NC := NC + 1; END 224 ELSE IF UPCASE(AA) = ’M' THEN BEGIN NM := NM + 1; END ELSE IF UPCASE(AA) = '8' THEN BEGIN NS := NS + 1; END ELSE IF UPCASE(AA) = '1" THEN BEGIN NT := NT + 1; END ELSE IF UPCASE(AA) = 'K' THEN BEGIN NK := NK + 1; END ELSE IF UPCASE(AA) = 'R' THEN BEGIN NR := NR + 1; END ELSE IF UPCASE(AA) = 'D' THEN BEGIN ND := ND + 1; END ELSE IF UPCASE(AA) = 'E' THEN BEGIN NE := NE + l; 225 END ELSE IF UPCASE(AA) = 'N' THEN BEGIN NN := NN + 1; END ELSE IF UPCASE(AA) = 'Q' THEN BEGIN NQ := NQ + 1; END; TOT := TOT + l; TOTR := TOT; WRITE(AA); END; WRITELN; CIDSE(FH—EIN); WITH COMPARE[i] DO BEGIN LABELC := NAMEIN; NGC:: N GIT OTR; NPC:: NP/TOT'R; NAC:: NA/I‘OTR; NV C:= NV/l'OTR; NLC:= NIJTOTR; NIC:= NI/I'OTR; NFC:= NF/I‘OTR; NW C:= NW/FOT‘R; NYC:= NY [I OTR; 226 NHC:= NH/IOTR; NCC:= NC/I'OTR; NMC:= NMfI'OTR; NSC:= NS/FOTR; NTC:: NT/TOTR; NKC:= NK/TOTR; NRC:= NR/TOTR; NDC:: ND/TOTR; NEC:= NE/I‘OTR; NNC:= NN/I'OTR; NQC:: NQ/I‘OTR; END; WRITELNCPRESS ANY KEY TO CONTINUE); READ(PAUSE); CLRSCR; X := FALSE; IF X = TRUE THEN BEGIN WRITELNCGLY = ',NG); WRITELN('PRO = '.NP); WRITELNCALA = ',NA); WRITELNCVAL = ',NV); WRITELNCLEU = ',NL); WRITELNCILE = ',NI); WRITELNCPHE = ‘,NF); WRITELNCTRP = ',NW); WRITELNCTYR = ',NY); 227 WRITELNCHIS = '.NH); WRITELNCCYS = ',NC); WRITELNCMET = ',NM); WRITELNCSER = ',NS); WRITELNCTHR = ',NT); WRITELNCLYS = ',NK); WRITELNCARG = ',NR); WRITELNCASP = ',ND); WRI'I‘ELN('GLU = ',NE); WRITELNCASN = ',NN); WRITELNCGLN = ',NQ); WRITELN; WRITELNCTOT = ',TOT); READ(PAUSE); END; CLRSCR; GOTOXY(1,1); WRITE(NAMEIN); GOTOXY (1,2); WRI'I‘E('AMINO ACID'); GOTOXY (15,2); WRITE(NUMBER'); GOTOXY (30,2); WRITE ('PERCENT COMPOSITION); GOTOXY ( 1,3); WRITELN('GLY',' ',NG,‘ WRITELNCPR ',' ',NP,’ '.RD(NGfl‘OTR)); ',RD(NPrrOTR)); 228 WRITELN('ALA',‘ ',NA,' '.RD(NAfrOTR)); WRITELN('VAL‘,‘ ',NV,‘ ',RD(NV/TOTR)); WRITELN('LEU‘,‘ ',NL,‘ ',RD(NLfl‘OTR)); WRITELN('ILE',‘ ',NI,‘ .RD(NI/TOTR)); WRITELN('PI—IE',‘ '.NF.' '.RD(NF/TOTR)); WRITELN('TRP',‘ '..'NW '.RD(NW/1‘ CTR»; WRITELN('TYR',' '.NY.' '.RD(NY/TOTR)); WRITELN('HIS',‘ '.NH.' '.RD(NH/l'OTR)); WRITELN('CYS',‘ ',NC,‘ ',RD(NC/TOTR)); WRITELN('MET',‘ ,,'NM '.RD(NM/rOTR)); WRITELN('SER',‘ ',NS,‘ ',RD(NS/I‘OTR)); WRITELN('THR',‘ ',NT,’ ',RD(NT/TOTR)); WRITELN('LYS',‘ '..'NK '.RD(NK/I‘OTR)); WRITELN('ARG',’ ',NR,’ '.RD(NR/I‘OTR)); WRITELN(‘ASP',’ '.'.ND '.RD(ND/r0TR)); WRITELN('GLU',’ ',NE,‘ ',RD(NE/I‘OTR)); WRITELN(’ASN',‘ '.'.NN '.RD(NN/I‘OTR)); WRITELN('GLN',‘ '.NQ.' '.RD(NQfl‘0TR)); WRITELN('H-PHOBIC',’ ',RD((NG+NA+NL+NI+NV+NP+NF+NW+NY)/TOTR)); WRITELN('POLAR',‘ ',RD((NS+NT+NH+NM+NC)/TOTR)); WRITELN('CHARGED',‘ '. RD((NK+NR+NQ+NN+NE+ND)/TOTR)); X := FALSE; IF X = TRUE THEN BEGIN WRITELN(LST,NAMEIN); 229 WRITELN(LST,‘AMINO ACID',‘ ','NUMBER',‘ 3% COMP'); WRITELN(LST,‘GLY',' ',NG,’ ',PRD(NG/TOTR»; WRITELN(LST,‘PRO',' ',NP,‘ ',PRD(NP/TOTR»; WRITELN(LST,'ALA',' '.NA,' ',PRD(NA/TOTR)); WRITELN(LST,'VAL',' ',',NV ',PRD(NV/I‘OTR)); WRITELN(LST,'LEU',' ‘,NL,‘ ',PRD(NL/TOTR»; WRITELN(LST,'ILE',' ',NI,' ',PRD(NI/I‘OTR»; WRITELN(LST,‘PHE',' ',NF,' ',PRD(NF/I‘OTR»; WRITELN(LST,'TRP',' '..'NW ',PRD(NWrrOTR»; WRITELN(LST,'TYR',' ",NY, ',PRD(NY/I‘OTR»; WRITELN(LST,'HIS',' ',,'NH ',PRD(NH/I‘OTR»; WRITELN(LST,‘CYS',' ',NC,’ ',PRD(NC/I‘OTR»; WRITELN(LSTHMET ',,'NM ',PRD(NM/TOTR»; WRITELN(LST,'SER',' ',NS,‘ ',PRD(NS/TOTR»; WRITELN(LST,"I'I-IR',' ',NT,‘ ',PRD(NT/I‘OTR»; WRITELN(IST,’I.YS',' '..'NK ',PRD(NKfl‘OTR»; WRITELN(LST,‘ARG',' ',NR,' ',PRD(NR/TOTR»; WRITELN(LST,’ASP',' ',ND,‘ ',PRD(ND/rorR»; WRITELN(LST,‘GLU',' ',NE,‘ ',PRD(NE/TOTRD; WRITELN(LST,‘ASN',' ',,'NN ',PRD(NN/I‘OTR»; WRITELN(LST,'GLN',' ',NQ,‘ ',PRD(NQ/TOTR»; WRITELN(LST,‘H-PHOBIC’,' ‘,PRD((NG+NA+NL+NI+NV+NP+NF+NW+NY)/TOTR)); WRITELN(LST,’POLAR',' ',PRD((N S +NT+NH+NM+NC)/I‘OTR)); WRITELN(LST,'CHARGE ',' ',PRD((NK+NR+NQ+NN+NE+ND)ITOTR)); 230 END; READ(PAUSE); CLRSCR; WRITELN('ADD ANOTHER SEQUENCE? (Y/N)'); READ(YNT); WRITELN; YESNO; J := i; i := i + l; UNTIL YN = FALSE; WRITELN('NUMBER OF SEQUENCES = ',J); READ(PAUSE); 231 NRT := 0; NDT := 0; NET := O; NNT := 0; NQT := 0; SQNAT := 0; SQNGT := 0; SQNPT := 0; ' SQNVT ;= 0; SQNLT := O; SQNIT := 0; SQNFI‘ := 0; SQNWT := 0; SQNYT := 0; SQNHT := 0; SQNCI‘ := 0; SQNMT := 0; SQNST := 0; SQN'I'I‘ := 0; SQNKT := 0; SQNRT := 0; SQNDT := 0, SQNET := O; SQNNT := 0; SQNQT := 0; FORi := 1 TO I DO BEGIN 232 WITH COMPARE[i] DO BEGIN NGT := NGT + NGC: SQNGT := SQNGT + SQR(NGC); NPT := NPT + NPC: SQNPT := SQNPT + SQR(NFC); NAT := NAT 4» NAC; SQNAT := SQNAT + SQR(N AC); NLT := NLT + NLC; SQNLT := SQNLT + SQR(NLC); NIT := NIT + NIC; SQNIT := SQNIT + SQR(NIC); NVT := NVT + NV C; SQNVT := SQNVT + SQR(NVC); NYT := NYT + NYC; SQNYT := SQNYT + SQR(NY C); NFI‘ := NFI‘ + NFC; SQNPT := SQNFI‘ + SQR(NFC); NWT := NWT + NWC: SQNWT := SQNWT + SQR(NWC); NHT ;= NHT + NHC; SQNHT := SQNHT + SQR(NHC); NCT := NCI‘ + NCC: SQNCT := SQNCI‘ + SQR(NCC); NMT := NMT + NMC; SQNMT := SQNMT + SQR(NMC); NST := NST + NSC; 233 SQNST := SQNST + SQR(NSC); NTI‘ := N'I'I‘ + NTC; SQN'I'I‘ := SQNTT + SQR(NT C); NRT := NRT + NRC; SQNRT := SQNRT + SQR(NRC); NKT ;= NKT + NKC; SQNKT := SQNKT + SQR(NKC); NNT := NNT + NNC; SQNNT := SQNNT + SQR(NNC); NQT := NQT + NQC; SQNQT := SQNQT + SQR(NQC); NET := NET + NEC; SQNET := SQNET + SQR(NEC); NDT := NDT + NDC; SQNDT := SQNDT + SQR(NDC); END; END; MGT := NGT/J; MPT := NPT/J; MAT := NAT/J; MLT := NLT/J; MIT := NIT/J; MVT := NVT/J; MYT := NYT/J; MFT := NFI‘IJ; MWT := NWT/J; MHT := NHT/J; I! 234 MCT := NCI’IJ; MMT := NMT/J; MST := NST/J; MTI‘ := N'IT/J; NIRT := NRT/J; MKT := NKT/J; MNT := NNT/J; MQT := NQT/J; MET := NET/J; MDT := NDT/J; GTSD := SQRT((J*SQNGT - SQR(NG'I'))/(J*(J-1))); PT SD := SQRT((J*SQNPT - SQR(NPD)/(J*(J-l))); ATSD := SQRT((J*SQNAT - SQR(N AT»/(J *(J -1))); LTSD := SQRT((J*SQNLT - SQR(NLD)/(J*(J-l))); ITSD 3= SQRT((J*SQN1T ' SQR(NH'))/(J*(J-1))); VTSD := SQRT((J*SQNVT - SQR(NVT))/(J*(J-l))); YTSD := SQRT((J*SQNYT - SQR(NYD)/(J*(J-l))); FI‘SD := SQRT((J*SQNFT - SQR(NFI'))/(J*(J-l))); WTSD := SQRT((J*SQNWT - SQR(NW)/(J*(J-l))); HTSD := SQRT((J*SQNHT - SQR(NH'I'))/(J*(J-1))); CT SD := SQRT((J*SQNCI' - SQR(NCI')N(J*(J-l))); MT SD := SQRT((J*SQNMT - SQR(NMT))/(J*(J-l))); STSD := SQRT((J*SQNST - SQR(NST))/(J*(J-l))); TI'SD := SQRT((J*SQN'IT - SQR(Nm)/(J*(J-1))); RTSD := SQRT((J*SQNRT - SQR(NRDWPU-l)»; KTSD := SQRT((J*SQNKT - SQR(NKT))/(J *(J -1))); NT SD := SQRT((J*SQNNT - SQR(NNT))/(J*(J-l))); 235 QTSD == SQRT((J*SQNQT - SQR(NQD)/(J*(J-1))); ETSD := SQRT((J*SQNET - SQR(NED)/(J*(J-1))); DTSD := SQRT((J*SQNDT - SQR(NDD)/(J*(J-1))); CLRSCR; GOTOXY(60,2); WRITE(CATEGORY); GOTOXY(1,2); WRITE(‘AMINO ACID'); GOTOXY(22,2); WRITE(‘MEAN'); GOTOXY(42,2); WRITE ('STD DEV'); GOTOXY(1,3); WRITELN('GLY',‘ '.RD(MGT). '.RD(GTSD»; WRITELN('PRO',’ .RD(MPT).' '.RD(PI‘SD»; WRITELN('ALA',‘ RD'(MA.'T) '.RD(ATSD»; WRITELN('VAL',‘ ,RD(MVT) ' '.RD(VTSD»; WRITELN('LEU',‘ '.RD(MLT)‘ '.RD(LTSD»; WRITELN('ILE',‘ ',RD(MIT),' '.RD(ITSD»; WRITELN('PHE',‘ ,RD(MFI‘),' '.RD(FTSD»; WRITELN('TRP',‘ ','.RD(MWT) '.RD(WTSD»; WRITELN('TYR',' ‘.RD(MYT).' '.RD(YTSD»; WRITELN('HIS',‘ ',RD(MHT),' '.RD(HTSD»; WRITELN('CYS',‘ ',RD(MCI'),' '.RD(CTSD»; WRITELN('MET',‘ ',RD(MMT),' '.RD(MTSD)); WRITELN('SER',‘ ',RD(MST).' '.RD(STSD)); WRITELN('I'HR'.’ '.RD(M'IT).' '.RD(TTSD»; 236 WRITELN('LYS',’ '.RD(MKT),' '.RD(KTSD»; WRITELN('ARG',’ ',RD(MRT).' '.RD(RTSD»; WRITELN('ASP',’ ',RD(MDT),' '.RD(DTSD»; WRITELN('GLU',‘ ',RD(MET),' ',RD(ETSD)); WRITELN('ASN',‘ ',RD(MNT),' ',RD(NTSD»; WRITELN('GLN',‘ ’,RD(MQT),' ',RD(QTSD)); WRITELN('H-PHOBIC',’ ',RD(MGT+MAT+MLT+MIT+MVT+MPT+MFI‘+MWT+MYT)); WRITELN('POLAR',‘ ',RD(MST+MTI‘+MH'I‘+MMT+MCI')); WRITELN('CHARGED',’ ',RD(MT+MRT+MQT+MNT+MET+MDT)); GOTOXY (60,4); WRITE(FILES USED'); FORi := 1T0 I DO BEGIN WITH COMPARE[i] DO BEGIN GOTOXY(60,(i + 4)); WRITE(LABELC); END; END; READ(PAUSE); CLRSCR; WRITELN('ANOTHER SET OF FILES (Y/N)?'); READ(YNT); WRITELN; YESNO; UNTIL YN = FALSE; 237 END. program Arrheniussimulation; VAR DATOUT: TEXT; X,Y,Z: STRING[20]; TOPT,EA,DT,T,FK,TOLD: REAL; TF: BOOLEAN; BEGIN ASSIGN(DATOUT,'ARRSIM.DAT'); WRITELN('INPUT T0pt (K)'); READ(I’OPT); WRITELN; WRITELN('INPUT Ea (kl/MOLD; READ(EA); WRITELN; TOLD := 200, TE := FALSE; RESET(DATOUT); APPEND(DATOUT); WRITELN(DATOUT,'T'0pt = ',TOPT,‘ '.'Ea = '.EA); WRITELN(DATOUT); REPEAT DT := TOPT-TOLD; FK := EXP(((EA*1000)/8.314)*((1/TOPT) - (l/TOLD)»; 238 STR(TOLD,X); STR(DT,Y); STR(FK.Z); WRITELN(DATOUTX,Y,Z); T := TOLD + 2; IF TOLD >= (TOPT) THEN BEGIN TF := TRUE; END; TOLD := T; UNTIL TF=TRUE; QOSEmAmUD; END. APPENDIX C 239 Appendix C Construction and Kinetic Characterization of Thermoanaerobacter ethanolicus 39E 2° ADH Proline Deficient Mutants INTRODUCTION Expression of the genes encoding thermOphilic and thermostable enzymes in mes0philic hosts typically yields recombinant enzymes with native-like thermal properties. Therefore, the molecular determinants of these characteristics appear to be in the amino acid sequences themselves. The general amino acid compositional similarity between mesophilic and thermophilic proteins ftu'ther suggests that the positions of specific amino acids and not their number are critical to enzyme thermal stability. Finally, the significant energetic contributions of individual amino acids to the marginal stabilizing energy of folded protein su-uctures indicates that. similar to proper protein folding, enzyme thermostability is due predominantly to energetic contributions from a subset of peptide amino acids. However, the identities and positions of these critical amino acid residues cannot yet be predicted even with exact 3-dimensional structure information. Proline residues have been proposed to constrain protein surface loops and this is proposed to stabilize the interactions between the associated hydrophobic core elements [1]. This proline-zipper hypothesis agrues that by preventing the disassociation of the surface most inter-core-element interactions, a constrained loop thermostabilize the folded protein. Loop engineering has been shown to thermostabilize protein folded structures (see chapter 1; prolines and loop regions). Furthermore, introducing Pro residues into tum regions has 240 thermostabilized both mesophilic and therm0philic proteins (see chapter 1; prolines and loop regions). The identification of thermophilic enzymes containing additional Pro residues compared to their mesophilic counterparts has suggesterd that this is also a protein thermostabilization strategy used in nature (see chapter 1; prolines and loop regions). The availability of functionally analogous thermophilic and mesophilic proteins with highly similar peptide sequences would allow comparative analysis to identify potentially critical Pro residues. The recently cloned thermophilic Thermoanaerobacter ethanolicus 2° ADH amino acid sequence differs from that deduced for the mesophilic Closm'dium beijerincla’i 2° ADH fiom its cloned gene at 89 positions (see chapter 2). Of these, approximately 12 are nonconservative and 9 of these are Pro residues in putative loop regions of the more thermophilic enzyme. The overall sequence similarity among these 2 peptides is >85% (See chapter 2). This is significantly higher than the similarity seen between the T. ethanolicus 2° ADH and the horse liver 1° ADH (54%), yet mutagenic structure-function analysis suggested that they were also architecturally analogous (see chapter 4). Because the mesophilic C. beijerincla’i 2° ADH shares greater peptide sequence identity to the thermophilic T. ethanolicus enzyme than the horse liver 1° ADH, the mesophilic C. beijerinckii 2° ADH is also predicted to share greater folded structural similarity with the thermophilic T. ethanolicus 2° ADI-I. Therefore, this is an excellent model system for mutagenic analysis to test the hypothesis that proline residues, constraining surface loops, can prevent hydrophobic core element separation and so thermostabilize proteins. The data presented here examine the effect of mutating 3 individual T. ethanolicus 2° ADH added Pro residues to the corresponding C. beijerinckii enzyme residues predicted based on sequence alignments. The thermOphilicities and thermostabilities of the wild type and proline mutant enzymes are reported 241 MATERIALS AND METHODS Chemicals and reagents All chemicals were of at least reagent/molecular biology grade. Oligonucleotide synthesis and amino acid sequence analysis were performed by the Macromolecular Structure Facility (Department of Biochemistry, Michigan State University). The kanamycin resistance GenBlock (EcoRI) DNA cartridge used in expression vector construction was purchased from Pharmacia (Uppsala, Sweden). DNA for sequencing was isolated using the Wizard Miniprep kit (Promega; Madison, WI). Media and strains Escherichia coli (DHSOt) containing the 2° ADH recombinant plasmids were grown in rich complex medium (20 g l'1 tryptone, 10 g l'1 yeast extract, 5 g 1'1 NaCl) at 37°C in the presence of 25 ug ml-1 kanamycin and 100 ug ml:1 ampicillin. Muta genesis All DNA manipulations were performed using established protocols [2,3]. Point. mutations were introduced into the (1th gene by PCR [2] using the T. ethanolicus 39E 2° ADH gene clonal plasmid pADHB25-kan (see chapter 2) as template DNA. An oligonucleotide primer (KA4 N-end) was synthesized to bind the noncoding strand and included a Kpnl restriction enzyme site, the native ath gene ribosome assembly site, and the initiation codon for the ath gene. An oligonucleotide primer (KA4 C-end) was synthesized to bind the coding strand, it included the compliment of the ath termination codon and an Apa! restriction enzyme site. Complimentary 30-45 base oligonucleotide primers that contained the mutated bases were used in conjunction with 2 KA4 end primers to amplify the N-terminal and C- terminal segments of the ath gene. PCR syntheses of partial and complete mutated genes were performed using the Taqplus, exonuclease containing polymerase (Strategene; La Jolla, CA). All clones were expressed in pBluescriptII KS(+) with a kanamycin resistance cartridge introduced into the polylinker EcoRI site. Mutations were verified by DNA sequencing using the method of Sanger et al. [4]. 242 Enzyme kinetics The standard 2° ADH activity assay was defined as NADP“ reduction coupled to propan-2- ol oxidation at 60°C as previously described [5]. The enzyme was incubated at 55°C for 15 min prior to activity determination unless otherwise indicawd. Tris buffer pH was adjusted at 25°C to be pH 8.0 at assay temperature (thermal correction factor = -0.031 ApH °C'1). Protein concentrations were measured using the bicinchoninic acid (BCA) procedure (Pierce; Rockford, IL). Thermostability Recombinant 2° ADH thermostabilities were evaluated in E. coli cell extracts by timed incubation at the desired temperatures, followed by incubation for 30 min at 25°C. Incubations were performed in 100 p1 PCR tubes (cat. #72.733.050, Sarstedt; Newton, NC) using 0.2 mg ml'1 protein in 100 ul of 50 mM TriszHCl (pH 8.0). Activity was determined using the unfractionated samples. RESULTS The percent activities of the single and duplicated N—terminal Met containing wild type T. ethanolicus 2° ADHs and the H022 to Ala, Prol49 to Thr, and H0222 to His mutants were compared at temperatures from 30°C to 90°C (Fig. 1A). The catalytic activities calculated for thecellextracts at90°Crangedfrom 6.0unitspermg total protein forthePro149 toThr mutant to over 60 units per mg protein for the wild type enzymes. However, the percent activities increased similarly for all enzymes from 30°C to 90°C. The enzyme residual activities after timed incubations at 90°C were determined (Fig. 1B), indicating that thermoinactivation of both the wild type and mutant enzymes was not pseudo-first order in E. coli cell extracts. Also, the proline deficient mutant enzymes demonstrated greater thermostability than either of the wild type 2° ADHs. 243 Figure l. Thermophilicity and thermostability profiles for T. ethanolicus 2° ADH Pro residue mutants. (A) The effect of temperature on propan-2-ol oxidation by recombinant wild type (0). Single N-terminal Met (0), Pro24 to Ser cl), Prol49 to Thr (A), and Pr0222 to His fl), enzyme containing E. coli DHSa cell extracts. (B) The efi'ect of timed incubations at 90°C on wild type (0), Single N-terminal Met (0), Pr024 to Ser Cb, ’ Prol49 to Thr (A), and Pm222 to His C), enzyme containing E. coli DHSa cell extract residual percent activities. 244 1.4 1.2d 08 6 .4 024 u 0. 1 3.2.2. 0:358 .2552: ._e .8322..— 0.0 100 20 Temperature (°C) 245 o 2 1 0 ID 0 IO 1 AID 0 l0 8 AID .. 0 AID I6 AID . lo I D O 4 AI D O .. AID O o I D .0 I2 I D00 I n! .. ID. F. . u 0 0 an. 9.. 3 2e: 53.8 .32825 Time (min) 246 DISCUSSION These preliminary data suggest that the 3 single Pro mutations tested do not significantly affect T. ethanolicus 2° ADH thermophilicity and thermostability. The wild type enzyme temperature activity profile is similar to that reported for pmified enzyme (chapter 3). The temperature dependence of mutant and wild type enzyme activities were similarly except for the Prol49 to Thr mutant which may be inactivating at temperatures between 70°C and 90° C. The decreased slope for average percent activity versus temperature above 70°C compared to that below 70°C for this mutant is not consistent with wild type enzyme behavior and suggests loss of optimal active enzyme structure in at least part of the enzyme population. Furthermore, because the maximum Prol49 to Thr activity is less than that expected from extrapolating the lower temperature activity data to 90°C, dividing the lower temperature specific activities by the maximal activity (that at 90°C) to obtain the percent of maximum catalytic rate would artificially raise the percent activity values at lower temperatures. Recalculating the data using the temperature at 70°C as 100% activity indicates that all ofthe mutants and the wild type enzyme had similar temperature activity profiles below 70°C. Therefore replacing Pro 24 and Pr0222 with the corresponding C. beijerinckii 2° ADH amino acids did not appear to significantly alter enzyme thermophilicity but replacing Prol49 may have reduced enzyme thermophilicity. However, activity determinations must be repeated using pmified enzyme to confirm these findings. The Pro residue deficient mutants all displayed similar thermostabilities which were greater than either wild type enzyme. The relationship between the logarithm of residual percent activity and incubation time however, was nonlinear for all 5 enzymes. Precipitate was seen in the thermoinactivated samples, consistent with data reported for the purified wild type enzyme. The failure of even the wild type enzyme data collected using cell extracts to fit a pseudo-first order rate equation suggests that the thermoinactivation slow step under these conditions differs from that for purified enzyme. The relatively rapid loss of 2° ADH activity in all 5 extracts compared to the purified enzyme further suggests that 247 purification stabilizes the 2° ADH against precipitation. While none of the Pro deficient mutants demonstrated a substantial loss of thermostability, the inconsistency of the cell extract data with that reported for the purified wild type enzyme makes analysis difficult. The entire set of nine added T. ethanolicus 2° ADH Pro residue mutations must be examined individually and together before drawing any conclusions but the preliminary evidence presented here using these 3 Pro mutants indicates that T. ethanolicus 2° ADH thermostability must be measured using purified enzyme. 248 REFERENCES Vieille, C., Burdette, D. S. and Zeikus, J. G. [1996] Thermozymes, in Biotechnology Annual Reviews, vol. 2, (M. R. El Geweley, ed.) pp. 1-83. Elsevier Press, Amsterdam, Neth. Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, J. A. and Struhl, K. (1993) Current Protocols in Molecular Biology (Janssen, K., ed.), Current Protocols, NY Sambrook, J ., Fritsch, E. F. and Maniatis, T. (1989) Molecular cloning: A laboratory manual 2051 edition. (Nolan, C., ed.), Cold spring Harbor Press, NY Sanger, F., Nicklen, S. and Coulson, A. R. (1977) Proc. Natl. Acad. Sci. USA 74, 5463-5467 Burdette, D. S. and Zeikus, J. G. (1994) Biochem. J. 302, 163-170 MICHIGAN STATE UNIV. LIBRQRIES lllWWIIWI”11111111“llNIHIHIHIIHWIlHI 31293014214930