, f 21x. .3?“ a: z . 3‘53 w. 5.3.)? n 31:21.3... . , V... A, _ _ ..._2 1r..7_ _. 7; i1 1 W‘SIS ‘ l This is to certify that the dissertation entitled GENETIC ENGINEERING 0F XYLOSE ISOMERASE THERMOZYMES FOR ENHANCED ACTIVITY, STABILITY, AND UTILITY presented by DINLAKA SRIPRAPUNDH has been accepted towards fulfillment of the requirements for Doctoral Food Science degree in y ‘sz 7- M Date ,2'13‘ 2002 MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 GENE GENETIC ENGINEERING OF XYLOSE ISOMERASE THERMOZYMES FOR ENHANCED ACTIVITY, STABILITY, AND UTILITY By Dinlaka Sriprapundh A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Food Science and Human Nutrition 2002 GEI isome identi‘ less ABSTRACT GENETIC ENGINEERING OF XYLOSE ISOMERASE THERMOZYMES FOR ENHANCED ACTIVITY, STABILITY, AND UTILITY By Dinlaka Sriprapundh Molecular determinants responsible for high thermostability of the xylose isomerase from hyperthermophilic eubacterium Thermotoga neapolitana (TNXI) were identified by comparative thermostability and site-directed mutagenesis studies with the less thermostable counterpart enzyme from thermophilic eubacterium T hermoanaerobacterium thermosulfitrigenes (TTXI). Despite their highly similar structures and amino acid sequences (70.4 % identity), no obvious differences in the enzyme structures can explain the differences in TNXI’s stability compared to that of 'ITXI except for a few additional prolines and fewer Asn+Gln in TNXI. TNXI has 2 additional prolines in the Phe59 loop (Pr058 and Pro62). This loop helps forming another enzyme subunit’s active site. When the 2 prolines in TNXI were substituted with the corresponding amino acids present in its less thermostable counterpart, TTXI, all mutant enzymes showed significant loss in thermostability compared to the wild-type TNXI. These data confirmed the hypothesis that prolines indeed play important roles in TNXI thermostability by reducing its entropy of unfolding. TNXI’s active site was engineered to improve its catalytic efficiency toward glucose. The TNXI V185T mutant derivative was three times more efficient in glucose isomerization than the wild-type TNXI. Although this mutant derivative was highly thermostable and highly active at 97°C, it was less than 10% as active at 60°C and required net evolution at low temper: low tempera exhibited dr low pH as mutation rel; active site of pHs. TNXI l resulth fror neighboring } conformation Bloch: awmmaou higher cataln gm“? Iherm. aClll’lty. Tl'lls glucose Wilde] Process based, required neutral pH to work. To customize this TNXI mutant derivative, a directed evolution approach was applied to the TNXI V185T to improve its activity on glucose at low temperature and low pH. After two successive rounds of random mutagenesis and low temperature/low pH activity screening, a new mutant, TNXI 1F1, was obtained that exhibited dramatic improvement of glucose isomerase activity at low temperature and low pH as compared to TNXI V185T. TNXI 1F 1 (V185T/L282P/F186S) with one mutation relatively distant (L282P) from the active site and the other (F1868) within the active site of the enzyme, was more active than TNXI V185T over all temperatures and pHs. TNXI 1F] was also more stable than TNXI and TNXI V185T and this may have resulted from additional H-bond formation between Ser186’s sidechain and the neighboring L229 residue’s mainchain structure. This H-bond would strengthen local conformation and the affinity of E231 co-ordination with the structural metal. Biochemical properties and fructose productivities of TNXI 1F] and GensweetTM, a commercially available glucose isomerase were also compared. TNXI lFl displayed higher catalytic efficiencies on glucose at low or high temperature and pH ranges and had greater thermal stability than GensweetTM despite having similar temperature optima of activity. This greater thermal stability together with the superior kinetic parameters on glucose render TNXI lFl an excellent candidate for the industrial glucose isomerization process based on the lifetime fructose productivity estimation. Copyright by Dinlaka Sriprapundh 2002 Dedicated to my family, Mr. Udom and Mrs. Achara Sriprapundh ml J. Grcgt guidance Motenoi research DTS. IOhI th‘enic lhroughr Kalli-aim COmPUIt slide- m.- M. Kelly Ontne p, immobil OPPOnUn I ACKNOWLEDGEMENTS I would like to express my sincere appreciation and deepest gratitude to Professor J. Gregory Zeikus, who is my scientific advisor and philosophical mentor, for his guidance, encouragement, and patience through my graduate school years. I cannot thank him enough for all the support he has provided me for the successful completion of this research project. I would also like to thank all of the members of my guidance committee, Drs. John Linz, James Pestka, and Gale Strasburg, for their advice and invaluable time. I am grateful to my laboratory members especially Dr. Claire Vieille, Maris Laivenieks, and Su-il Kang for all their support, friendship, advice, and encouragement throughout my graduate program. I would also like to acknowledge the help from Dr. Kaillathe Padmanabhan for computer aided molecular modeling in Silicon Graphic Computer System, insightful discussion of protein structure, and assistant in poster and slide- making as well as other computer-related troubles. I also should thank Dr. Robert M. Kelly and Kevin Epting at the North Carolina State University for collaboration work on the project and providing facilities and advice for the DSC measurement and enzyme immobilization. I owe my success to my former advisor Dr. Sirirat Rengpipat for giving me the opportunities, her enormous support and encouragement through tough times. Without her assistance, this work could not have been completed. I am most grateful to my best friend Paweena Limjaroen for her dedication and encouragement. I could not have undertaken this endeavor without her help and vi UEJC while understanding. My deepest appreciation goes to my family and friends who supported me while accomplishing this work. vii L151 L157 (TIA Liter. l CHAP‘ ‘ Role of l Themu: ‘ .\\ lOsc Xylem , TABLE OF CONTENTS page LIST OF TABLES .................................................................................... x LIST OF FIGURES ................................................................................... xi CHAPTER I Literature Review ..................................................................................... l Xylose isomerase and high fructose corn syrup ......................................... 2 X1 structure, sequence homology and reaction mechanisms ........................... 5 Thermozymes and their thermal stability ................................................ 13 Genetic and protein engineering of XIs ................................................. 16 Thermostability mutations ....................................................... 16 Active site mutations .............................................................. 20 Directed evolution of enzymes ........................................................... 22 References .................................................................................. 30 CHAPTER II Role of Proline in Xylose Isomerase Thermal Stability: Comparative Thermostability of a Thermophilic Thermoanaerobacterium thermosulfurigenes Xylose Isomerase (TTXI) and a Hyperthermophilic Thermotoga neapolitana Xylose Isomerase (TNXI) .......................................................................... 36 Abstract ...................................................................................... 37 Introduction ................................................................................. 38 Materials and Methods ..................................................................... 41 Results ....................................................................................... 45 Discussion ................................................................................... 52 References ................................................................................... 57 CHAPTER HI Role of Substrate Binding Pocket in Thermotoga neapolitana Xylose Isomerase (TNXI) on Glucose Isomerase Activity .......................................................... 59 Abstract ...................................................................................... 60 Introduction ................................................................................. 6 1 Materials and Methods ..................................................................... 63 Results ....................................................................................... 67 Discussion ................................................................................... 70 References ................................................................................... 73 CHAPTER IV Directed Evolution of Thermotoga neapolitana Xylose Isomerase: High Activity on Glucose at Low Temperature and Low pH ................................................... 74 Abstract ...................................................................................... 75 Introduction ................................................................................. 77 viii CHAI Bioch- an Ind Ewfl\ CHAP Role 0 CH AP’ COnclu page Materials and Methods ..................................................................... 80 Results ........................................................................................ 85 Discussion .................................................................................... 99 References ................................................................................. 105 CHAPTER V Biochemical Comparison of Stability, Activity, and Fructose Synthesis by an Industrial Glucose Isomerase (GENSWEETW) versus a Laboratory- Evolved Thermo-Acid Stable Xylose Isomerase (TNXI lFl) ............................... 108 Abstract ..................................................................................... 109 Introduction ................................................................................ 1 10 Materials and Methods ................................................................... 1 13 Results ...................................................................................... 1 17 Discussion ................................................................................. 13 1 References ................................................................................. 1 36 CHAPTER VI . Role of Metals in TNXI Stability at Extremely High Temperatures ........................ 138 Abstract ..................................................................................... 139 Introduction ................................................................................ 1 40 Materials and Methods ................................................................... 142 Results ...................................................................................... 143 Discussion ................................................................................. 147 References ................................................................................. 149 CHAPTER VII Conclusions and Directions for Future Research .............................................. 150 ix .11 CHAPT Tbkl. CHAPT Lmkl. luHeZ. IuMe3. CHAPT' IaMcl, CHAPT, lame]. CHAPT; Tdble 1. LIST OF TABLES CHAPTER I Table 1. Examples of enzymes that were successfully optimized using directed evolution ................................................................................ 28-29 CHAPTER II Table 1. Oligonucleotides and DNA templates used for site-directed mutagenesis ........ 42 Table 2. Effect of mutations on enzyme precipitation temperatures .......................... 50 Table 3. Catalytic parameters of TTXI (at 65°C) and TNXI (at 80°C) and of their mutant derivatives ......................................................................... 51 CHAPTER HI Table 1. Oligonucleotides and DNA templates used for site-directed mutagenesis ................................................................................. 64 Table 2. Catalytic parameters of TTXI (at 65°C) and TNXI (at 80°C) and of their mutant derivatives ......................................................................... 69 CHAPTER IV Table 1. Kinetic parameters of TNXI and its mutant derivatives .............................. 91 CHAPTER V Table 1. Comparison of glucose isomerization kinetic parameters of TNXI 1F 1 vs. GensweetTM .............................................................. 125 Table 2. Comparison of thermal activity and stability properties of TNXI lFl vs. GensweetTM ............................................................ 132 Table 3. Comparison of fructose production parameters of TNXI lFl vs. GensweetTM ............................................................. 134 CHAPTER VI Table 1. Melting temperatures (Tm) of TNXI and its mutant derivatives in the presence (holo-enzyme) and absence (apo-enzyme) of 5 mM MgSO4 and 0.5 mM COCl2 as determined by DSC ......................... 146 CHA Figur Figure Figure Figure FJSUre. CHAPT Figure 1 ‘EUre 2 LIST OF FIGURES “Images in this dissertation are presented in color” CHAPTER I Figure 1. Illustration of the reactions carried out by xylose isomerases (XIs) ................ 3 Figure 2. Starch processing into high fructose corn syrup (HFCS) ............................ 4 Figure 3. Three-dimensional structure of a subunit of Thermotoga neapolitana xylose isomerase (type 11 XI) ............................................................ 7 Figure 4. Homology alignment of xylose isomerase amino acid sequences from different organisms by the ClustalW analysis ..................................... 9-10 Figure 5. Schematic representation of the active site of TNXI ................................ 1 1 Figure 6. A metal-assisted hydride-shift mechanism of xylose in the Streptomyces rubiginosus xylose isomerase (type I XI) ............................. 14 CHAPTER 11 Figure 1. Three-dimensional structure of TNXI’s Phe59 loop region ........................ 46 Figure 2. Effect of temperature on the specific activities of TFXI and TNXI and their Phe59 loop mutant derivatives .................................................. 47 Figure 3. Inactivation curves of TTXI and TNXI and their Phe59 loop mutant derivatives at 85°C (A) (TTXI and derivatives) and 95°C (B) (TNXI and derivatives) ................................................................. 48 Figure 4. Three-dimensional structure of the Phe59 loop mutations in TTXI ............... 54 Figure 5. Van der Waals contacts between Lys61 and Pro62 in 'ITXI Ala62Pro mutant derivative ........................................................................ 55 CHAPTER HI Figure 1. Activity and stability of TNXI and its active site mutant derivatives ................................................................................. 68 Figure 2. Superposition of TTXI and TNXI active sites ....................................... 71 xi CHAP Figure Figure Figure CHA] CHAPTER IV Figure l. A) Effect of temperature on the specific activities of TNXI and its mutant derivatives on glucose at pH 7.0 and B) Ln (specific activity) versus 1/Temperature ................................... 87-88 Figure 2. Effect of pH on the specific activities of TNXI and its mutant derivatives on glucose at 80°C ......................................................... 89 Figure 3. Inactivation curves of TNXI and its mutant derivatives at A) 80°C, pH 7.0 and B) 80°C, pH 5.5 ........................................................ 92-93 Figure 4. Three—dimensional model of the TNXI 1F] monomer showing the positions of mutations V185T, F1868, and L282P .................................... 96 Figure 5. Three-dimensional model of part of the area surrounding TNXI 3A2 and 1F1’s Leu282Pro mutation ......................................................... 97 Figure 6. Three-dimensional model of part of the TNXI 1F 1 active site showing . hydrogen bonds among 8186, L229, and E231 ....................................... 98 CHAPTER V Figure 1. Effect of temperature on the specific activities of TNXI 1F1 and GensweetTM on glucose at pH 7.0 .................................................... 1 18 Figure 2. Effect of pH on the specific activities of TNXI 1F1 and GensweetTM on glucose at pH 7.0 ................................................................... 1 19 Figure 3. Inactivation curves of TNXI 1F1and GensweetTM at 60°C (pH 7.0) (A) and at 60°C (pH 5.5) (B) .......................................................... 121-122 Figure 4. Thermal unfolding of TNXI 1F1 (A) and GensweetTM (B) in the presence of 5 mM MgSO., and 0.5 mM CoClz followed by DSC ............... 123 Figure 5. Estimated fructose productivity of TNXI 1F] and GensweetTM in different conditions .................................................................... 126 Figure 6. Experimental fructose conversion of TNXI 1F1 and GensweetTM at pH 7.0 and 5.5 at 80°C (A) and at 60°C(B) ...................................... 128 Figure 7. Browness of syrups from experimental fructose conversion of TNXI 1F 1 and GensweetTM ........................................................ 129 xii C HAPTEF Figure 1.T it Figure 3: T lit‘ CHAPTER VI Figure 1. Thermal unfolding of the holo-forms of TNXI and its mutant derivatives in the presence of 5 mM MgSO4 and 0.5 mM CoClz followed by DSC ........ 144 Figure 2: Thermal unfolding of the apo-forms of TNXI and its mutant derivatives followed by DSC ........................................................................ 145 xiii CHAPTER I LITERATURE REVIEW XYLOSE ISOMERASE AND HIGH FRUCTOSE CORN SYRUP Xylose isomerase (XI) (D-xylose ketol isomerase; EC 5.3.1.5), generally known as glucose isomerase (GI), catalyzes the interconversion of D-xylose to D-xylulose in vivo and D-glucose to D-fructose in vitro (Figure 1). Interconversion of xylose to xylulose provides a nutritional requirement in bacteria that thrive on decaying plant materials and also aids in the bioconversion of hemicelluloses to ethanol. Bioconversion of renewable biomass to fermentable sugars and ethanol is important in view of the rapid depletion of fossil fuels. Isomerization of glucose to fructose is of commercial importance in the production of high fructose corn syrup (HFCS) (Bhosale et al., 1996). Xylose isomerase is one of the most important industrial enzymes in the food industry and also one of the three highest tonnage value enzymes, amylase and protease being the other two (Bhosale et al., 1996). The production of HFCS (Figure 2) from cornstarch comprises three major processes: (1) liquefaction of starch by OL-amylase, (2) saccharification of starch slurry by the combined action of amyloglucosidase and debranching enzymes, and (3) isomerization of glucose by X1 (or G1). The final product is a corn syrup containing a mixture of glucose and fructose with a greater sweetening capacity than that of sucrose. An equilibrium mixture of glucose and fructose (1:1) is 1.3 times sweeter than sucrose and 1.7 times sweeter than glucose. The sweetening capacity of glucose is 70 to 75 ‘70 that of sucrose, whereas fructose is twice as sweet as sucrose (Barker, 1976). The price of HFCS is 10 to 20 % lower than that of sucrose on the basis of its sweetening power. H H O H H O H- C-OH 0 OH H OH H OH OH H xylopyranose xylulofuranose CH20H H H H o H CH2OH o 'H- C-OH 01 OH H OH H OH OH H glucopyranose fructofuranose Figure 1: Illustration of the reaction carried out by xylose isomerases (XIs) ED 42 % Fructose , Glucose 1 isomerase ‘iJLI .2.',- ' 9 NaOH Rig“ Deionize [a -Amy1ase I Flucoamylase WW Liquefaction Saccharification Isomerization pH 6.0 pH 4-2 — 4-6 pH 8.0 — 8.2 95-105°c 60°C 55 — 60°C 90 min 48'96 hr Streptomyces, Bacillus licheniformis ASPergillus mm, Bacillus, B. .rtearmhermnnhilus Aspergillus ”i8” Actinnnlanes Figure 2: Starch processing into high fructose corn syrup (HFCS) HFCS is preferred by the food industry since it does not pose the problem of crystallization, as is the case with sucrose. Moreover, fructose plays an important role as a diabetic sweetener because it is only slowly reabsorbed by the stomach and does not influence the glucose level in blood. The major uses of HFCS are in the beverage, baking, canning, and confectionery industries. The annual world consumption of HFCS was estimated to have reached 10 million tons (dry weight) in 1995 (de Raadt et al., 1994). In 1996, HFCS has almost completely replaced sucrose in beverage market in the United States. Beside its industrial importance, xylose isomerase also serves as an interesting model for studying structure- function relationships by advanced biochemical and genetic engineering techniques such as site-directed mutagenesis and directed evolution. XI STRUCTURE, SEQUENCE HOMOLOGY, AND REACTION MECHANISMS To elucidate structure-function relationships of xylose isomerases, the genes encoding XIs (xylA) sequences from various organisms have been compared. XIs can be classified into two groups based on their amino acid sequences and a 40-50 residue N- terminal extension (Vangrysperre et al., 1988). Class I XIs contain about 390 amino acids and consist of those from Streptomyces spp., Actinoplanes spp., Ampullariella spp., Arthrobacter spp., and Thermus thermophilus. The enzymes from Escherichia coli, Bacillus spp., Lactobacillus spp., Lactococcus spp., Thermoanaerobacterium thennosulfurigenes, and Thermotoga spp. contain approximately 440 amino acids and are grouped as class II XIs. The enzymes being studied extensively in this work (Thermoanaerobacterium thermosulfurigenes XI: TTXI and Thermotoga neapolitana XI: TNXI) belong to the class H XIs. The XI monomer consists of an eight—stranded parallel B-barrel surrounded by eight helices with an extended C-terminal tail that provides extensive contacts with a neighboring monomer (Figure 3). The active site pocket is defined by an opening in the barrel of which the entrance is lined with hydrophobic residues while the bottom of the pocket consists mainly of glutamate, aspartate, and histidine residues coordinated to two bivalent cations (MgZT, Mn”, or Coz+)(Whitlow et al., 1991). Three type I and five type H XI’s xylA sequences were aligned based on their sequence homology (Fig 4). In spite of the low homology between classes I and II enzymes, the amino acids involved in the substrate (H100, T140, E231, K233, D338) and metal ions binding (E231, E267, H270, D295, D306, D308, D338), as well as in catalysis (H100, D103, D338), are completely conserved (Fig. 5). Thus, the essential structure at the catalytic center of XIs appears to be analogous in class H XIs so far compared. The subunit structure and amino acid composition of XI reveal that it is a tetramer or a dimer of similar or identical subunits associated with non-covalent bonds and is devoid of interchain disulfide bonds. Most known XIs are homotetramers with molecular masses of approximately 45 to 50 kDa per subunit, although some XIs have been found to be dimeric. The dissociation and unfolding of the tetrameric XI from Streptomyces sp. Strain NCIM 2730 revealed that the tetramer and the dimer are the active species whereas the monomer is inactive (Ghatge et al., 1994). The native Thermotoga neapolitana homotetrameric xylose isomerase is also eXpressed as a catalytically active and thermostable dimer in E. coli (Hess et al., 1998). Figure 3: Three-dimensional structure of a subunit of Thermotoga neapolitana xylose isomerase (type 11 XI). l - 1“ re! ; Figure 4: Homology alignment of xylose isomerase amino acid sequences from different 3 T- Li organisms by the ClustalW analysis (http://www.ebi.ac.uk/clustalw). The sequences are f ‘ T.m., Thermotoga maritima; T.n., Thermotoga neapolitana: T.t., Thermoanaerobacterium thennasulfurigenes; B.s., Bacillus subtilis; lllil 1; B.l., Bacillus licheniformis; E.c., Escherichia coli; A.m., Actinoplanes missouriensis; ' l S.r., Streptomyces rubiginosus; Ar, Arthrobactor sp. (*) means amino acids are identical, 1" “5 7|" m tn " . . ‘1 (2) means conserved substitutions, and (.) means semi-conserved substitutions. ‘ H} l' ' In 7" (I! m In .‘ L! 1') " i h " ii In I ; . .‘I' IN ”I “y I) I ‘l w m y m w w a a a H 3 n F-m n 5 3 v m > m w w a a a H B n F'm n 3 3 H B n H‘m n S 3 y m y m m w 6 6 H > m > m m m a a a H B n H‘m n D S > m y m w m H 6 a H 3 n H‘m n D 3 y m > m m m H a H H 8 0 »:m n S B ------- MAEFFPEIPKIQFEGKESTNPLAFRFYDPNEVIDGKPLKDHLKFSVAFWHTFV ------- MAEFFPEIPKVQFEGKESTNPLAFKFYDPEEIIDGKPLKDHLKFSVAFWHTFV ------- MNKYFENVSKIKYEGPKSNNPYSFKFYNPEEVIDGKTMEEHLRFSIAYWHTFT MAQSHSSSINYFGSANKVVYEGKDSTNPLAFKYYNPQEVIGGKTLKEHLRFSIAYWHTFT --------- MFFRNIGMIEYEGADSENPYAFKYYNPDEFVGGKTMKEHLRFAVAYWHTFD ——————— MQAYFDQLDRVRYEGSKSSNPLAFRHYNPDELVLGKRMEEHLRFAACYWHTFC ———————————————————————————— MSVQATREDKFSFG-~——---——--——LWTVG ———————————————————————————— MNYQPTPEDRFTFG—--—--—-——-—-LWTVG ———————————————————————————— MSVQPTPADHFTFG-—-—-——-—-——-LWTVG * . * NEGRDPFGDPTAERPWNRFSDPMDKAFARVDALFEFCEKLNIEYFCFHDRDIAPEGKTLR NEGRDPFGDPTADRPWNRYTDPMDKAFARVDALFEFCEKLNIEYFCFHDRDIAPEGKTLR ADGTDQFGKATMQRPWNHYTDPMDIAKARVEAAFEFFDKINAPYFCFHDRDIAPEGDTLR ADGTDVFGAATMQRPWDHYKG-MDLAKMRVEAAFEMFEKLDAPFFAFHDRDIAPEGSTLK ADGKDPFGDGTMFRAWNRLTHPLDKAKARAEAAFEFFEKLGVPYFCFHDVDIVDEGATLR WNGADMFGVGAFNRPWQQPGEALALAKRKADVAFEFFHKLHVPFYCFHDVDVSPEGASLK WQARDAFGDATRT ------------- ALDPVEAVHKLAEIGAYGITFHDDDLVPFGSDAQ WQGRDPFGDATRR ------------- ALDPVESVRRLAELGAHGVTFHDDDLIPFGSSDS WTGADPFGVATRK ------------- NLDPVEAVHKLAELGAYGITFHDNDLIPFDATEA 'k it . . . *‘kt *. ETNKILDKVVERIKERMKDSNVKLLWGTANLFSHPRYMHGAATTCSADVFAYAAAQVKKA ETNKILDKVVERIKERMKDSNVKLLWGTANLFSHPRYMHGAATTCSADVFAYAAAQVKKA ETNKNLDTIVAMIKDYLKTSKTKVLWGTANLFSNPRFVHGASTSCNADVFAYSAAQVKKA ETNQNLDMIMGMIKDYMRNSGVKLLWNTANMFTNPRFVHGAATSCNADVFAYAAAQVKKG ETFTYLDQMSSFLKEMMETSHVQLLWNTANMFTHPRYVHGAATSCNADVYAYAAAKVKKG EYINNFAQMVDVLAGKQEESGVKLLWGTANCFTNPRYGAGAATNPDPEVFSWAATQVVTA TR----DGIIAGFKKALDETGLIVPMVTTNLFTHPVFKDGGFTSNDRSVRRYAIRKVLRQ ER----EEHVKRFRQALDDTGMKVPMATTNLFTHPVFKDGGFTANDRDVRRYALRKTIRN ER----EKILGDFNQALKDTGLKVPMVTTNLFSHPVFKDGGFTSNDRSIRRFALAKVLHN *.* *..* . i i . . .. . . . .. LEITKELGGEGYVFWGGREGYETLLNTDLGLELENLARFLRMAVEYAKKIGFTGOFLIEP LEITKELGGEGYVFWGGREGYETLLNTDLGFELENLARFLRMAVDYAKRIGFTGQFLIEP LEITKELGGENYVFWGGREGYETLLNTDMEFELDNFARFLHMAVDYAKEIGFEGQFLIEP LETAKELGAENYVFWGGREGYETLLNTDLKFELDDLARFMHMAVDYAKEIGYTGQFLIEP LDIAKELGAENYVFWGGREGYETLLNTDMKLELENLSSFYRMAVEYAREIGFDGQFLIEP MEATHKLGGENYVLWGGREGYETLLNTDLRQEREQLGRFMQMVVEHKHKIGFQGTLLIEP MDLGAELGAKTLVLWGGREGAEYDSAKDVSAALDRYREALNLLAQYSEDRGYGLRFAIEP IDLAVELGAETYVAWGGREGAESGGAKDVRDALDRMKEAFDLLGEYVTSQGYDIRFAIEP IDLAAEMGAETFVMWGGREGSEYDGSKDLAAALDRMREGVDTAAGYIKDKGYNLRIALEP .." . * *i**** ”k i. . s t. . .** a o. o a o . . o - KPKEPTKHQYDFDVATAYAFLKNHGLDEYFKFNIEANHATLAGHTFQHELRMARILGKLG KPKEPTKHQYDFDVATAYAFLKSHGLDEYFKFNIEANHATLAGHTFQHELRMARILGKLG KPKEPTKHQYDFDVANVLAFLRKYDLDKYFKVNIEANHATLAFHDFQHELRYARINGVLG KPKEPTAHQYDTDAATTIAFLKQYGLDNHFKLNLEANHATLAGHTFEHELRMARVHGLLG KPKEPTKHQYDFDAATTIAFLETYGLKDHFKLNLEANHATLAGHTFEHELRVAALHDMLG KPQEPTKHQYDYDAATVYGFLKQFGLEKEIKLNIEANHATLAGHSFHHEIATAIALGLFG KPNEPRGDILLPTAGHAIAFVQELERPELFGINPETGHEQMSNLNFTQGIAQALWHKKLF KPNEPRGDILLPTVGHALAFIERLERPELYGVNPEVGHEQMAGLNFPHGIAQALWAGKLF KPNEPRGDIFLPTVGHGLAFIEQLEHGDIVGLNPETGHEQMAGLNFTHGIAQALWAEKL **.it i. i i i .. i . . i SIDANQGDLLLGWDTDQFPTNIYDTTLAMYEVIKAGG ----- FTKGGLNFDAKVRRASYK SIDANQGDLLLGWDTDQFPTNVYDTTLAMYEVIKAGG ----- FTKGGLNFDAKVRRASYK SIDANTGDMLLGWDTDQFPTDIRMTTLAMYEVIKMGG ----- FDKGGLNFDAKVRRASFE SVDANQGHPLLGWDTDEFPTDLYSTTLAMYEILQNGG ----- LGSGGLNFDAKVRRSSFE SIDANQGDLLLGWDTDEFPTDLYSAVLAMYEILKAGG ----- FKTGGINFDAKVRRPSFA SVDANRGDAQLGWDTDQFPNSVEENALVMYEILKAGG ----- FTTGGLNFDAKVRRQSTD HIDLNGQHGPKFDQDLVFGHGDLLNAFSLVDLLENGP-DGAPAYDGPRHFDYKPSR-TED HIDLNGQNGIKYDQDLRFGAGDLRAAFWLVDLLES ------ AGYSGPRHFDFKPPR-TED . * * . * HIDLNGQRGIKYDQDLVFGHGDLTSAFFTVDLLENGFPNGGPKYTGPRHFDYKPSR-TDG i .ti’ * i n 53 53 53 6O 51 53 19 19 19 113 113 113 119 111 113 66 66 66 173 173 173 179 171 173 122 122 122 233 233 233 239 231 233 182 182 182 293 293 293 299 291 293 242 242 242 348 348 348 354 346 348 300 295 301 '1‘ >m3wmmm'av-3ra H a n H‘m n S B >m>mww~3~3~a H S 0 Few n D 3 VEDLFIGHIAGMDTFALGFKIAYKLAKDGVFDKFIEEKYRSFKEGIGKEIVEGKTDFEKL 408 VEDLFIGHIAGMDTFALGFKVAYKLVKDGVLDKFIEEKYRSFREGIGRDIVEGKVDFEKL 408 PEDLFLGHIAGMDAFAKGFKVAYKLVKDRVFDKFIEERYASYKDGIGADIVSGKADFRSL 408 PDDLIYAHIAGMDAFARGLKVAHKLIEDRVFEDVIQHRYRSFTEGIGLEIIEGRANFHTL 414 DEDLFHAHIAGMDTYAVGLKVASRLLEDKALDQVIEERYESYTKGIGLEIKEGRTDLKKL 406 KYDLFYGHIGAMDTMALALKIAARMIEDGELDKRIAQRYSGWNSELGQQILKGQMSLADL 408 YDGVWESAKANIRMYLLLKERAKAFRADPEVQEALAASKVAELKTPTLNPGEGYAELLAD 360 FDGVWASAAGCMRNYLILKERAAAFRADPEVQEALRASRLDELARPTA-—ADGLQALLDD 353 YDGVWDSAKANMSMYLLLKERALAFRADPEVQEAMKTSGVFELGETTLNAGESAADLMND 361 ‘ : * .:. : EEYIIDKED-IELPSGK—QEYLESLLNSYIVKTIAELR —————— 444 EEYIIDKET—IELPSGK-QEYLESLINSYIVKTILELR ------ 444 EKYALERSQ-IVNKSGR—QELLESILNQYLFAE ----------- 439 EQYALNHKS—IKNESGR-QEKLKAILNQYILEV ----------- 445 AAYALENDH-IENQSGR-QERLKATVNRYLLNALREAPAGKETH 448 AKYAQEHHLSPVHQSGR-QEQLENLVNHYLFDK ----------- 440 RSAFEDYDADAVGAKGFGFVKLNQLAIEHLLGAR ---------- 394 RSAFEEFDVDAAAARGMAFERLDQLAMDHLLGARG --------- 388 SASFAGFDAEAAAERNFAFIRLNQLAIEHLLGSR ---------- 395 t . 10 CE Ti... 6 C .. - 5...: 62865 2m $233 .985 meson :32: 98 @053 2?: 3:03 SEES: .559? as 88593 05 53, new 32 ESE 53> masoEBE 83.60% .395. mo 26 268 05 mo 5335858 ocmfionom ”m 8sz a NMNA ll tetr; nati' don for c hu\e grout grou; (Whiz Valid, ilenkn of sub: Slums:- with Cc Other 1]; 101'. and f iiCllthe 150mm £184 will the ( Obsm'ed . is also me The recombinant TNXI existed as a dimer as well as a tetramer with the ratio of dimer to tetramer of approximately 20:1. Although the structural features of the recombinant and native forms of TNXI differ in the degree of subunit assembly, their functional prOperties do not differ. The dimer is a catalytically viable and stable form of the enzyme. Typically, two divalent cations (MgZT, Mn2+, or C0,“) per monomer are required for catalytic activity and stability of XIs. Two distinct metal binding sites, M1 and M2, have been identified in X13: (1) the metal in site M1 is coordinated to four carboxylate groups; (ii) the metal in site M2 is coordinated to one imidazole and three carboxylate groups. These were initially referred to as the structural and catalytic metals, respectively (Whitlow et al., 1991, and Marg and Clark, 1990). But these assumptions are no longer valid, since later studies showed that both metals are directly involved in catalysis (Jenkins et al., 1992 and Allen et al., 1994). Metal specificity depends on both the nature of substrate and on the enzyme type. Thermus aquaticus XI, a type I enzyme isomerizes glucose most efficiently in the presence of Mn2+, but its activity toward xylose is highest with Co2+ (Lehmacher and Bisswanger, 1990). The type II Bacillus coagulans XI, on the other hand, isomerizes xylose most effectively in the presence of Mn2+, while its activity toward fructose is highest with Co2+ (Marg and Clark, 1990). In our case, the three metals activate the TNXI. At any concentration, C02+ is the best activating metal. Glucose isomerase activity in the presence of Mg2+ is approximately 40 % of the activity observed with the Co2+ enzyme. Poorly active, the Mn2+ enzyme show only 16 % of the activities observed with the Coz+ enzyme (Vieille et al., 2001). The thermal stabilization of TNXI is also metal-specific: the Mn2+ enzyme is significantly more stable than the Co2+ and Mg2+ enzymes at 101°C. 12 r3} 198 enz3 The”! inter. in [hi intem' 3 met; al., 19‘ (Lee er for the molecule differenci Vt‘l'SUS X \‘1 THERMC lSOlaled ”Om lllllgllf’ “mull“ A mechanism of xylose isomerase catalysis was identified based on results of x- ray crystallographic studies on Arthrobacter or Streptomyces enzymes (Farber et al., 1989; Collyer et al., 1990) and biochemical properties exhibited by thermophilic enzymes obtained by site-directed mutagenesis of the xylA gene from Thermoanaerobacterium thermosulfurigenes (Lee et al., 1990). The enzymatic interconversion of aldose to ketose by xylose isomerases involves binding of the substrate in the ring form, substrate ring opening, isomerization of the linear intermediate, intermediate ring closure, and release of the product. The isomerization step proceeds by a metal ion-assisted hydride-shift mechanism (Figure 6) (Farber et al., 1989: Collyer et al., 1990; Lee et al., 1990), and this step, rather than ring opening, is rate determining (Lee et al., 1990). D-xylose and D-glucose have identical atomic configuration, except for the presence of an additional —CHzOH group at the C-6 position in the glucose molecule. This extra hydroxymethyl group must therefore be responsible for the differences in the catalytic efficiency exhibited by xylose isomerase toward glucose versus xylose. THERMOZYMES AND THEIR THERMAL STABILITY Thermozymes are enzymes that evolved in thermophiles (organisms thriving at 50—80°C) and hyperthermophiles (organisms thriving at 80°C or above). Most archea and some bacteria are thermophiles and hyperthermophiles. These organisms have been isolated from all types of terrestrial and marine environments. Thermozymes developed unique structure-function properties above 70°C (Vieille and Zeikus, 1996; Vieille and 13 *\ Figure rublgim COFFEspi D308 an OH OH H K183 OH OH H l K183 Figure 6: A metal—assisted hydride-shift mechanism of xylose in the Streptomyces rubiginosus xylose isomerase (type 1 X1). In type II XIs, Both Mn2+ are replaced by Co“. Corresponding residues of D257 and K183 in the Thermotoga neapolitana XI (TNXI) are D308 and K233, respectively. 14 Zhfll pol} e\ce inch and; therr 10pm: once Uiltl and t conce reactu SOlUbl: used n1 COUnter StllJCtUr they 5h. s.» C0Unt er? menh02\ Zeikus, 2001). They have already been used in molecular biology (e.g. Taq DNA polymerase), starch-processing (e.g. a-amylases, glucose isomerases) industry, and are excellent catalytic candidates for several additional applications that require high stability including organic syntheses, diagnostics, waste treatment, pulp and paper manufacture, and animal feed (Vieille et al., 1996). Intrinsically stable and active at high temperatures, thermozymes offer major biotechnological advantages over meSOphilic enzymes (optimally active at 25-50°C) or psychrophilic enzymes (optimally active at 5-25°C): (i) once expressed in mesophilic hosts, thermozymes are easier to purify by heat treatment, (ii) their thermostability is associated with a higher resistance to chemical denaturants, and (iii) performing enzymatic reactions at high temperatures allow higher substrate concentrations, lower viscosity, fewer risks of microbial contamination, and higher reaction rates. Protein stability has been actively studied for several decades starting with small, soluble, and monomeric enzymes (e.g. lysozyme and ribonuclease)(Dill, 1990). Having access to thermozymes allows us to determine what protein stabilization mechanisms are used in nature to gain extreme stability. Thermozymes are relatively similar to mesophilic enzymes: (1) their amino acid sequences are 40-85 % similar to those of their mesophilic counterparts (Vieille et al., 1995 and Burdette et al., 1996), (ii) their three-dimensional structures are superimposable (Davies et al., 1993 and Fujinaga et al., 1993), and (iii) they share the same catalytic mechanisms (Vieille et al., 1995 and Voorhorst et al., 1995). Therefore, their increased stability (as compared with their mesophilic counterparts) must be a result of differences in specific amino acid sequences. Since thermozymes are optimally active under more severely denaturing conditions than their 15 enZ}: efficr oxera unde- xleSc mesophilic counterparts, they need to be more rigid. This increased rigidity is essential for preserving their active center structures and protect them from unfolding. Such rigidity is demonstrated by lower hydrogen exchange rates and by lower susceptibility to proteolytic degradation and chemical denaturant or thermal unfolding (Veronese et al., 1984, Wrba et al., 1990, Kanaya and Itaya, 1992). Observations from large groups of enzymes indicate that stabilizing substitutions tend to improve the enzymes’ packing efficiency (through cavity filling and increase in core hydrophobicity), and to increase overall enzyme rigidity through helix stabilization, electrostatic interaction optimization and conformational strain reduction. Well-known and studied mechanisms of protein thermostabilization (Vieille and Zeikus, 1996; Vieille and Zeikus, 2001) are (i) hydrophobic interactions, (ii) packing efficiency and reduction in solvent-accessible hydrophobic surface, (iii) aromatic interactions, (iv) salt bridges, (v) disulfide bridges, (vi) hydrogen bonds, (vii) metal binding, (viii) reduction of conformational strain, (ix) reduction of the entropy of unfolding by proline substitution, (x) helix stabilization, (xi) stabilization of loops (xii) intersubunit interactions and oligomerization, (xiii) resistance to covalent destruction, and (xiv) post-translational modifications. GENETIC AND PROTEIN ENGINEERING OF Xls Thermostability mutations A cluster of aromatic amino acid residues is present in the active site pocket of xylose isomerases isolated from different sources and the hydrophobic interactions 16 (111101 mum t’l al. residt them; heat I mutan import subsUt by 43J protrud reducer tl1501105 increase TINSAJ fullCllOn Unchang Wildqypc A (199]) by residues_ 5 from Chem among these aromatic amino acids were postulated to be one of the important forces that maintains the monomers associated into active dimers (Collyer et al., 1990 and Whitlow et al., 1991). Meng et a1. (1993) explored the thermostability of mutants of aromatic residues (Trp48, PheS9, Trpl38, Phel44, and Trp187) in Thermoanaerobacterium thermosulfurigenes XI) with Co2+ as a co-factor. The Trp187His and Phel44Lys mutant enzymes were no longer resistant to the heat treatment at 75°C, which was one of the purification steps of the wild type and mutant enzymes. This indicates that the hydrophobic character of Trp187 and Phe144 is important in maintaining the enzyme structure at high temperature. The Trp138 substitutions with smaller hydrophobic residues had increased the thermostability at 85°C by 43-91 % without affecting the enzymes’ activities. Since the indole group of Trp138 protrudes into the active site cavity, replacement of Trp138 with Phe, Met, or Ala reduced the area of active site hydrophobic surface and therefore enhanced thermostability. Trp48Arg mutation did not change the activity of the enzyme but it increased the stability of the enzyme by 60 %. The enhancement of thermostability in the Trp48Arg mutant enzyme was brought about because Arg48 could presumably fulfill the function of hydrogen bonding to Asp338 and thus leave the active site structure unchanged. The Phe59His mutant was relatively less stable with shorter half-live than the wild-type enzyme. A different approach to improving thermostability was adopted by Quax et al. (1991) by converting lysine residues in Actinoplanes missouriensis XI (AMXI) to Arg residues. Special attention was given to preventing deleterious effects that could result from chemical modification of xylose isomerase by sugar components of high fructose l7 C011 AM inte inter inert impt‘ of th uhicl explu. MI li solutic isomer; Compdr. Actinop. thermW thennmL (ii A. Tc 1 ii) A the corn syrup, and in particular, from non-enzymatic glycation of key amino groups. Each AMXI subunit contains 20 lysine residues and two of them are located at the dimer/dimer interface, Lys252 and Ly5294 (Gly304 and Lys340 in TNXI, respectively). Lys252 is buried in the XI tetramer and is also involved in electrostatic interactions across the interdimer interface. The mutation Lys252Arg resulted in a 30 ‘70 increased in thermostabilty of the Mg-enzyme in solution with no loss in activity. More importantly, the half-life of the immobilized mutant at 70°C was 3-fold longer than that of the wild type. Lys252 would be accessible for glycation when the dimer dissociates, which would inactivate the enzyme by preventing reassociation of the dimer. This would explain why the immobilized mutant enzyme is more resistant to glucose inactivation. . Lys294 forms a salt bridge with the M1 ligand Asp256 (Asp308 in TNXI) and the M2 ligand Asp292 (Asp338 in TNXI). Lys294Arg had decreased thermostability in solution but very little after immobilization. Chang et al. (1999) studied the structures of highly thermostable type I xylose isomerases from Thermus thermophilus (TthXI) and Thermus caldophilus (TcaXI) compared to those of less thermostable XIs from Arthrobacter B3728 (AXI) and Actinoplanes missouriensis (AMXI). Analyses of various factors that may affect protein thermostability indicate that the possible structural determinants of the enhanced thermostability of TcaXI/TthXI over AXI/AMXI are (i) An increase in ion pair networks: the total ion pairs per tetramer of TcaXI/TthXI/AXI/AMXI are 150/ 1 63/93/ 1 07, respectively. (ii) A decrease in the large intersubunit cavities: there is a clear correlation between the decreased number of large internal cavities and thermostability. The total 18 lllll m) (iii) (iV) volume and surface area of cavities are 389/373/896/591 A3 and 938/887/ 1702/ 1 136 A2 for TcaXI/TthXI/AXI/AMXI, respectively. All cavities in TcaXI/TthXI tetramers are small, whereas some of the cavities in AXI/AMXI tetramers are very large. The volume and surface area of the largest cavities in TcaXI/TthXI/AXI/AMXI are 38/30/193/122 A3 and 74/64/304/207 A3. A removal of potential deamidation and isoaspartate formation sites: Deamidation and isoaspartate formation have been found to be one of the important processes leading to irreversible heat denaturation of protein at neutral pH (Ahern and Klibanov, 1985 and Aswad, 1990). They occur frequently when the sequences Asn-Gly, Asn-Ser, and Asp-Gly lie in highly flexible regions of the polypeptide chain. The analysis indicated that they are much fewer in TcaXI/TthXI (2/1) compared to AXI/AMXI (7/7). Shortened loops and Proline residues: TcaXI and TthXI have a significantly shortened loop due to deletions, compared with AXI and AMXI. Compared with AXI and AMXI, TcaXI and TthXI have approximately five more proline residues. A site-directed mutagenesis study of His219 residue of Streptomyces rubiginosus XI (His270 in TNXI) and its effect on activity and thermostability was done by Cha et al. (1994). This residue is conserved in all xylose isomerases. The three dimensional structure of the enzyme revealed that His219 is part of the octahedral coordination sphere of M2, One of the two metal ions (MnZT) in the active site. Substitutions of His219 with Ser, Glu, and Asn resulted in enzymes with the kcat values of only 0.3—0.5 % of that of the wild-type enzyme. The Km values of these mutant enzymes increased by 30-40 fold over 19 the \V actix ii enZ} it stable et‘t‘ecti interat betu e structt carbo “3511 Coord medst 1’6 Spry reflex enZ\j Cfih the wild-type value. The His219Lys mutant enzyme did not exhibit any measurable activity. Thermal denaturation studies indicated that the His219Ser and His 219Asn mutant enzymes are 5-8°C less stable, whereas Hi5219Glu and Hi5219Lys are 13-24°C less stable than the wild-type enzyme. In the His219Ser structure, a water molecule effectively replaced the Ne-2 atom of the imidazole ring of His219 and mediated the interaction between Mn2+ at the M2 site and Ser219. A similar water-mediated interaction between the metal ion and Asn219 was observed in the His219Asn mutant enzyme structure. On the other hand, no direct or water mediated interactions between the carboxyl group of Glu219 and the metal were observed. Whereas octahedral coordination was maintained for the metal at the M2 site in His219Ser and Hi5219Asn, a pentahedral coordination with the metal at the M2 site was observed in His219Glu. Metal activation measurements supported the observation that metal binding is perturbed and is responsible for thermal lability of His219 mutant enzymes. Active site mutations Most active site mutations reduce or destroy activity but some are also relevant to enzyme thermostability (Hartley et al., 2000). However, Meng et al. (1990) successfully used site-directed mutagenesis to switch the substrate preference of Themloanaerobacterium thermosulfurigenes XI (TTXI) from xylose to glucose by redesigning its substrate binding pocket. In the Arthrobacter enzyme, the structure of the enzyme complex with the six carbon competitive inhibitor, sorbitol, indicated that the C6-hydroxymethyl group of the substrate (glucose) is oriented toward the bottom of the 20 \kfll 'Ihrl sntul deem for x substrate-binding pocket and is adjacent (3.4-3.7 A) to the residues Met87, Thr89, and Va1134. These residues are highly conserved among X15 and correspond to Trpl38, Thr140, and Va1185, respectively, in the TTXI. Each of these residues was replaced with smaller amino acids to prove that they are part of the substrate-binding pocket in TTXI. Replacement of Trp138 with Phe, a smaller residue, produced an enzyme that had a higher catalytic efficiency (kw/Km) for glucose than the wild-type enzyme due to both a decrease in Km and an increase in km. On the other hand, this mutation increased the Km for xylose and decreased the catalytic efficiency for this substrate. This was suggested to be due to a more spacious pocket and this increased the freedom of movement of the xylose molecule, decreasing its binding efficiency. The enlargement of the binding pocket also decreased binding energy between the enzyme and the transition state resulting in the decrease of km. The Vall85Thr mutant enzyme had a slightly lower Km and a higher kcat for glucose. Substitution with a Ser residue, which has a smaller sidechain but otherwise is equivalent to Thr, did not improve the glucose catalytic efficiency. Likewise, replacement with Ala did not significantly change either Km or kw. Thus these data suggested that Va1185 does not hinder glucose binding, but the Thr substitution may provide an additional hydrogen bonding, presumably to the C6-OH group of glucose. Replacement of Thr140 with Ser increased the KM for both xylose and glucose and resulted in lower catalytic efficiency for glucose. It can be concluded that Thr140 hydrogen bonds to the substrate but does not strictly hinder the binding of glucose. The double mutant enzymes Trp138Phe/Vall85Thr and Trp138Phe/Vall85Ser had a higher catalytic efficiency for glucose than the wild-type enzyme of 5- and 2-fold, respectively. 21 'Ihc rest isor I)lF llbfar appro ”56d 5 [hmllg Vumu They also exhibited 1.5-and 3-fold higher catalytic efficiency for glucose than for xylose, respectively, thus they can be called the real “glucose isomerase” rather than xylose isomerase. DIRECTED EVOLUTION OF ENZYMES Directed evolution, also termed evolutionary engineering, has recently emerged as a key technology for biomolecular engineering and generating impressive results in the functional adaptation of enzymes to artificial, non-natural environments (Reetz and Jaeger, 1999; Pluckthun et al., 2000; and Wintrode and Arnold, 2000). The approach is highly attractive because its principles are simple and do not require detailed knowledge of structure, function, or mechanism. Essentially like natural (Darwinian) evolution, directed evolution comprises the iterative implementation of (1) the generation of a “library” of mutated genes, (2) its functional expression, and (3) a sensitive assay to identify individuals showing the desired properties, either by selection or by screening. After each round, the genes of improved variants are deciphered and subsequently serve as parents for another round of optimization (Brakmann, 2001). A series of experimental strategies have been developed for generating mutant libraries in the laboratory which differ in diversity. They can be divided into two main approaches, random mutagenesis and recombination. Random mutagenesis is a widely used strategy, which target whole genes. This may be achieved by passing cloned genes through mutator strains (Cox, 1976; and Greener et al., 1996), by treating DNA with various chemical mutagens (Shortle and Nathans, 1978; Kadonaja and Knowles, 1985; 22 and Dc' 19921. emerg- per nt alterat suhsu 11111111 and t perm lli'ldl‘ leId COW rec‘ emI Seq- Pare al.. and Deshler, 1992), or by error-prone PCR (Leung et al., 1989; and Cadwell and Joyce, 1992). Due to its simplicity and versatility, error-prone PCR (random PCR mutagenesis) emerged as the most common technique, which can result in mutation rate as high as 2 % per nucleotide position. The mutation rate may also be adjusted to lower values with alterations of PCR conditions. However, only a limited number of amino acid substitutions is accessible by this method since this reaction biases the distribution of mutation type in favor of transitions (A(—>G and T<——>C) over transversions (A/GHC/T), and because multiple substitutions within a single codon are extremely rare. Complete permutation of a single amino acid position (saturation mutagenesis) may enable the finding of non-conservative replacements that are inaccessible by random PCR mutagenesis (Miyazaki and Arnold, 1999). Recombination of DNA represents an alternative or additional approach for generating genetic diversity that is based on mixing and concatenation of genetic material from a number of parent sequences. Recombination may be advantageous in concentrating beneficial mutations which have arisen independently and may be additive, and likewise, in concentrating deleterious mutations which subsequently might be more efficiently purged from the population by selection (Zeyl and Bell, 1997; and Moore and Maranas, 2000). DNA shuffling was the first technique introduced for random in vitro recombination of gene variants created by random mutagenesis (Stemmer, 1994). It employs the PCR reassembly of whole genes from a pool of short overlapping DNA sequences (SO-300 bp) generated by random enzymatic fragmentation of different parental genes. Alternative protocols include staggered extension process (StEP) (Zhao et al., 1998) and random-priming recombination (Shao et al., 1998). StEP recombination is 23 a PC of p stru prir for €\lt 0f the CU ph bu EX a PCR reaction with very short annealing and extension steps that promote the formation of premature extension products. The truncated strands may anneal randomly to a parent strand, thus combining the different sequences from different parental strands. Random- priming recombination, as an alternative to DNA shuffling, produces random fragments for reassembly by annealing of short, random primers to a certain template gene and extension by a polymerase. In directed enzyme evolution, diversity is created on the DNA level, but selection or screening acts on the level of the encoded protein. Therefore, functional expression of the DNA libraries is a necessary prerequisite for the detection of improved enzyme variants. The most common approaches for recombinant protein expression employ the cellular transcription/translation machineries of well-established organisms such as E. coli, S. cerevisiae, or B. subtilis. Alternatively, a physical link between genotype and phenotype may be established by generating fusions between the protein of interest and a bacteriophage coat protein. Following intracellular assembly, recombinant phages express the protein variants on their surface while enclosing the appertaining genetic information within their genome (Johnson and Ge, 1999; and Smith and Petrenko, 1997). The most challenging step in directed evolution experiments is to develop a screening or selection scheme that is sensitive to the desired properties. Selection can be used either in vivo or in vitro. In vivo selection is most often achieved by genetic complementation of hosts that are deficient in a certain pathway or capacity. In vitro enrichment procedures that are detached from cell survival may also be termed selection. These techniques have been developed for the biopanning of phage-displayed peptide libraries by binding to a ligand that is immobilized on an appropriate column matrix 24 enr alt: &5\ Oil *3 E; mitt mo COT. 6E2 dire Oil: Tegl lFlt; (Brakmann, 2001). Recently, the approach has also been applied to the selective enrichment of phage-displayed functional enzyme libraries. Screening is an important alternative to selection. It enables a better control of the applied constraints, and also is more versatile, predominantly in unnatural environments, or with unnatural substrates. It usually requires that the mutant libraries are diluted and well-distributed which can be achieved by conventional plating of transfonnants on agar plates or filter membranes. This time-consuming step is sometimes accelerated using robotic systems. Common assays are based on visual or spectroscopic detection, for example formation, alteration, or destruction of colors or fluorescence characteristics. The determination of the optical parameters can also be accomplished by using automatic plate-readers, which enable a normalization of measured values to respective cell densities and may also be used to monitor the reaction kinetics. It should be noted that it is important to choose selective constraints that precisely reflect the desired property since the first law of directed enzyme evolution is “you get what you screen for” (Schmidt-Dannert and Arnold, 1999). During the past few years, many enzymes have been successfully improved by directed evolution (Table 1). The narrow range of substrates accepted by natural enzymes often prevents their use in new synthetic and commercial applications. Thus, by far most results were efficient tuning of catalytic efficiency toward non-natural substrates. Thermostability of enzymes and enantioselectivity of specific bioconversions have also been improved by these approaches. These examples showed that directed evolution is a powerful and reliable tool for improving biocatalysts in reasonably short periods of time. The only directed evolution approach (or random mutagenesis) performed on xylose isomerase reported to date was by Lonn et al. (2002). The thermophilic Thermus 25 N r. \\ th re 211‘: 21d ter thermophilus XI was subjected to one round of random PCR mutagenesis. It was screened for xylose isomerase activity at lower temperature than optimal by expression of the mutated genes in E. coli and replica—plated on McConkey agar plates, complemented with 1 % xylose followed by incubation at 30°C for 2 days. Three transformant colonies that were deeper red (suggesting higher XI activity) than the wild—type were selected and further characterized. Three amino acid substitutions were identified as Phe163Leu in domain I (C-terminal tail), and Glu372Gly/Val379Ala in domain 11 ([Ol/Bls barrel). These mutant enzymes showed improved catalytic rate constants (km) by up to nine times on both xylose and glucose with up to 26 times higher Km values on xylose but relatively unchanged for glucose. All enzyme variants’ relative activities on xylose are higher than the wild-type at low temperatures. These results suggested that amino acid substitutions distant from the catalytic center could lead to cold adaptation. There is a close relationship between molecular flexibility and function. Thermophilic enzymes are rigid and require high temperatures in order to gain sufficient molecular flexibility for activity. Their molecular structure must therefore be balanced between the requirements for stability and dynamics. They suggested that the sequence changes underlying the adaptation of T. thermophilus XI variants to temperatures lower than their optimal temperature, allow a higher degree of flexibility in areas that move during catalysis. Kinetic analysis demonstrated that the increase in the relative activity in the enzyme variants for xylose at low temperatures was indeed caused by an increase in kcal and not by a decrease in the Km value. This suggests that the mutant enzymes did not acquire higher affinity for the substrate than the wild-type enzyme at lower temperatures. 26 1n and with Ilienm in )2 directed m COUlp‘dl’lWl Tllr’mluum. themiostah glucose at biochemica. commercial. industrial up In this study, the molecular determinants that are responsible for thermostability and activity toward glucose of a xylose isomerase from a hyperthermophilic eubacterium Thermotoga neapolitana (TNXI) will be identified. The method used in such work is site- directed mutagenesis based on the amino acid sequence and three dimensional structure comparisons between TNXI and a xylose isomerase from a thermophilic eubacterium Thermoanaerobacterium thermosulfurigenes (TTXI). Furthermore, the highly thermostable TNXI will be engineered by directed evolution to improve its activity on glucose at low temperature and pH. Finally, the resulting laboratory-evolved enzyme’s biochemical properties and fructose productivity will be compared to those of the commercially available glucose isomerase, GensweetTM, to further ascertain its utility in industrial applications. 27 Table 1: Examples of enzymes that were successfully optimized using directed evolution. activity in organic solvent + screening Target enzyme Target property Change evolved Approach Reference subtilisin E activity in organic ~ 170-fold increase error-prone PCR + Chen & solvents in 60% screening Arnold dimethylformamide (1993); Arnold & Chen (1994) B-lactamase activity towards 32,000-fold greater DNA shuffling + Stemmer new substrate resistance to selection (1994) cefotaxime para-nitrobenzyl activity towards 60- 150 fold increase error-prone PCR Moore & esterase pNB esters; and DNA shuffling Arnold (1996): Moore et al. (1997); Arnold & Moore (1998) B-galactosidase activity towards 66-fold increased DNA shuffling + Zhang et al. new substrate; activity; 1000-fold screening (1997) substrate increase in substrate specificity specificity aminoacyl-tRNA aminoacylation of 55-fold increase in DNA shuffling + Liu et al. synthetase a modified tRN A activity selection (1997) aspartate activity towards B 105 increase DNA shuffling + Yano et al. aminotransferase branched amino selection (1997) and 2-oxo acids lipase enantioselectivity increase in error-prone PCR + Reetz et al. in hydrolysis of p- enantiomeric excess screening ( 1997) nitrophenyl 2- from 2% to 81% methyldecanoate pNB esterase thermostability 14 0C increase in Tm error-prone PCR, Giver er al. + increased activity DNA shuffling + (1998) at all temperatures screening subtilisin E thermostability 17 °C increase in Tm error-prone PCR, Zhao & + increased activity DNA shuffling + Arnold (1999) at all temperatures screening subtilisin BPN’ activity at 10°C 2-fold increase chemical Taguchi et al. mutagenesis + (1998) screening cephalosporinases activity towards 270-540-fold DNA shuffling of Crameri et al. moxalactam increased resistance homologous genes (1998) + selection kanamycin thermostability increase 20°C DNA shuffling + Hoseki nucleotidyl screening/selection et al. (1999) 28 .1wl w. l_.l _ll transferase B. remove fructose active without FBP random Allen & stearothermophilus 1,6 bisphosphate mutagenesis + Holbrook LDH requirement (FBP) screening (2000) phospholipase A, thermostability increase Tm by 1 1°C error prone PCR + Song & Rhee without screening (2000) compromising activity TEM-l [3- activity towards 20,000 fold increase high frequency Zaccolo et a1. lactamase cefotaxime random ( 1999) mutagenesis myoglobin peroxidase activity 25-fold increase error-prone PCR + Wan et al. screening (1998) hydantoinase enantioselectivity inverted error-prone PCR + May et al. + total activity enantioselectivity, 3 screening (2000) x increase in total activity xylose isomerase activity at low improved kcan at low error prone PCR + Lonn er temperature temperatures screening a1.(2002) 29 10. ll. 12. 13. 14. REFERENCES . Allen, K. N ., Lavie, A., Glasfeld, A., Tanada, T. N., Gerrity, D. P., Carlson, S. C., Farber, G. K., Petsko, G. A., and Ringe, D. (1994) Isotopic exchange plus substrate and inhibitor kinetics of D-xylose isomerase do not support a proton-transfer mechanism. Biochemistry 33, 1481-1487. Allen, S. J., and Holbrook, J. J. (2000) Production of an activated form of Bacillus stearothermophilus L-2-hydroxyacid dehydrogenase by directed evolution. Protein Eng. 13, 5-7. Arnold, F. H., and Chen, K. (1994) US patent 5,316,935. Arnold, F. H., and Moore, J. C. (1998-1999) US patents 5,741,691 and 5,906,930. Barker, S. A. (1976) Pure fructose syrups. Process Biochem. 11, 20-25. Bhosale, S. H., Rao, M. B., and Deshpande, V. V. (1996) Molecular and industrial aspects of glucose isomerase. Microbial. Reviews 60, 280- 300. Brakmann, S. (2001) Discovery of superior enzymes by directed molecular evolution. Chembiochem. 2, 865-871. Burdett, D. S., Vieille, C., and Zeikus, J .G. (1996) Cloning and expression of the gene encoding the Thermoanaerobactor ethanolicus 39E secondary-alcohol dehydrogenase and biochemical characterization of the enzyme. Biochem. J. 316, 115-122. Cadwell, R. C., and Joyce, G. F. (1992) Randomization of genes by PCR mutagenesis. PCR Methods Appl. 2, 28-33. Cha, J., Cho, Y., Whitaker, R. D., Carrell, H. L., Glusker, J. P., Karplus, P. A., and Batt, C. A. (1994) Perturbing the metal site in D-xylose isomerase: effect of mutations of His-220 on enzyme stability. J. Biol. Chem. 269, 2687-2694. Chang, C., Park, B. C., Lee, D. -S., and Suh, S. W. (1999) NCBI Protein Database CAB 4930288. Chen, K., and Arnold, F. H. (1993) Tuning the activity of an enzyme for unusual environments: sequential random mutagenesis of subtilisin E for catalysis in dimethylforrnamide. Proc. Natl. Acad. Sci. USA 90, 5618-5622. Collyer, C. A., Henrick, K., and Blow, D.M. (1990) Mechanism for aldose-ketose interconversion by D-xylose isomerase involving ring opening followed by a 1,2- hydride shift. J. Mol. Biol. 212, 21 1-235. Cox, E. C. (1976) Bacterial mutator genes and the control of spontaneous mutation 3O 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. Ann. Rev. Genet. 10, 135-156. Crameri, A., Raillard, S. A., Berrnudez, E., and Stemmer, W. P. C. (1998) DNA shuffling of a family of genes from diverse species accelerates directed evolution. Nature 391, 288-291. Davies, G. J., Gamblin, S. J ., Littlechild, J. A., and Watson, H. C. (1993) The structure of a thermally stable 3-phosphoglycerate kinase and a comparison with its mesophilic equivalent. Protein Struct. F unct. Genet. 15, 283-289. de Raadt, A., Ebner, M., Ekhart, C. W., Fechter, M., Lechner, A., Stobl, M., and Stutz, A. E. (1994). Glucose isomerase (E.C.5.3. 1.5) as a reagent in carbohydrate synthesis: success and failures with the isomerisation of non-natural derivatives of D- glucose into the corresponding 2-ketoses. Catalysis Today 22, 549-561. Deshler, J. O. (1992) A simple method for randomly mutating cloned DNA fragments by using chemical mutagens and the polymerase chain reaction. Gen. Anal. Techn. Appl. 9, 103-106. D111, K. A. (1990) Dominant forces in protein folding. Biochemistry 29, 7133-7155. Farber, G. K., Glasfeld, A., Tiraby, G., Ringe, D., and Petsko, G. A. (1989) Crystallographic studies of the mechanism of xylose isomerase. Biochemistry 28, 7289-7297. Fujinaga, M., Berthet,-Colominas, C., Yaremchuk, A. D., Tukalo, M. A., and Cusak, S. (1993) Refined cr stal structure of the seryl-tRNA synthetase from Thermus thermophilus at 2.5 resolution. J. Mol. Biol. 234, 222-233. Ghatge, M. S., Phadatare, S. U., Bodhe, A. M., Deshpande, V. V. (1994) Unfolding and refolding of glucose/xylose isomerase from Streptomyces sp. NCIM 2730. Enzyme Microb. Technol. 16, 323-327. Giver, L., Gerhenson, A., Freskgard, P. O., and Arnold, F. H. (1998) Directed evolution of a thermostable esterase. Proc. Natl. Acad. Sci. USA 95, 12809-12813. Greener, A., Callaghan, M., and Jerpseth, B. (1996) An efficient random mutagenesis technique using an E. coli mutator strain. Methods Mol. Biol. 57, 375-385. Hartley, B. S., Hanlon, N., Jackson, R. J., and Rangarajan, M. (2000) Glucose isomerase: insights into protein engineering for increased thermostability. Biochim. Biophys. Acta. 1543, 294-335. Hess, M. J., Tchemajenko, V., Vieille, C., Zeikus, J. G., and Kelly, R. M. (1998) Thermotoga neapolitana homotetrameric xylose isomerase is expressed as a catalytically active and thermostable dimer in Escherichia coli. Appl. Environ. 31 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. Microbiol. 64, 2357-2360. Hoseki, J., Yano, T., Koyama, Y., Kuramitsu, S., and Kagamiyama, H. (1999) Directed evolution of thermostable kanamycin-resistance gene: a convenient selection marker for Thermus thermaphilus. J. Biachem.(Takya) 126, 951-956. Jenkins, J., Janin, J., Rey, F., Chiadmi, M., Van Tilbeurgh, H., Lasters, I., DeMaeyer, M., VanBelle, D., Wodak, S. J ., Lauwereys, M., Stanssens, P., Mrabet, N. T., Snauwaert, J ., Matthyssens, G., and Lambeir, G. A. -M. (1992) Protein engineering of xylose (glucose) isomerase from Actinaplanes missouriensis .1. Crystallography and site directed mutagenesis of metal binding sites. Biochemistry 31, 5449-5458. Johnson, K., Ge, L. (1999) Phage display of combinatorial peptide and protein libraries and their application in biology and chemistry. Curr. Top. Microbial. Immunol. 243, 87-105. Kadonja, J. T., and Knowles, J. R. (1985) Nucleic Acids Res. 13, 1733-1745. Kanaya, S., and Itaya, M. (1992) Expression, purification, and characterization of a recombinant ribonuclease H from Thermus thermaphilus HB8. J. Biol. Chem. 267, 10184-10192. Lehmacher, A., and Bisswanger, H. (1990) Comparative kinetics of D-xylose and D- glucose isomerase activities of the D-xylose isomerase from Thermus aquaticus H88. Biol. Chem. 371, 527-536. Lee, C., Bagdasarian, M., Meng, M., and Zeikus, J. G. ( 1990a) Catalytic mechanism of xylose (glucose) isomerase from Clastridium thet'mosulfurafenes. J. Biol. Chem. 265, 19082-19090. Leung, D. W., Chen, E., and Goeddel, D. V. (1989) Systems for heterologous gene expression. Technique 1, 11-15. Liu, D. R., Magliery, T. J., Pastemk, M., and Schultz, P. G. (1997) Engineering a tRNA and aminoacyl-tRNA synthetase for the site-specific incorporation of unnatural amino acids into proteins in vivo. Proc. Natl. Acad. Sci. USA 94, 10092-10097. Lonn, A., Gardonyi, M., van Zyl, W., Hahn-Hagerdal, B., and Otero, R. C. (2002) Cold adaptation of xylose isomerase from Thermus thermaphilus through random PCR mutagenesis. Eur. J. Biochem. 269, 157-163. Marg, G. A., and Clark, D. S. (1990) Activation of the glucose isomerase by divalent cations: evidence for two distinct binding sites. Enzyme Microb. Technol. 12, 367- 373. May, 0., Nguyen, P., and Arnold, F. H. (2000) Inverting enantioselectivity by 32 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. directed evolution of hydantoinase for improved production of L-methionine. Nat. Biotechnol. 18, 317-320. Meng, M., Lee, C., Bagdasarian, M., and Zeikus, J. G. (1991) Switching substrate preference of thermophilic xylose isomerase from D-xylose to D-glucose by redesigning the substrate binding pocket. Proc. Natl. Acad. Sci. USA 88, 4015-4019. Meng, M., Bagdasarian, M., and Zeikus, J. G. (1993) Thermal stabilization of xylose isomerase from Thermoanaerabacterium thermosulfurigenes. Proc. Natl. Acad. Sci. USA 90, 8459-8463. Miyazaki, K., and Arnold, F. H. ( 1999) Exploring nonnatural evolutionary pathways by saturation mutagenesis: rapid improvement of protein function. J. Mal. Eval. 49, 716-720. Moore, G. L., and Maranas, C. D. (2000) Modeling DNA mutation and recombination for directed evolution experiments. J. Theor. Biol. 205, 483-503. Moore, J. C., and Arnold, F. H. (1996) Directed evolution of a para—nitrobenzyl esterase for aqueous-organic solvents. Nat. Biotechnol. 14, 458-467. Moore, J. C., Jin, H. M., Kuchner, O., and Arnold, F. H. (1997) Strategy for the in vitro evolution of protein function: enzyme evolution by random recombination of improved sequences. J. Mal. Biol. 272, 336-347. Pluckthun, A., Schaffitzel, C., Hanes, J ., and Jermutus, L. (2000) In vitro selection and evolution of proteins. Adv. Prat. Chem. 55, 367-403. Quax, W. J., Mrabet, N. T., Luiten, R. G. M., Schuurheizen, P. W., Stanssens, P., and Lasters, I. (1991) Enhancing the thermostability of glucose isomerase by protein engineering. Biotechnology 9, 738-742. Reetz, M. T., Zonta, A., Schimossek, K., Liebeton, K., and Jaeger, K. E. (1997) Directed evolution of an enantioselective lipase. Angew. Chem. Int. Ed. Eng]. 36, 2830-2832. Reetz, M. T., and Jaeger, K. E. (1999) Directed evolution of enantioselective enzymes for organic chemistry. Top. Curr. Chem. 200, 32-57. Schmidt-Dannert, C., and Arnold, F. H. (1999) Directed evolution of industrial enzymes. Trends Biotechnol. 17, 135-136. Shao, Z., Zhao, H., Giver, L., and Arnold, F. H. (1998) Random priming in vitro recombination: an effective tool for directed evolution. Nucleic Acids Res., 26, 681- 683. 33 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. Shortle, D., and Nathans, D. (1978) Local mutagenesis: a method for generating viral mutants with base substitutions in preselected regions of the viral genome. Proc. Natl. Acad. Sci USA 75, 2170-2174. Smith, G. P., and Petrenko, V. A. (1997) Phage display. Chem. Rev. 97, 391-410. Song, I. K., and Rhee, J. S. (2000) Simultaneous enhancement of thermostability and catalytic activity of phospholipase A(l) by evolutionary molecular engineering. Appl. Environ. Microbiol. 66, 890-894. Stemmer, W. (1994) Rapid evolution of a protein in vitro by DNA shuffling Proc. Natl. Acad. Sci. USA, 91, 10747-10751. Taguchi, S., Ozaki, A., and Momose, H. (1998) Engineering of a cold-adapted protease by sequential random mutagenesis and screening system. Appl. Environ. Microbiol. 64, 492-495. Vangrysperre, W., Callens, M., Kersters-Hilderson, H., and De Bruyne, C. K. ( 1988) Evidence for an essential histidine residue in D-xylose isomerases. Biochem. J. 250, 153-160. Veronese, F. M., Boccu, E., Schiavon, O., Grandi, C., and Fontana, A. (1984) General stability of thermophilic enzymes: studies on 6-phosphogluconate dehydrogenase from Bacillus stearathermophilus and yeast. J. Appl. Biochem. 6, 39-47. Vieille, C., Hess, M., Kelly, R. M., and Zeikus, J. G. (1995) xylA cloning and sequencing and biochemical characterization of xylose isomerase from Thermataga neapolitana. Appl. Environ. Microbiol, 61, 1867—1875. Vieille, C., Burdette, D. S., and Zeikus, J. G. ( 1996) Thermozymes. Biatechnol. Annu. Rev. 2, 1-83. Vieille, C. and Zeikus, J. G. (1996) Thermozymes: identifying molecular determinants of protein structural and functional stability. Trends Biotechnol. 14, 183-190. Vieille, C., and Zeikus, J. G. (2001) Hyperthermophilic enzymes: sources, uses, and molecular mechanisms for thermostability. Microbial. Mol. Biol. Rev. 65, 1-43. Vieille, C., Sriprapundh, 8., Kelly, R. M., and Zeikus, J. G. (2001) Xylose isomerases from Thermataga. Methods Enzymal. 330, 245-249. Voorhorst, W. G. B., Eggen, R. I. L., Luesink, E. J., and de Vos, W. M. (1995) Characterization of the celB gene coding for B-glucosidase from the hyperthermophilic archaeon Pyracaccusfuriasus and its expression and site—directed mutation in Escherichia cali. J. Bacterial. 177, 7105-71 1 1. 34 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. Wan, L., Twitchett., M. B., Eltis, L. D., Mauk, A. G., and Smith, M. (1998) In vitro evolution of horse heart myoglobin to increase peroxidase activity. Proc. Natl. Acad. Sci. USA 95, 12825-12831. Whitlow, M., Howard, A. J., Finzel, B. C., Poulos, T. L., Winbome, E., and Gilliland, G. L. ( 1991) A metal-mediated hydride shift mechanism for xylose isomerase based on the 1.6 A Streptomyces rubiginasus structures with xylitol and D-xylose. Proteins 9, 153—173. Wintrode, P. L., Miyazaki, K., and Arnold, F. H. (2000) Cold adaptation of a mesophilic subtilisin-like protease by laboratory evolution. J. Biol. Chem. 275, 31635-31640. Wrba, A., Schweiger, A., Schultes, V., Jaenicke, R., and Zavodsky, P. (1990) Extremely thermostable D-glyceraldehyde-3-phosphate dehydrogenase from the Eubacterium Thematoga maritima. Biochemistry 29, 7584-7592. Yano, T., Oue, S., and Kagamiyama, H. (1998) Directed evolution of an aspartate aminotransferase with new substrate specificities. Proc. Natl. Acad. Sci. USA 95, . 5511-5515. Zaccolo, M., and Gherardi, E. (1999) The effect of high-frequency random mutagenesis on in vitro protein evolution: a study on TEM-l beta-lactamase. J. Mal. Biol. 285, 775-783. Zeyl, C., and Bell, G. (1997) The advantage of sex in evolving yeast populations. Nature 388, 465-468. Zhang, J-H., Dawes, G., and Stemmer, W. P. C. (1997) Directed evolution of a fucosidase from a galactosidase by DNA shuffling and screening. Proc. Natl. Acad. Sci. USA 94, 4504-4509. Zhao, H., Giver, L., Shao, Z., Affholter, J ., and Arnold, F. H. (1998) Molecular evolution by staggered extension process (StEP) in vitro recombination. Nat. Biotechnol., 16, 258-261. Zhao, H., and Arnold, F. H. (1999) Directed evolution converts subtilisin E into a functional equivalent of therrnitase. Protein Eng. 12, 47-53. 35 CHAPTER II ROLE OF PROLINE IN XYLOSE ISOMERASE THERMAL STABILITY: COMPARATIVE THERMOSTABILITY OF A THERMOPHILIC Thermaanaerobacterium thermasulfurigenes XYLOSE ISOMERASE (TTXI) AND A HYPERTHERMOPHILIC Thermatoga neapolitana XYLOSE ISOMERASE (TNXI) 36 ABSTRACT Xylose isomerases (Xls) from Thermaanaerabacterium thermasulfurigenes (TTXI) and Thermataga neapolitana (TNXI) are 70.4% identical in their amino acid sequences and have a nearly superimposable crystal structure. Nonetheless, TNXI is much more thermostable than TTXI. Except for a few additional prolines and fewer Asn+Gln in TNXI, no other obvious differences in the enzyme structures can explain the differences in their stability. TNXI has 2 additional prolines in the Phe59 loop (Pr058 and Pro62). Mutations Gln58Pro, Ala62Pro, and Gln58Pro/Ala62Pro in TTXI and their reverse counterpart mutations in TNXI were constructed by site-directed mutagenesis. Surprisingly, only the Gln58Pro mutation enhanced thermostability of TTXI (43 % longer half-life at 85°C), whereas, the Ala62Pro and Gln58Pro/Ala62Pro mutations both dramatically lowered the TTXI’s stability. Analysis of the three-dimensional (3D) structure of TTXI and the predicted structure of its Ala62Pro mutant derivative indicated that a steric hindrance between Pro62-C8 and Lys6l-CB (2.92 A) is responsible for the reduced thermostability of the mutant enzyme compared to the native TTXI. All the reverse counterpart mutations destabilized TNXI thus confirming that these 2 prolines play important roles in TNXI’s thermostability. 37 INTRODUCTION Xylose isomerase (D-xylose ketol isomerase; EC 5.3.1.5) (X1) is an intracellular enzyme found in a number of bacteria that utilize xylose as carbon substrate for growth (Chen, 1980). XI converts D-xylose to D-xylulose in viva and also catalyzes the conversion of D-glucose to D-fructose in vitro (Takasaki et al., 1969). This latter activity is used in industry for the production of high fructose corn syrup (HFCS) and xylose (i.e., glucose) isomerase is one of the largest volume commercial enzymes used today (Lee and Zeikus, 1991). Therrnostable XIs with neutral or slightly acidic pH optima have a potential for industrial applications with the advantages of faster reaction rates, higher fructose concentrations at equilibrium, decreased viscosity of substrate and product streams, and less problems of by-products formation (Lee and Zeikus, 1991). Two XI groups have been identified; type I enzymes are shorter than type II enzymes by about 50 amino acids at their N-terminus (Vangrysperre et al., 1990). Two type II XIs have been studied extensively in our laboratory: one from a thermophile, Thennoanaerabacterium thermasulfurigenes (TTXI) and the other from a hyperthermophile, Thermataga neapolitana (TNXI). The genes encoding these enzymes (xylA genes) were cloned, sequenced, and expressed in Escherichia cali (Lee et al., 1990a, 1990b, Meng et al., 1991, and Vieille et al., 1995). TNXI has a higher turnover number, a lower Km for glucose, and is more thermostable than any other known type II xylose isomerases (Vieille et al., 1995). XIs are significantly stabilized and activated in the presence of divalent cations, especially by Mg“, Co“, and Mn”. Although TNXI and TTXI are highly similar (70.4% amino acid sequence identity), TNXI is significantly 38 “1016 ll for II Tle‘1 the dif in TX. al.. 1‘: lExllli enho] residt does inter. three and muta addn unto lodri 91581 ( Allen Serif) stahih.‘ compo! more thermostable than TTXI with an optimum temperature for activity of 95°C (85°C for TTXI) and melting temperature in the presence of 0.5 mM Co2+ of 126°C (82°C for TTXI) (Vieille et al., 1995). No obvious differences in the enzyme structures can explain the differences in their stabilities except for a few additional prolines and fewer Asn+Gln in TNXI (as seen from its alignment with TTXI). Proteins can be stabilized by decreasing their entropy of unfolding (Matthews et al., 1987). Prolines, with their pyrrolidine ring, can only adopt a few configurations. They restrict the configurations allowed for the preceding residue, and they can decrease the entropy of a protein’s unfolded state. Substituting Pro for another carefully chosen residue can thus increase protein stability, provided that the newly introduced proline does not create volume interferences, and does not destroy stabilizing non-covalent interactions. The effect of prolines on protein stabilization has been studied by site- directed mutagenesis. The two T4 lysozyme mutants Ala82Pro (Matthews et al., 1987) and Ile3Pro (Dixon et al., 1992) illustrate the importance of carefully selecting the mutation location. With mutation Ala82Pro, the two conditions listed above were addressed, and the mutation stabilized the protein mainly by decreasing its entropy of unfolding. In mutant Ile3Pro, the substitution eliminated a hydrogen bond and the hydrophobic interactions created by lie. The degree of enthalpic destabilization was greater than the entropy gained resulting in destabilization of the mutant enzyme. In Allen et al.’s study of Aspergillus awamari glucoamylase, three proline mutants: Ser30Pro, Asp345Pro, and G1u408Pro were constructed. The Ser3OPro mutation Stabilized the protein because the mutation site allowed a residue conformation compatible with a proline, and because residue 29 (a valine) could adopt one of the 39 conlor Gil-11.1 conio stahil Ptolll proli conformations allowed for the residue preceding a proline (Matthews et a1, 1987). The G1u408Pro mutation was destabilizing because the mutation site did not allow any conformation required for proline, whereas the Asp345Pro mutation did not affect stability due to the fact that the substitution destroyed the Ol-helix dipole in that region. Prolines 58 and 62 in TNXI are present in a large loop in which 14 hydrophilic residues surround Phe59. The corresponding TTXI residues are Gln58 and Ala62. The Phe59 loop participates in building the neighboring subunit’s active site (Farber et al., 1989: Whitlow etaL,l99l) In the present report we test the hypothesis that substituting Gln58 and Ala62 with prolines will stabilize TTXI, and that the reverse mutations will destabilize TNXI. 40 MATERIALS AND METHODS Bacterial strains and chemicals: E. cali strain BL21 (DE3) (Novagen, Madison, WI.) was used to overexpress the recombinant T. thermasulfurigenes and T. neapolitana xylA genes cloned in the pET23a and pET22b+ vectors, respectively (Novagen). Media and growth conditions were the same as described by Vieille et al. (1995). Medium components and all other chemicals were reagent grade. Site-directed mutagenesis and other DNA techniques: A11 DNA manipulations were performed using established protocols (Sambrook et al., 1989; Ausubel et al., 1993). Point mutations were introduced into the T. neapolitana and T. thermasulfurigenes xylA genes using the QuikChangeTM site-directed mutagenesis kit (Stratagene, La Jolla, CA). Mutagenic Oligonucleotides (Table 1) were synthesized by the Macromolecular Structure Facility, Department of Biochemistry, Michigan State University. Mutations were verified by DNA sequencing using the Thermosequenase sequencing kit (United States Biochemical, Cleveland, OH). Protein purification: Recombinant enzymes were purified using the procedure of Vieille et al. (1995) followed by two additional steps. Partially purified enzymes were applied to a DEAE-Sepharose column (2.5x15 cm) equilibrated with 50 mM MOPS (pH 7.0) containing 5 mM MgSO., and 0.5 mM CoClz (buffer A), and enzymes were eluted using 500 m1 of 0—250 mM NaCl gradient in buffer A. The pooled fractions from the DEAE-Sepharose 41 Table 1: Oligonucleotides and DNA templates used for site-directed mutagenesis. Bold and underlined nucleotides are the mutation sites. DNA templates Gln58Pro/Ala62Pro 5’-GGAACAGATCCAI I IGGCAAACC-3’ 3’-CCTTGTCTAGGTAAACCGI l lGG—S’ Mutations Oligonucleotides TTXI: 5’-GGAACAGATCCATTTGGCAAAGC-3’ 01°58!“ 3’-CCTI‘GTCTAGGTAAACCGTITCG-5’ “XI g°°° Ala 62Pro 5’-GATCAAT'ITGGCAAACCTACCATGC-3’ TTXI we 3’-CTAGTTAAACCG’I'ITGGATGGTACG-5’ g TTXI (Ala62Pro) gene TNXI: ProS8Gln Pro62Ala ProSSGln/Pro62Ala 5’ -GGGAAGGGATCAGTTCGGAGACCC-3’ 3’-CCCTTCCCTAGTCAAGCCTCTGGG-5’ 5’-CCTTCGGAGACGCAACGGCCGATC-3’ 3’ -GGAAGCCTCTGCGTTGCCGGCTAG-S’ 5’-AGTTCGGAGACGCAACGGCCGATC-3’ 3’ -TCAAGCCTCTGCGTTGCCGGCTAG-5’ TNXI gene TNXI gene TNXI (Pr058Gln) gene 42 column were concentrated in a stirred ultrafiltration cell (MW cut-off 30 kDa)(Amicon, Beverly, MA), dialyzed twice against buffer A, applied to a Polybuffer column (Pharrnacia, Uppsala, Sweden) equilibrated with 25mM histidine-HCl (pH 6.2), and eluted using a pH (6.0 to 4.0) gradient according to the manufacturer’s instructions. The active fractions were pooled, concentrated in a stirred ultrafiltration cell (Amicon), and dialyzed twice against buffer A. Concentrated and homogenous enzymes were dispensed and stored frozen at -70°C. Xylose isomerase assays: XI activity was routinely assayed with glucose as the substrate. The enzyme (0.06 mg/ml) was incubated in 50 mM MOPS (pH 7.0 at room temperature) containing 1 mM CoClz and 1 M glucose at 60°C for 20 min. The reaction was stopped by cooling the tubes in ice. The amount of fructose produced was determined by the cysteine-carbazole-sulfuric acid method (Dische and Borenfreund, 1951). To determine the effect of temperature on XI activity, the assay mixtures were incubated at the temperatures of interest in a Perkin-Elmer Cetus GeneAmp PCR system 9600 (Perkin Elmer, Norwalk, CT) for 20 min. To determine the kinetic parameters, assays were performed in the presence of either 80 to 1,400 mM glucose or 20 to 900 mM xylose. The amounts of fructose and xylulose produced were determined as above. Absorbances were measured at 537 nm and 560 nm for xylulose and fructose, respectively. One unit of isomerase activity is defined as the amount of enzyme that produced 1 umole of product per min under the assay conditions. 43 Thermostability assays: The time course of irreversible thermoinactivation was measured by incubating the enzyme (0.5-5 mg/ml) in 10 mM MOPS buffer (pH 7.0) containing 50 11M CoClg (buffer B) at 85°C (TTXI derivatives) or 95°C (TNXI derivatives) in a Perkin-Elmer Cetus GeneAmp PCR system 9600 for various amounts of time, and by determining the residual glucose isomerase activity at 65°C (TTXI derivatives) or 80°C (TNXI derivatives). The first order rate constant, k, of irreversible thermoinactivation, was obtained by linear regression in semi-log coordinates of residual activity. Enzyme half-life was calculated from the equation: tum = ln 2/k. Heat-induced enzyme precipitation: Heat-induced enzyme precipitation was monitored from 25°C to 100°C by light scattering (71. = 580 nm) using protein solutions (0.2 mg/ml) in buffer B. Absorbance measurements were conducted in 0.3 m1 quartz cuvettes (path length = 1.0 cm), using a Gilford Response spectrophotometer (Corning, Oberlin, OH) equipped with a Peltier cuvette heating system. The increasing thermal gradient was 10°C min]. The temperature of 50% precipitation was the temperature at which the OD at 580 nm equals half of the difference between the baseline and the maximum ODs. Analysis of TTXI and TNXI three-dimensional (3D) structures: Enzymes were visualized on an IRIS-4D25 computer (Silicon Graphics Computer System, Mountain View, CA) using the INSIGHT II graphic program (Biosym Technologies, San Diego. CA). Proteins Data Bank (PDB) files (#lAOC for TTXI and #1AOE for TNXI) were obtained from the Protein Data Bank website (www.rcsb.org/pdb). 44 RESULTS TNXI Pr058 and Pro62 are present in a large loop in which fourteen hydrophilic residues surround Phe59 (Figure 1). According to crystallographic data, Phe59 (Phe26 in Actinoplanes XI) participates in the architecture of the neighboring subunit’s active site (Farber et al., 1989; Whitlow et al., 1991). The corresponding residues in TTXI, Gln58 and Ala62, have backbone dihedral angles ([-66.57, -l4.49] and [-57.91, 138.87], respectively) allowed for prolines (Nicholson et al., 1988). With backbone dihedral angles of (-139.73, -175. 10) and (-93.32, 176.20), respectively, TTXI residues Asp57 and Lys61 are in extended conformations, the most common conformation for residues preceding prolines (Nicholson et al., 1988). This structural information suggests that mutations Gln58Pro and Ala62Pro would not create unfavorable backbone conformations and that they could stabilize TTXI. Mutation Gln58Pro did not affect the optimum temperature for TTXI activity (i.e., 85°C), whereas mutation Ala62Pro and double mutation Gln58Pro/Ala62Pro decreased it by 7°C and 12°C, respectively (Figure 2A). Mutation Gln58Pro stabilized TTXI at 85°C: the enzyme’s half-life was extended from 69 to 99 min (a 43% increase). With half-lives of 6.2 and 21 min at 85°C, respectively, Ala62Pro and Gln58Pro/Ala62Pro mutant TTXIs were significantly less thermostable than the wild-type enzyme (Figure 3A). These surprising results indicated that Pro in position 62 destabilized TTXI. Similar results were obtained in precipitation experiments. Mutation Gln58Pro increased TTXI’s temperature of 50% precipitation by 6°C, whereas mutations Ala62Pro and Gln58Pro/Ala62Pro decreased it by approximately 4°C and 3°C, respectively (Table 2). 45 Figure 1: Three-dimensional structure of TNXI’s Phe59 loop region. Only parts of the tetramer’s subunits A and C are shown. Subunit C, shown in yellow ribbon, is interacting with subunit A’s Phe59 loop. Residues in subunit A’s Phe59 loop are colored based on their hydrophilicity (Blue-hydrophilic, Red-hydrophobic). 46 Specific activity (U/mg) Specific activity (U/mg) 40 50 60 70 80 90 100 50 60 70 80 90 100 Temperature (°C) Figure 2: Effect of temperature on the specific activities of TTXI and TNXI and their Phe59 loop mutant derivatives. The substrate used was glucose. (A) TTXI and its mutant derivatives. Symbols: —l:l—: TTXI; ---<>---: Gln58Pro; ----O----: Ala62Pro; "um-A Gln58Pro/Ala62Pro. (B) TNXI and its mutant derivatives. Symbols: —El—: TNXI; <> Pr058Gln; --------O: Pro62Ala; --------A. Pr058Gln/Pro62Ala. 47 Ln (residual activity) “.3 O/ Ln (residual activity) Ln (residual activity) -2 I T l l r 0 25 50 75 100 0 50 100 150 Time (min) Figure 3: Inactivation curves of TTXI and TNXI and their Phe59 loop mutant derivatives (A) TTXI and its mutant derivatives at 85°C. Symbols: —Cl—: TTXI; -----'<> Gln58Pro; --------:O Ala62Pro; ----A----. Gln58Pro/Ala62Pro. Half-lives of TTXI, Gln58Pro, Ala62Pro, and Gln58Pro/Ala62Pro are 69.3, 99.0, 6.2, and 21.0 min, respectively. (B) TNXI and its mutant derivatives at 95°C. Symbols: —Cl—: TNXI; mom; Pr058Gln; --------:O Pro62Ala; A. Pr058Gln/Pro62Ala. Half-lives of TNXI, Pr058Gln, Pro62Ala, and ProS8Gln/Pro62Ala are 69.3, 49.5, 7.1, and 7.1 min, respectively. 48 d, The reverse counterpart mutations in TNXI (Pr058G1n and Pro62Ala, and double mutation Pr058Gln/Pro62Ala) decreased the optimum temperature for TNXI activity (Figure 28). Not surprisingly, thermoinactivation curves revealed that mutation ProS8Gln decreased TNXI’s half-life from 69.3 min to 49.5 min (29% decrease). Both mutations Pro62Ala and Pr058Gln/Pro62Ala decreased TNXI’s half-life to 11.6 min (83% decrease), which confirmed that these mutations were all destabilizing as we expected (Figure 3B). Again, these results were confirmed by precipitation experiments. The temperature of 50% precipitation of both Pro62Ala and Pr058Gln were 95.7°C, 1.1°C lower than that of the wild-type TNXI. Double mutant Pro62Ala/Pr058Gln precipitated at an even lower temperature than each single mutant enzyme, suggesting that these destabilizing effects are additive. The kinetic features of the wild-type and mutant xylose isomerases with glucose and xylose as substrates were determined at 65°C for the TTXI series and at 80°C for the TNXI series (Table 3). With the exceptions of mutations Ala62Pro and Gln58Pro/Ala62Pro in TTXI, the mutations in TTXI and TNXI Phe59 loops did not significantly alter the enzymes’ catalytic properties. The Ala62Pro mutation increased both TTXI’s Vmax and Km on glucose, leaving its catalytic efficiency on glucose almost unchanged. This mutation had a much stronger effect on TTXI activity on xylose. It increased its affinity for xylose and its catalytic efficiency approximately 2.9 times. Mutation Gln58Pro/Ala62Pro had a more pronounced effect on TTXI activity on glucose. A 2.6-fold increase in its Km for glucose decreased its catalytic efficiency almost 3-fold. These results indicate that mutations Ala62Pro and Gln58Pro/Ala62Pro altered TTXI’s catalytic features and at the same time destabilized the enzyme. 49 Table 2: E of three in Table 2: Effect of mutations on enzyme precipitation temperatures. Results are the means of three independent experiments. Denaturation temperature for 50% Enzymes precipitation (°C) Wild-type TTXI 83.7 i 0.2 Gln58Pro 89.9 i 0.6 Ala62Pro 79.9 i 1.0 Gln58Pro/Ala62Pro 80.8 i 0.7 Wild-type TNXI 96.8 i 0.4 ProS8Gln 95.7 i 0.4 Pro62Ala 95.7 r 0.6 ProS8Gln/Pro62Ala 94.0 t 0.3 50 Caz—92-238532 of 3 33. 3_u:_o> EEC.— oLCC szc 32.6 swash: 32.536 9.03 33:5 56> wen EM .mEoEtono Eoc:..i.dz: 09F: .76 acne:— 05 was $237.91 80315.53? «:93»: :95 mo flan AUoOm any CAZF ficm abowc .3 CALLE~O 33.42552» 9.3.2339 um. 03:8 w.mm finev ad n mm 5; a fix Ev Wm; ad a md 5.3 a 0.8. £ A2810. :3 .3. ans: §> A28 ex ex} ex} moauam one—xx 08020 8208598 E02032: 00:: mo 9808 05 0S £30m 285850 5802820322 05 8 3% 5823 REE 05 mo mafia 8050 $585 35890 0.63 329’ §e> 28 av. £88598 5385005 ooh: 0o 885 20 on £33. 83388 cause use no one 608 as szn o5 6.8 5 Ext 0o 8388 2:36 ”m 23. The different I enzymes : stability 2 identified thermostu stabiliza'i T10 Obvic therm0>t COmp-are 1 "- Thes engineel entropy pTOIEin. 0f the c adOpt 0: Should : non‘Cm backbor dihedral DISCUSSION The high degree of similarity between TTXI and TNXI and their significantly different thermostabilities and thermophilicities make comparative studies of these enzymes attractive for understanding the key molecular features responsible for their stability and activity differences (Vieille et al., 1995). Multiple factors have been identified (Vieille and Zeikus, 1996) that can be responsible for a protein’s high thermostability. They include packing efficiency, hydrophobic interactions, loop stabilization, reduction of entropy of unfolding, electrostatic interactions, etc. In our case, no obvious differences in 'I'TXI and TNXI structures could explain their different thermostabilities except for a few additional prolines and fewer Asn+Gln in TNXI compared with TTXI. Two additional prolines are found in a large “Phe59 loop” region in TNXI (Figure 1). These prolines are substituted with Gln and Ala in TTXI. Proteins can be stabilized by engineering proline into selected sites thereby decreasing the proteins’ conformational entropy of unfolding (Allen et al., 1998). In order for a proline substitution to stabilize a protein, at least four criteria have to be considered: (i) the mutation site should allow one of the conformations allowed for proline; (ii) the preceding residue should be able to adopt one of the conformations allowed for the residue preceding a proline; (iii) proline should not create volume interferences; and (iv) proline should not destroy stabilizing non-covalent interactions (Nicholson et al., 1988). TTXI Gln58 and Ala62 have backbone dihedral angles allowed for prolines, and Asp57 and Lys61 have backbone dihedral angles allowed for residues preceding prolines. Despite these first two 52 confomiati‘or As seen in F proline p)TTC the wild-t}?t coxalent inte Gln58Pro. 'dl reduction of destabilizing and it is not eliminate an) mutation HlOt ring (C5 ator CarbOn atom: Waals interac The unfavoral [O [Oca] Conft EifftEct [he dCllt defiabilizing destabilizgmOn decrease in Um 1“ COnc [0 identify an (o a _ clueose’, 15mm conformational requirements being satisfied, only mutation Gln58Pro stabilized TTXI. As seen in Figure 4A, the conformation of Gln58’s sidechain is very close to that of the proline pyrrolidine ring. No volume interference is created by the Gln58Pro mutation. In the wild-type TTXI structure, Gln58 is not involved in any potentially stabilizing, non- covalent interactions (not shown). No enthalpic destabilization is expected with mutation Gln58Pro, and the stabilization it provides to TTXI probably entirely results from a reduction of the entropy of unfolding. On the other hand, the Ala62Pro mutation had a destabilizing effect on TTXI. In TTXI, Ala62’s sidechain points toward a large cavity, and it is not in close vicinity of any other residues. So the Ala62Pro mutation does not eliminate any stabilizing non-covalent interactions. Detailed analysis of the Ala62Pro mutation modeled into the 'I'TXI structure (Figure 5) suggests that Pro62’s pyrrolidine ring (C5 atom) is in close contact (within 2.92 A) with Lys6l’s sidechain (CB atom). Carbon atoms have Van der Waals radii of 1.70-1.78 A in protein. Optimal Van der Waals interactions between 2 carbon atoms would take place at approximately 3.4-3.5 A. The unfavorable Van der Waals contact between Pro62-C5 and Lys6l-CB probably leads to local conformational changes. Not only are these changes destabilizing, they also affect the active site structure and the enzyme’s interaction with the substrate. The overall destabilizing nature of mutation Ala62Pro indicates that the conformational destabilization of the native enzyme more than cancels the benefits of a potential decrease in unfolding entropy. In conclusion, our data show that genetic engineering approaches can be utilized to identify amino acid residues responsible for extreme thermal stability of xylose (glucose) isomerase thermozymes. Both proline58 and 62 appear to stabilize TNXI’s 53 .5on E 2a £53852 #6:?» E m_ moo. omega m.< zcanzm 6:3 E 2m 852mm: 88 9:8 0 255.6 do. E mm D :Sfism dimes—xx 5:938 Amy ”Exams—O 5:82: 2v AXE. E 225:8 n82 omega 2: Co 23023 65558662.; ”v Eswfi V“ 54 Figure 5: Van der Waals contacts between Lys61 and Pro62 in TTXI Ala62Pro mutant derivative. Carbon atom Van der Waals radii were arbitrarily fixed at [.7 A. Lys6l and Ala62 are in yellow. Pro62 is in green. 55 Phe59 lOOl mutation. Phe59 loop; whereas TTXI’s thermal stability can be enhanced by the Gln58Pro mutation. 56 I '5 >~ 30- —1.5 ‘3 .‘2: El ct: > u— .8 g S 20- - 1 '53 95 :2, 0 <1.) C: :2“ 104 -o.5 ”1 0 50 6O 70 80 90 1001100 20 4O 60 80 100120 Temperature (°C) Time (min) Figure 1: Activity and stability of TNXI and its active site mutant derivatives. (A) Effect of temperature on the specific activities of TNXI and its active site mutant derivatives. The substrate used was glucose. Symbols: —Cl—: TNXI; ---<>---: Trp138Phe; --------:O Vall85Thr; -------.A Trpl38Phe/Va1185Thr. (B) Inactivation curves of TNXI and its active site mutant derivatives at 95°C. Symbols: same as in (A). Half-lives of TNXI, Trp138Phe, Va1185Thr, and Trp138Phe/Va1185Thr are 69.3, 87.0, 69.3, and 99.0 min, respectively. 68 :8: _e e 082 sec 9.5. ofim 5me wd H 0.2 0; H 93: NE v.05 m.o H o.m_ _.m H mév Emw_~a>\0gmwm_&._i aém vanes 0.0 H 92 m._ H 0.2 «.2 swan: c; H Nom 2: H meme EH33; mdm wdvv ed H Em n; H 92 fie odan Wm H 52 Yam H v52 23min; VS ng a; H 0.3 h; H 0.: Wm Owen N._ H _.m_ mdm H 1.2: RE 095-253 QNM own 02 Wm H on odm o3 0.2 Nam H mm .Ehhw:0>\0amwm_fih vdm o: 03 n; H 2 5.0 own 0.: cs H 3 mar—.325» 0.2 owe ~.N_ : H 8 92 o3 v.2 v.5 H 3 .05528... «.3 co: 0.; Na H N_ w.m own ”9 as H o: .CCL. 090-33 $28.3 958 $2873 $52 $5 .30. A25 av. p-331 9:5 em 33 59> 23 ae> 8:555 083x 08050 05250098 5030355 00:: $0 5005 05 0.8 530M .8533 :0502.50552 05 9 £50 .5023 355 05 mo w5=m 00050 nwsefi 005030 0.83 85? §e> 53 ev— .3:05._0nx0 505.035 8.5 0e 28.: 2: ea ease .eeéeéee Ease :2: 0e e5 608 3 02h e5 6.8 3 it .8 85888 e536 ”N 23. 69 DISCUSSION The key molecular structure-function feature previously studied in Thermoanaerobacterium thermosulfurigenes xylose isomerase (TTXI) is residues in the active sites. Meng et al., (1991) have shown that mutation Trp138Phe significantly increased TTXI catalytic efficiency on glucose (2.7 fold increase). This increase was suggested to result from a better accommodation of glucose in the substrate-binding site because Trp was substituted by a smaller residue, Phe. This substitution was also shown to double TTXI’s half-life at 85°C (Meng et al., 1993). The higher thermostability may be explained as the consequence of a reduction of the water-accessible hydrophobic surface area. Another mutation that has been shown to increase TTXI’s catalytic efficiency on glucose was Va1185Thr. This increase was attributed to additional hydrogen bonding of Thr to glucose’s C6—OH. When these mutations were introduced into TNXI, we observed the same trends in terms of catalytic features. All the mutant enzymes, including the double mutant Trp138PheNa1185Thr, had a higher catalytic efficiency for glucose than the wild-type TNXI. These increases mainly resulted from a much lower Km for glucose than that of TNXI. On the other hand, all these mutant TNXIs showed a lower catalytic efficiency for xylose like they also did in TTXI. W138F was the mutation that had the most significant effect on TTXI’s catalytic efficiency on glucose. Here, TNXI V185T mutant derivative had the highest catalytic efficiency on glucose. This result came as a surprise since the two XIs’ active sites were almost completely superimposable. As seen in figure 2, Trpl38 and Val l 85 are almost completely superposed in TTXI and TNXI. The 70 Figure 2: Superposition of TTXI and TNXI active sites. Single letter code is used for amino acid residues. TTXI residues are in yellow. TNXI residues are in red. Metal site I (i.e., structural metal, M1) and metal site II (i.e., catalytic metal, M2) are represented by crosses. Both metal sites are occupied by C02+in both enzymes. Co2+ in M2 of TTXI and TNXI are 1.23 A apart. 7] conformations of the neighboring residues are also extremely conserved. The most significant difference between TTXI and TNXI active sites is the position of metal 11 (i.e., the catalytic metal, a Co2+ in both enzymes): the two cations are 1.23 A apart. This shift of the catalytic metal might affect the enzyme-substrate binding properties and/or the catalytic metal’s reactivity during catalysis, thus explaining why there was a difference in mutation effects on TTXI and TNXI catalytic activities on glucose. All TNXI mutant derivatives were as stable as (Va1185Thr) or more stable than the wild-type TNXI (Trp138Phe and Trp138Phe/Va1185Thr). These mutations did not increase TNXI stability to the same extent as they did in TTXI (Meng et al., 1993). Mutation Trp138Phe might have the same stabilizing potential in TNXI as in TTXI, but another TNXI molecular feature probably becomes limiting for stability before the full stabilization potentially provided by the mutation can be reached. These findings prove that it is possible to further stabilize hyperthermophilic proteins. To our best knowledge, with a maximal activity on glucose at 97°C of 45.4 Units/mg, TNXI Va1185Thr mutant derivative is the most active type II xylose isomerase ever reported. In conclusion, our data show that genetic engineering approaches can be utilized to identify amino acid residues responsible for high catalytic activity of xylose (glucose) isomerase thermozymes. TNXI’s thermal stability was further enhanced by the Trpl38Phe substitution. The Va1185Thr mutation significantly enhanced TTXI’s and TNXI’s Vmax and kcm/Km for glucose isomerization to fructose. This significant catalytic enhancement of glucose isomerase activity was made possible in large part by the template enzyme’s naturally evolved function as xylose isomerase. Genetic engineering altered the active site only to better accommodate glucose as the substrate. 72 10. REFERENCES Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, IA, and Struhl, K. (1993) Current Protocols in Molecular Biology. (Struhl, K., ed.), Greene Publishing & Wiley-Interscience, New York. Chen, W. (1980) Glucose isomerase. Process Biochem. 15, 30-35. Collyer, C. A., Henrick, K., and Blow, D.M. (1990) Mechanism for aldose-ketose interconversion by D-xylose isomerase involving ring opening followed by a 1,2- hydride shift. J. Mol. Biol. 212, 211-235. Dische, Z. and Borenfreund, E. (1951) A new spectrophotometric method for the detection and determination of keto sugars and trioses. J. Biol. Chem. 192, 583-587. Lee, C., Bagdasarian, M., Meng, M., and Zeikus, J. G. (1990a) Catalytic mechanism of xylose (glucose) isomerase from Clostridium thermosulfurofenes. J. Biol. Chem. 265, 19082-19090. Lee, C., Bhatnagar, L., Saha, B. C., Lee, Y. E., Takagi, M., Imanaka, T., Bagdasarian, M., and J. G. Zeikus. (1990b) Cloning and expression of the Clastridium thermosulfurogenes glucose isomerase. Appl. Environ. Microbiol. 56, 2638-2643. Meng, M., Lee, C., Bagdasarian, M., and Zeikus, J. G. (1991) Switching substrate preference of thermophilic xylose isomerase from D-xylose to D-glucose by redesigning the substrate binding pocket. Proc. Natl. Acad. Sci. USA 88, 4015-4019. Meng, M., Bagdasarian, M., and Zeikus, J. G. (1993) Thermal stabilization of xylose isomerase from Thermoanaerobacterium thermosulfurigenes. Proc. Natl. Acad. Sci. USA 90, 8459-8463. Sambrook, J ., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: a laboratory manual (2nd ed.), Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York. Vieille, C., Hess, M., Kelly, R. M., and Zeikus, J. G. (1995) xylA cloning and sequencing and biochemical characterization of xylose isomerase from Thermotoga neapolitana. Appl. Environ. Microbiol. 61, 1867-1875. 73 CHAPTER IV DIRECTED EVOLUTION OF Thermotoga neapolitana XYLOSE ISOMERASE: HIGH ACTIVITY ON GLUCOSE AT LOW TEMPERATURE AND LOW pH 74 ABSTRACT The Thermotoga neapolitana xylose isomerase (TNXI) is extremely thermostable and highly active at 95°C and above. Its mutant derivative, TNXI V185T, was the most active type 11 XI previously reported, with a catalytic efficiency (kw/Km) of 25.1 5'1 mM'1 toward glucose at 80°C (pH 7.0). To further optimize this enzyme’s potential industrial utility, two rounds of random mutagenesis and low temperature/low pH activity screening were performed using the TNXI Vl85T-encoding gene as the template. Mutants TNXI 3A2 (V185T/L282P) and IF] (V185T/L282P/F186S) were obtained after rounds one and two of random mutagenesis, respectively. TNXI lFl was more active than 3A2, which. in turn was more active than TNXI V185T at all temperatures and pHs tested. TNXI 3A2 and 1F1 high activities at low temperatures were due to significantly lower activation energies (57 and 44 kJ/mole, respectively) than that of TNXI and V185T (87 kJ/mole). This observation suggested that TNXI 3A2 and lFl’s increased activity at low temperature is a consequence of their increased flexibility in the active sites. Although 3A2 was more active than TNXI and V185T, its kinetic stability (based on the enzymes’ half life in different incubation conditions) was inferior to those of TNXI V185T possibly due to unfavorable van der Waal contacts of Pr0282’s pyrrolidine ring with neighboring mainchain atoms. This would, in turn, lead to conformational changes and eventually destabilize the enzyme. Unlike TNXI 3A2, 1Fl is more kinetically stable than TNXI and TNXI V185T. lFl’s enhanced stability is thought to be a result of additional H-bond formation between Ser186’s sidechain and the neighboring L229 residue’s mainchain structure. This, in turn, strengthens local conformation and the affinity of E231 co- 75 ordination with the structural metal, hence restoring the thermostability lost in 3A2. We showed here that low temperature/low pH activity of a hyperthermostable enzyme could be enhanced without costs to extreme thermal stability by directed enzyme evolution. 76 INTRODUCTION Xylose isomerase (XI) (EC.5.3.1.5) is an intracellular enzyme found in bacteria that can utilize xylose as a carbon substrate for growth (Chen, 1980). Due to its ability to use glucose as substrate and convert it to fructose, X1 is often referred to as glucose isomerase, and it is widely used in the industrial production of high fructose corn syrup (HFCS) (Meng et al., 1993; Bhosale et al., 1996). X1 is one of the three highest tonnage value enzymes, amylase and protease being the other two (Bhosale et al., 1996). Most industrially used XIs are isolated from mesophilic organisms (e.g., Streptomyces spp. and Actinoplanes spp.). The reaction temperature used in the current industrial glucose isomerization process is limited to 60°C because of by-product and color formations that occur at high temperature and alkaline pH, and because the isomerases themselves are not highly thermostable (Lee and Zeikus, 1991; Vieille and Zeikus, 2000). Thermostable XIs with neutral or slightly acidic pH optima have a potential for industrial applications. Performing isomerization at higher temperature than 60°C and neutral/slightly acidic pH with thermostable XI would allow faster reaction rates, higher fructose concentrations at equilibrium, higher process stability, decreased viscosity of substrate and product streams, and reduced by-products formation (Lee and Zeikus, 1991; Vieille and Zeikus, 2000). The XI from the hyperthermophile Thermotoga neapolitana (TNXI) has been studied extensively in our laboratory. The gene encoding TNXI (xylA) was cloned, sequenced, and overexpressed in Escherichia coli (Vieille et al., 1995). TNXI’s active site was engineered by site-directed mutagenesis to increase its activity on glucose 77 (Sriprapundh et al., 2000). The TNXI Vall85Thr (V185T) mutant derivative is more active, more glucose-efficient, and as stable as the wild-type TNXI. It was also the most active type II XI ever reported. Although TNXI V185T is highly thermostable and highly active at 97°C, it is very poorly active (10 % of maximal activity) at the current industrial isomerization temperature (60°C) and it requires a neutral pH for optimal activity. Rules for engineering protein activity and stability by rational design are likely to be protein-specific, and any such design effort would require prior detailed structural information. Numerous and intensive site-directed mutagenesis studies have probed this issue. Despite these efforts, considerable disagreement remains over which forces dominate stabilization mechanisms, and no generally applicable rules have been established (Giver et al., 1998; Vieille and Zeikus, 2001). Although protein chemists continue to elucidate the relationships between the sequence, structure, and function of proteins, the extensive knowledge that is necessary for the application of rational engineering approaches is available for only a tiny fraction of known enzymes. Directed evolution, on the other hand, has proved to be useful for modifying enzymes in the absence of such knowledge (Kuchner and Arnold, 1997). In directed evolution, the process of natural evolution is accelerated in a test tube for selecting proteins with the desired properties (Moore and Maranas, 2000). A typical experimental cycle of directed evolution begins with the creation of a library of mutated genes. Among the methods that introduce mutations randomly along the entire length of a gene (Leung et al., 1989, Stemmer, 1994, Zhao and Arnold, 1997, Shao et al., 1998, Zhao et al., 1998, and Ostermeier et al., 1999), error-prone PCR has been used the most extensively. The mutated genes are then ligated into an expression vector and transformed into suitable 78 bacterial cells. A screening procedure is next employed to identify the few transforrnants expressing proteins/enzymes with improved properties. Random mutagenesis and screening are repeated several times depending on the extent to which the properties of the protein should be altered and on the effects of mutations observed in each generation. Interest in engineering enzymes using directed evolution has grown significantly in the past few years. It has been used to increase enzyme thermostability, activity on novel substrates, substrate specificity, and enantioselectivity. For example, six generations of random mutagenesis, recombination, and screening stabilized Bacillus subtilis p- nitrobenzyl esterase significantly (>14°C increase in Tm) without compromising its catalytic activity at lower temperature (Giver et al., 1998). Here we use the TNXI V185T-encoding gene as the template for directed evolution to develop an enzyme active at 60°C and acidic pH. We show that activity can be increased significantly at low temperature and acidic pH without cost to the enzyme thermal stability. 79 MATERIALS AND METHODS Random mutagenesis: Random mutations were introduced into the TNXI V185T- encoding gene cloned between the NdeI and HindIII restriction sites of pET23a(+). PCR was performed with primers 5’-CGACTCACTATAGGGAGAC-3’ and 5’- GGTGGTGCTCGAGTGCG-3’ encoding sequences upstream of the Mid site and downstream of the HindIII, respectively, in pET23a(+). The reaction mixture contained 100 ng plasmid DNA, 50 pmol of each primer, 50 mM KCl, 10 mM Tris-HCl (pH 8.3), 1.5 mM MgC12, 0.4 mM dCTP, 0.4 mM dTTP, 0.08 mM dATP, 0.08 mM dGTP, and 2.5 Units Taq DNA polymerase (Roche, Nutley, NJ) in a 50 ul reaction volume. Cycling parameters were 36 cycles of 95°C for 45 sec, 50°C for 45 sec, and 72°C for 3 min. Amplification of the 1.4-kb product was checked by running a small aliquot of the reaction on a 1 % agarose gel. The PCR product was purified using the Geneclean III kit (BiolOl, Carlsbad, CA) and cloned back into the NdeI and HindIII sites of pET23a(+) using standard molecular biological techniques (Ausubel et al, 1993). For the second round of random mutagenesis, the gene encoding TNXI 3A2 was used as the template. Construction of a mutant library: The plasmids resulting from random mutagenesis were transformed into electrocompetent E. coli HB101(DE3) cells (XI-deficient) created using the ADE3 lysogenation kit (Novagen, Madison, WI). Transformants were plated on Luria-Bertani (LB) agar containing 100 ttg/ml ampicillin. After 16 hr of growth, single colonies were picked with sterile toothpicks and transferred into 24-well plates, each well containing 2 ml of LB plus 100 ug/ml ampicillin. Plates were then incubated overnight at 37°C on a shaker at 175 rpm to allow for cell growth. One hundred fifty microliters of 80 each culture were transferred to sterile 96-well plates. These plates were used to quantify bacterial growth by reading absorbance of the bacterial suspensions (OD595) in a microplate reader (Dynatech, McLean, VA), before being stored at 4°C to save the original cultures. The rest of the cultures were pelleted by centrifugation at 1,000 g for 10 min and resuspended in 200 ul of 50 mM MOPS (pH 7.0) containing 5 mM MgSO.; and 0.5 mM CoClg (i.e., buffer A). Bacterial suspensions were incubated with 50 pl of a 1 % (w/v) lysozyme solution at 37°C for 1 hr before being subjected to 3 freeze/thaw cycles (5 min in a dry ice-ethanol bath and 5 min in a 50°C water bath) to break the cells and release the enzymes into the supernatant. Cell-free crude extracts were then obtained by centrifugation at 1,000 g for 10 min and stored at 4°C for further use. Screening the mutant library for increased activity on glucose at 60°C and low pH: The crude extracts were assayed. for glucose isomerization in two conditions: 60°C (pH 7.0) and 80°C (pH 5.2). Assays were performed in microtiter plates with 150 pl of 100 mM MOPS (pH 7.0) or 100 mM sodium acetate (pH 5.2) containing 1 mM CoClg, 0.4 M glucose, and 10 til of crude extract. The plates were incubated at 60°C (pH 7.0) or 80°C (pH 5.2) for 10 min and placed on ice to stop the reactions. The fructose produced was assayed using the resorcinol-ferric ammonium sulfate-hydrochloric acid method (Schenk and Bisswanger, 1998). Ten microliters of each reaction were transferred to a new set of microtiter plates and mixed with 40 ul of distilled water and 150 pl of a freshly prepared 1:] mixture (v/v) of solution A (0.05 % resorcinol in ethanol) and solution B (0.216 g of FeNH4(SO4)2.12HzO in IL HCl). The plates were incubated in an 80°C water bath for 30 min to develop the color. The OD490 was measured with a microplate reader (Dynatech) 81 with 0-2.5 mM fructose as standards. A crude extract of HB101(DE3)pET23a(+) was used as the negative control on each plate. Crude extracts of HB101(DE3) expressing TNXI V185T and TNXI 3A2 were the positive controls in mutagenesis rounds one and two, respectively. Mutants with potentially higher activity on glucose than the positive control were selected on the basis of increases in both OD490 and OD490/OD595 relative to the positive controls in the two rounds of mutagenesis. Mutants showing increased activity were screened a second time using crude extracts prepared from 5 ml cultures. These crude extracts were prepared as described above, before being heat-treated at 80°C for 15 min and centrifuged. Oligonucleotide synthesis and DNA sequencing: PCR primers were synthesized by the Macromolecular Structure Facility, Department of Biochemistry and Molecular Biology at MSU. DNA sequences were determined either manually using the Thermosequenase kit (USB, Cleveland, OH) or automatically at the MSU Genomics Technology Support Facility. Protein Purification: Recombinant enzymes were purified as described (Vieille et al., 1995), followed by an additional ion-exchange chromatography step. Partially purified enzymes were applied to a DEAE-Sepharose column (2.5 x 15 cm) equilibrated with buffer A, and enzymes were eluted using a 500 ml linear 0—300 mM NaCl gradient in buffer A. The pooled fractions from the DEAE-Sepharose column were concentrated in a stirred ultrafiltration cell (30 kDa MW cut-off) (Amicon, Beverly, MA) and dialyzed 82 twice against buffer A. Concentrated, homogenous enzymes were dispensed and stored frozen at -70°C. Glucose isomerase assays: TNXI and its mutants were assayed routinely with glucose as the substrate. The enzyme (1-1.5 mg/ml) was incubated in 100 mM MOPS (pH 7.0) [or 100 mM sodium acetate (pH 5.5)] containing 1 mM CoClz and 0.4 M glucose at 80°C for 10 min. The reaction was stopped by transferring the tube to an ice bath. The amount of fructose produced was determined by the resorcinol-ferric ammonium sulfate- hydrochloric acid method (Schenk and Bisswanger, 1998). To determine the effect of temperature on activity, the enzymes were incubated in the reaction mixture at the temperatures of interest in a heated water (45-95°C) or oil bath (95-110°C) for 10 min. The effect of pH on activity was determined using the routine assay described above except that the MOPS buffer was substituted with 100 mM sodium acetate (pH 4.3-5.8), 100 mM PIPES (pH 6.1-7.0), or 100 mM EPPS (pH 7.2-8.1). All pHs were adjusted at room temperature, and the ApKa/At’s for acetate, PIPES, and EPPS (0, -0.0085, and - 0.011, respectively) (USB, Cleveland, OH) were taken into account for the results. To determine the kinetic parameters, assays were performed in 50 mM MOPS (pH 7.0) containing 10-1,500 mM glucose and 1 mM CoClg. One unit of glucose isomerase activity is defined as the amount of enzyme that produces 1 pmol of fructose per minute under the assay conditions. Thermal inactivation assays: To obtain the apo-enzymes (metal-free enzymes), the purified enzymes were incubated overnight at 4°C in 50 mM MOPS (pH 7.0) containing 83 10 mM EDTA. They were then dialyzed twice against 50 mM MOPS (pH 7.0) containing 2 mM EDTA, and they were finally dialyzed twice against 50 mM MOPS (pH 7.0) without EDTA. CoClz (0.5 mM) was added to the apo-enzymes and equilibrated at 4°C overnight before thermoinactivation assays. The time course of irreversible thermoinactivation was measured by incubating the enzymes (0.1-0.2 mg/ml) in either 10 mM MOPS (pH 7.0) or 10 mM sodium acetate (pH 5.5) at various temperatures for different periods of time in a heated water bath. Residual glucose isomerase activity was measured at 80°C as described above. The first order rate constant, k, of irreversible thermoinactivation was obtained by linear regression in semi-log coordinates. Enzyme half-lives were calculated from the equation: too) = ln2/k. Analysis of three-dimensional (3D) structures of TNXI and its variants: Enzymes were visualized on an IRIS-4D25 computer (Silicon Graphics Computer System, Mountain View, CA) using the INSIGHT 11 graphic program (Biosym Technologies, San Diego, CA). The TNXI pdb file (#lAOE) was obtained from the Protein Data Bank (www.rcsb.org/pdb). 84 RESULTS Construction of mutant TNXI libraries and screening for activity on glucose at low temperature and low pH: TNXI V185T is optimally active at 95°C - 97°C, but its activity at 60°C does not exceed 10% of its optimal activity (Sriprapundh et al., 2000). It retains only 20% of its optimal activity at pH 5.2. To increase this enzyme’s activity at 60°C and at acidic pH, and to gain insight into the factors determining the effects of temperature and pH on activity, we subjected the TNXI V185T-encoding gene to sequential random mutagenesis and to low temperature/low pH activity screening. Random mutations were introduced into the gene by error-prone PCR. The PCR conditions used were suggested to yield an average of 1-2 mutations per gene, conditions deemed optimal for the improvement of specific properties by mutagenesis and screening (Arnold and Moore, 1997). After the first round of random mutagenesis, 1,000 transforrnants were screened for their activity on glucose at low temperature (60°C, pH 7.0) and at low pH (pH 5.2, 80°C). Thirty mutants were identified that showed significantly higher activity (> 30% increase) than TNXI V18ST in both screening conditions. The phenotype of these mutants was tested again with heat-treated crude extracts prepared from 5 ml cultures. Higher activity on glucose was confirmed in only eleven out of the thirty crude extracts. XI expression level in these eleven crude extracts was checked by SDS-PAGE. Ten crude extracts showed higher XI content than the TNXI Vl85T control (data not shown). These ten mutants were discarded. The remaining mutant, TNXI 3A2, was purified to homogeneity. Once it was verified that TNXI 3A2 was significantly more active than TNXI V185T at 60°C and at pH 5.2, the gene 85 encoding TNXI 3A2 was used as the template in a second round of error-prone PCR and activity screening at low temperature and low pH. A library of ~l,500 transforrnants was screened using TNXI 3A2 as the positive control. A single mutant, TNXI 1F1, showed 80% and 40% increases in activity on glucose at 80°C (pH 5.2) and 60°C (pH 7.0), respectively, based on assays with heat-treated crude extracts. TNXI 3A2 and IF] were then purified to homogeneity. Their catalytic properties were studied in function of temperature and pH, and their thermostability was determined. Effects of temperature and pH on TNXI 3A2 and IF 1 activities: The effect of temperature on 3A2 and IF] glucose isomerase activities is shown in Figure 1A in comparison to the activities of TNXI and TNXI V185T. Both 3A2 and lFl show significantly higher specific activity on glucose than TNXI and TNXI V185T at all temperatures. At their optimal temperatures of activity (i.e., 90°C for lFl and 95°C for 3A2), both mutants are ~ 3-fold more active than TNXI V185T. Activation energies (Ea’s) for activity on glucose were calculated from the linear regressions shown in Figure 1B, using the equation A = Aoe'E‘VRT. Whereas TNXI V185T shows the same activation energy as TNXI (i.e., 87 kJ/mole), 3A2 and lFl show significantly decreased Ea’s (57 and 44 kJ/mole, respectively). These lower Ea’s explain why 3A2 and 1F1 are as much as 7.3 and 12.3 times more active, respectively, than TNXI at 60°C, but only 4.2 and 4.8 times more active, respectively, than TNXI at 90°C. The effect of pH on the activities of TNXI and its mutant derivatives is shown in Figure 2. 3A2 and IF] show significantly increased specific activity on glucose compared to TNXI and TNXI V185T over the entire active pH range. The activity 86 50 A . A O 40" A . A EB E o a 30- A ‘ 5’ o '5 .3 A «3 5 M.— '6' 0 a. U) 120 Temperature (°C) Figure 1: A) Effect of temperature on the specific activities of TNXI and its mutant derivatives on glucose at pH 7.0. (a): TNXI; (0): TNXI V185T; (O): TNXI 3A2: (A): TNXI 1F]. 87 Ln(specific activity) '1 l l l I 0.0026 0.0027 0.0028 0.0029 0.003 0.0031 l/Temperature ( l/ K) Figure 1: B) Ln (specific activity) versus l/Temperature. All linear regressions had r2 values above 0.97. (D): TNXI; (O): TNXI V185T; (O): TNXI 3A2; (A): TNXI lFl. Activation energies (E) for activity of TNXI, TNXI Vl85T, TNXI 3A2, and TNXI 1F] are 87, 87, 57, and 44 kJ/mole, respectively. 88 4O A A A A 30~ A go A \ O :3 C E O . O A :g 20- ‘ 8 A O .2 I": g c U) 107 ‘A' . .. . . . . I I. I i\- / 3‘. I v C A/ 0_ .l“- I I I 4 5 6 7 8 9 pH Figure 2: Effect of pH on specific activities of TNXI and its mutant derivatives on glucose at 80°C. (El): TNXI; (O): TNXI V185T; (O): TNXI 3A2; A TNXl 1F]. 89 increase is so significant that 3A2 and lFl are more active at pH 5.5 than TNXI and TNXI Vl 8ST are at pH 7.0. Kinetic parameters of TNXI 3A2 and IF]: The kinetic parameters on glucose of TNXI V185T, 3A2, and IF 1 were compared in different conditions of temperature and pH (Table 1). In all conditions tested, TNXI 3A2 and IF] showed higher Km and Vmax values than TNXI V185T did. At pH 7.0 (both at 60°C and 80°C), TNXI 3A2 and lFl ’s Vmax values increased more significantly than their Km’s for glucose, yielding important increases in catalytic efficiencies (up to 2.3 fold for lFl at 60°C [pH 7.0]). At 80°C (pH 5.5), the increases in TNXI 3A2 and lFl’s Vmax’s do not compensate for the major increases in their Km for glucose (i.e., 3.0 fold for 3A2 and 4.6 fold for lFl). In these conditions, TNXI 3A2 and lFl show catalytic efficiencies that are approximately half that of TNXI Vl8ST. At 60°C (pH 5.5), TNXI 3A2’s increase in Vnm does not compensate for a poor glucose affinity (high Km), resulting in a lower catalytic efficiency than that of TNXI V185T. Unlike TNXI 3A2, 1F1 has a higher catalytic efficiency on glucose than TNXI V185T does due to a dramatic increase (5 fold) in its Vmax that surpasses the increases in its Km (3.7 fold) in these conditions. Its 5-fold increase in Vmax makes lFl a 1.7 fold more active enzyme at 60°C (pH 5.5) than TNXI V185T is at 80°C (pH 7.0). Thermal stability of TNXI 3A2 and IF]: To determine whether the mutations present in 3A2 and IF] affected the kinetic stability of the mutated enzymes, the residual activities of 3A2 and lFl were measured after heat treatment at 80°C (pH 7.0) and 80°C 90 .5833 5202820222 05 9 9% 5623 325 05 mo mafia 886 smack: 8:350 203 82m> ee> 98 5% $258598 Eoufimuufi 025 we £808 2: 8a 333m. we 53 md H od— 3. H min 53 mum 5o H v.2 o.m H «.8 7: 5mg 3 new no H 3 ha H 33m m.~_ NS ho H ~12 Wm H mém N 39E. 9-28% :3 9:52 :25 p.263 A 73 ESE A25 EMNnov— «aux KQE> EM Eggs; «3 Kfl~=> EM 3 me 00% 3 mg 00% c6 ma: _.v H mwm 9g 2% 98. EN H 3% E. H can F: CmZF H QEN c.» mmfi to H Eva no H mdw— mom M62 ed H 3am YN H W? N ex eiém 3 me boom 3 mg 9.8 e.mo>zm>tou ESE: m: 98 Eng mo E80583 38:0. H. 2an 91 5 _ A I 4 5 Q ‘ ”>1 li‘Q :‘é " s I A ‘ To' U a g 4 - O . ‘ A o O A *- fl °\° I] 9 . O E“ [:1 O ._i 3.5 _, El 3 3 Cl 3 I I I I ' 0 100 200 300 400 500 Time (min.) Figure 3: A) Inactivation curves of TNXI and its mutant derivatives at 80°C (pH 7.0). Symbols used are the same as in Figure l. Half-lives of TNXI, TNXI V185T, TNXI 3A2, and TNXI lFl are 1.6 hr, 3.8 hr, 4.5 hr, 6.7 hr, respectively 92 B I 4.5 E Q E :3 2 § if, A 5 0 El 0 3 I I I O 100 200 300 400 Time (min.) Figure 3: B) Inactivation curves of TNXI and its mutant derivatives at 80°C (pH 5.5). Symbols used are the same as in Figure 1. Half-lives of TNXI, TNXI V185T, TNXI 3A2, and TNXI 1F1 are 1.3 hr, 2.3 hr, 1.7 hr, and 3.0 hr, respectively. 93 (pH 5.5) for various lengths of time (Figure 3). Stability experiments performed with the metal-free enzymes in 10 mM MOPS (pH 7.0) containing 0.5 mM CoClz showed that 3A2 and lFl (with “/2 of 4.5 hr and 6.7 hr, respectively) were kinetically more stable than TNXI (ti/2 of 1.6 hr) and TNXI V185T (II/2 of 3.8 hr). At pH 5.5, 1F] (tug of 3.0 hr) remained more stable than TNXI (ti/2 of 1.3 hr) and TNXI V185T (ti/2 of 2.3 hr); 3A2 was less stable (tug of 1.7 hr). Amino acid substitutions in TNXI 3A2 and IF]: The mutations present in 3A2 and IF] were identified by DNA sequencing. In addition to Va1185Thr already present in TNXI V185T, 3A2 contained a single additional mutation, Leu282Pro. The Leu282Pro mutation is located in helix 0t7 of the (Oi/[3)3-barrel structure, at approximately 12-14 A from the catalytic center (Figure 4). Helix a7 itself is located at the surface of a monomer and at the interface of the dimer. Neither Leu nor Pro’s sidechain can form hydrogen bonds with neighboring residues. Whenever a proline occurs in a peptide chain, it interrupts a—helices and creates a kink or bend (Lehninger, 1970). Detailed analysis of the Leu28Pro mutation modeled into the TNXI structure (Figure 5) suggests that Pr0282’s pyrrolidine ring (C7, and C5) is in close contact (in some cases ~1.7A) with mainchain atoms of residues Phe278 and Gln279. With van der Waal’s radii of 1.87 and 1.35 A for C and O atoms, respectively, in proteins, optimal van der Waal interactions between carbon atoms of Pro282 sidechain and the mainchain C and O atoms of residues Phe278 and Gln279 would take place at approximately 3.2 A to 3.7 A. The unfavorable van der Waal contacts (clashes) probably lead to local conformational rearrangements. These 94 changes might, in turn, affect the active site structure and dynamics, the enzyme’s interaction with the substrate, and probably inter-subunit interactions within the dimer. 1F1 contains the same two mutations as 3A2, plus mutation Phe186Ser. This last mutation is located in the active site, adjacent to mutation Vall85Thr (Figure 4). Serine’s sidechain is much less bulky than that of the original Phe. Residue 186’s sidechain points into the active site cavity, and it is close to the bulky sidechains of residues Tyr184, Phe228, Phe262, and Leu229. The Phe186Ser mutation probably leads to a rearrangement of the neighboring residues. This change in local packing may in turn be responsible for the large increase in low temperature activity of mutant 1F]. 95 Figure 4: Three-dimensional model of the TNXI lFl monomer showing the positions of mutations V1 8ST, F1863, and L282P. 96 62 E 550% 2m 2658an RC Nwmoi use A p— £11 0 l MHC (Kcal/mole*K) '— N O O o c U: C O I I 70 80 90 100 110 120 O\ O Te mpe rature (°C) Figure 4: Thermal unfolding of (A) TNXI 1F 1 and (B) GensweetTM in the presence of 5 mM MgSO4 and 0.5 mM COC12 followed by DSC. 123 Kinetic parameters of TNXI lFl and Gensweetm: The kinetic parameters on glucose of TNXI 1Fl and GensweetTM were compared in different conditions (Table 1). In all comparable conditions, GensweetTM has higher Km and Vmax than TNXI lFl (except at 80°C, pH 5.5 in which TNXI lFl’s Vmax is higher than that of GensweetTM). The difference in Km of GensweetTM and TNXI lFl is more pronounced than that of Vmax resulting in worse catalytic efficiency (kw/Km) on glucose for GensweetTM than TNXI lFl. The TNXI 1F1’s superiority of glucose catalytic efficiency on glucose is more noted at pH 7.0 than at pH 5.5 and at higher (95°C) or lower (60°C) than both enzymes’ optimal temperatures. Modeled fructose productivity: The lifetime fructose productivity of both enzymes was estimated using the one phase inactivation model at various combinations of temperatures and pHs. The modeled time course of fructose productivity of TNXI 1F] and GensweetTM at various conditions is shown in (Figure 5 and Table 3). Fructose productivity of GensweetTM at 80°C cannot be generated because no residual activity was detected at 80°C in buffer pH 7.0 and 5.5 after just 10 minutes. At 60°C, GensweetTM produced a maximum amount of 1.3 and 0.4 kilogram (kg) fructose per gram (g) enzyme at pH 7.0 and 5.5, respectively. With lifetime fructose production of 30.5 and 4.4 kg fructose/g enzyme at pH 7.0 and 5.5, respectively, TNXI 1F1 yielded approximately 24 and 12-fold 124 85:53 82 dz... he: mww Y: N60 04» Nwm m8 wwh 7.: Enz.—r 0; wow_ 0. :u ofio: 0d *2: m.mm _.©mw_ anokwmcomv 0000 Tum Oma_ aSm O.Nm 0.0 was; m.mm QVBN 7.: CHE Nam wabm mane vdav YN 5N2 hdm ode 53026.50 Doom CZ DZ OZ DZ 0.x. CONN _.mm w..vmm T: CHE DZ DZ DZ DZ DZ DZ DZ CZ Euooamcomv Uooo h6~ mwOM ode céwfi DZ DZ DZ DZ 7: USAF m; wwwm $60 méwNN DZ DZ DZ *DZ 552.550 Uoma nemwfic pee... £83 Ea} 5%me re} £83 Eaves. ease o.» ma 3 ma 2.803800 .m> _ n: Caz... mo 880883 0:35 nonmwtoaofl 883w mo cemwaafioo L 2an 125 _-. W, _ , ,,, ._ .. . a 1 +Gensweefi00lpH 7.0 l —O-—Genswee160CIpH 5.5 ! +TNXI 1F1 GOO/pH 7.0 —D—TNX| 1F1 SOC/pH 5.5 —I—TNX| 1F1 soc/pH 7.0 4000 3000 2000 Productlvlty (g Fructose/g Enzyme) 1 3654 7306 10959 1461 1 Time (min.) 35000 + GenswEt-BOCIpH 7.0 30000 , ' —o— Gensw eet 60C/pH 5.5 —I- TNXI 1 F1 60ClpH 7.0 ,--D— INXI1F1 SOC/pH 5.5 l —l— INXl1F1 BOOpH 7.0 20000 —D— TNXI 1 F1 BOCJpH 5.5 15000 10000 Productivity (9 Fructose/g Enzyme) 5000 0 E' 1— O) (O (V) 0 w to N O) h v ‘- w to t') O N w h (D m 0') N V- O') 00 Ix (D (O N ‘- O) N m w ‘- g N O N In w r- l~ O 0') LO (0 (D O, n O) 0) to O) N (D N (D 0') N v— F v- N N N (*) (fl 0') V V V In Time (min.) l 1 Figure 5: Estimated fructose productivity of TNXI 1F] and GensweetTM in different co nditions. 126 increases in fructose production compared to GensweetTM under the same conditions. The main reason for superior fructose productivity by TNXI 1F1 over GensweetTM is mainly due to its higher thermostability. It is also important to note that at 60°C the fructose production of GensweetTM reached the maximum points before 24 hr whereas it took approximately 15 days at pH 5.5 and more than 30 days at pH 7.0 for TNXI 1F1 to reach its maximum production. At 80°C, TNXI 1F1 produced 4.5 and 2.4 kg fructose/g enzyme at pH 7.0 and 5.5, respectively. The fructose production for TNXI 1F1 at 80°C, both pH 7.0 and 5.5, reached the maximum points at approximately 2.5 days. Fructose production experiments: Fructose production by TNXI 1F1 vs. GensweetTM with 45 % glucose syrup at various combinations of temperatures and pHs was performed to study the effect of both temperature and pH on fructose conversion ratio (compared to glucose) and potential browning reactions with each enzyme (Figure 6, 7 and Table 3). To prove that higher isomerization yield of fructose may be achieved by increasing the reaction temperature, TNXI 1F1 and GensweetTM were incubated with glucose syrups (pH 7.0 or 5.5) at 60, 80, and 90°C for up to 24 hr. At both pH, increases in fructose conversion was observed to be proportional to higher temperature for up to at least 24 hr in all cases with one exception. The syrup incubated with TNXI 1F1 at 90°C, pH 7.0 showed a higher conversion percentage for up to 6 hr after which the conversion rate remain relatively constant and its fructose conversion percentage was surpassed by that of the syrup incubated with TNXI 1F1 at 80°C, pH 7.0. The explanation for this event might be due to TNXI lFl’s relatively short half-life at high temperature above 85°C. The 127 OJ U! 00 O ; —I-TNXI 1F1 BOO/pH 7.0 ' :—D—TNXI 1F1 80C/pH 5,5 l .» +Gensweet 8OC/pH 7.0 l —O— Gensweet 80C/pH 5.5 i N (11 Hip—i % Fructose .1 N 01 o l l 10 l 5 0 l t l | I l l I l l T f J 0 2 4 6 8 10 12 14 16 18 20 22 24 Time (hr) 25 l B a, 3:. z . A- a) 15 _ ~—I—TNX11F160C/pH7.0 .3 —D-TNXI 1F1 60C/pH 5.5 .3. +Genswee160C/pH 7.0 I; 10 . —O—Gensweet SEQ/r1146; (J1 0: i l l 1 1T 1 l I in 0 2 4 6 81012141618202224 Time (hr) Figure 6: Experimental fructose conversion of TNXI 1F 1 and GensweetTM at pH 7.0 and 5.5. at 80°C (A) and at 60°C (B). 128 f: 1 l 2-0 —u- TNX|1F1 pH 7.0 90c l -D— TNXI1F1 pH 5.5 soc 1 5 +TNXI 1F1 pH 7.0 soc a ' —O—TNXI 1F1 pH 5.5 80C 3 + Gensweet pH 7 0 80C 0 1.0 , figs-censweet pH 5.5 800 0.5 0.0 . . . i . . . 0 2 4 6 8 10 12 14 16 18 20 22 24 Time (hr) Figure 7: Browness of syrups from experimental fructose conversion of TNXI 1F1 and GensweetTM. The time course of browning reactions was monitored by maximal absorbance at 425 nm. reactions were performed at two different pHs, 7.0 and 5.5, to also investigate the feasibility of prevention of browning the syrups that occurs at high temperature and pH. The brownness of the resulting syrups was monitored at the maximum wavelength of absorbency, which is 425 nm. As expected, the brownness of the syrups was observed to be most pronounced at the highest temperature tested (90°C) and pH (7.0). Syrups resulting from reactions at low pH (5.5) have dramatically fewer problems with browning. Fructose conversion percentage was compared between TNXI IF] and GensweetTM. At 80°C, TNXI 1F1 compared favorably with Gensweet with slightly higher conversion percentage at 24 hr (32 % for TNXI 1F1 vs. 28 % for GensweetTM), which is possibly due to GensweetTM’s less stable nature at high temperature. However, at 60°C, TNXI 1F1 was obviously better than Gensweet at converting glucose to fructose with higher conversion rate throughout the time course of 24 hr. 130 DISCUSSION The initial goal of this study was to compare biochemical and kinetic parameters as well as productivities of a laboratory-evolved xylose isomerase TNXI IF] and a commercially available glucose isomerase, GensweetTM to ascertain that TNXI 1F1 can be genuinely considered for industrial glucose isomerization. Table 2 summarizes key properties of the two enzymes. Not surprisingly, TNXI 1F1 compares favorably with GensweetTM in every aspect. The key factor that distinguishes the two enzymes was shown to be their thermal stability difference. Although TNXI 1F] and GensweetTM have almost the same apparent temperature optima (e.g., 90°C and 85°C, respectively), GensweetTM is much less thermostable than TNXI 1F1 by more than one order of magnitude at 60°C at pH 7.0 or 5.5. A mathematical model derived to account for the effect of temperature on reversible enzyme kinetics, inactivation rates, and the glucose-fructose chemical equilibrium (Bandlish et al., 2002) was used to estimate their lifetime fructose productivity. Because Km, km, and k1) are based on soluble enzyme data, the effect of immobilization is not taken into account. However, these estimates provide useful information concerning the potential of the enzyme for HFCS production under optimal conditions. TNXI 1F1 has the lifetime fi'uctose productivity at 60°C, pH 7 .0 of 30.5 kg fructose/ g enzyme whereas Gensweet”, which reached its maximum fructose conversion in less than a day due to its limited thermal stability, produced only 1.3 kg fructose/g enzyme. TNXI 1F1 ’s estimated greater fructose productivity mainly resulted fi'om both 131 Table 2: Comparison of thermal activity and stability properties of TNXI 1F1 vs. GensweetTM Properties GensweetTM TNXI 1F1 Tom (°C) 85 90 Optimal pH 7.5 6.7 Tm of Holo-enzyme (°C) 93.4 107.3 Tm of Apo-enzyme (°C) 76.1 78.4 Tm at 60°C/pH 7.0 (hr) 2.9 115.5 TI/z at 60°C/pH 5.5 (hr) 1.7 38.5 132 greater thermal stability and better catalytic efficiency on glucose compared to those of Gensweet”. Experimental fructose conversion was performed with 45 % glucose syrups incubated with 50 ug of either TNXI 1F1 or GensweetTM in various conditions to simulate industrial conditions and also to study the effect of temperature and pH on browning reactions resulting from interactions of enzymes with reducing sugars. TNXI 1F1 has a slight edge in term of fructose conversion ratio in every condition tested compared to GensweetTM with maximal conversion observed at 80°C, pH 7.0 (32 %). The browness of resulting syrups was monitored up to 24 hr at maximal absorbance of 425 nm. Browning of syrup occurs much more pronounced at 90°C, pH 7.0 and can be greatly reduced by either lowering the pH, the reaction temperature or both. It should also be noted that at 60°C, browness of the resulting syrup is marginal. When the enzyme concentration in the experiment was increased from 50 g to 100g and 1 mg, not only did we see more pronounced browning, but precipitates were also observed in the resulting syrups (data not shown). A study of a glucose isomerase from Streptomyces rubiginosus by Visuri et al. (1999) suggested that in an industrial process, glucose isomerase inactivation is caused mainly by a Maillard-type browning reaction between the enzyme and the reactive substrates glucose and fructose resulting in inactive glycated protein complexes. From our data, we can speculate that using TNXI 1F1 at 60°C at pH 7.0 or even at slightly higher temperature or lower pH would result in sufficiently high fi'uctose yield and productivity as well as reduced concern on Maillard browning reaction. Further detailed study of immobilized TNXI 1F1 and GensweetTM on the production of high fructose 133 8555.8 .2 dz... .d v _.N OZ mnmv own Wm oo _.o V a6 2 Ommom we: o8 oc mo age. 2 3mm VNM: Wm cm 0; 92 cm move wvcm o8 ow 56 5o DZ 02 OZ Wm om NN WE we Q2 Q2 o6 oo 7.: HXZH _.o v n; OZ vom wmm Wm o0 mo v ad 2 £2 00: ox. an _.o Wm _ _ DZ OZ Wm ow : 02 NN DZ .32 OS ow 3583300 C: m BEANS onEhNam onPSNcm A V Ea E .58 80 2: 80 2: 3.0 2: o Amsmwwmmcv 0 8885 \ aha. a we mean w» e... 2% :a amea. eEEm omowoam e\e 53:36. 2&0on 33:03:65 535355 5863980 .m> _ n: 59E. mo £80888 cote—68m 882E mo acmEquU “m 2an 134 syrup in different conditions in underway to ascertain the potential use of TNXI 1F1 in industrial application. ACKNOWLEDGEMENT This work was supported by the National Science Foundation, grant no. BES- 0115754. We express our gratitude to Dr. Jay Shetty of Genencor International (Rochester, NY) for providing GensweetTM for this work. 135 ll. 12. 13. REFERENCES Bandlish, R. K., Hess, J. M., Epting, K. L., Vieille, C., and Kelly, R. M. (2002) Glucose-to-fructose conversion at high temperatures with xylose (glucose) isomerases from Streptomyces murinus and two hyperthermophilic Thermotoga species. Biotechnol. Bioeng. 80, 185-194. Bentley, L. S. and Williams, EC. (1996) Starch conversion. In Industrial Enzymology (Godfrey, T and West, S. I. eds), pp. 339-357. Stockton Press, New York. Bhosale, S. H., Rao, M. B., and Deshpande, V. V. (1996) Molecular and industrial aspects of glucose isomerase. Microbial. Reviews 60, 280- 300. Bucke, E. (1981) Enzymes and Food Processing (G. G. Birch, N. Blakebrough, and K. J. Parker, eds), 51-72. Applied Science Publishers, London. Hartley, B. S., Hanlon, N., Jackson, R. J ., and Rangarajan, M. (2000) Glucose isomerase: insights into protein engineering for increased thermostability. Biachim. Biophys. Acta. 1543, 294-335. Klibanov, A. M. (19833) Approaches to enzyme stabilization. Science 219, 722. Klibanov, A. M. (1983b) Stabilization of enzymes against thermal inactivation. Advan. Appl. Microbial. 29, 1-28. Pedersen, S. (1993) Bioprocess technology Tanaka A., Tosa T., Kobayashi T., Eds. Marcel Dekker, New York 16, 185-208. Schenk, M. and Bisswanger, H. (1998) A microplate assay for D-xylose/D-glucose isomerase. Enz. Microbial. Technol. 22, 721-723. . Sriprapundh, D., Vieille, C., and Zeikus, J. G. (2000) Molecular determinants of xylose isomerase thermal stability and activity: analysis of thermozymes by site- directed mutagenesis. Protein Eng. 13, 259-265. Sriprapundh, D., Vieille, C., and Zeikus, J. G. (2002) Directed evolution of Thermotoga neapolitana xylose isomerase: high activity on glucose at low temperature and low pH. Protein Eng. (submitted) Vieille, C., Hess, M., Kelly, R. M., and Zeikus, J. G. (1995) xylA cloning and sequencing and biochemical characterization of xylose isomerase from Thermotoga neapolitana. Appl. Environ. Microbial. 61, 1867-1875. Vieille, C., Sriprapundh, 8., Kelly, R. M., and Zeikus, J. G. (2001) Xylose isomerases from Thermotoga. Methods Enzymol. 330, 245-249. 136 14. Visuri, K. and Klibanov, A. M. (1987) Enzymatic production of high fructose corn syrup (HFCS) containing 55% fructose in aqueous ethanol. Biotechnol. Bioeng. 30, 917-920. 15. Visuri, K., Pastinen, 0., Wu, X., and Makinen, K. (1999) Stability of native and cross-linked crystalline glucose isomerase. Biotechnol. Bioeng. 64, 377-380. 137 CHAPTER VI ROLE OF METALS IN TNXI STABILITY AT EXTREMELY HIGH TEMPERATURES 138 ABSTRACT TNXI and its directed evolution mutants’ melting temperatures (Tm) were determined by DSC in the presence and absence of metals. With the holo-enzymes, all scans except TNXI 1F1 revealed two thermal transitions. Surprisingly, TNXI 1F1 showed only one thermal transition, at a slightly lower temperature than those of the wild-type TNXI and TNXI V185T second thermal transitions showing that 1F1 lacks the first thermal transition common in the wild-type TNXI, TNXI V185T, and TNXI 3A2 mutant. DSC of apo-enzymes of all TNXIs revealed only one thermal transition. The result suggested that the additional mutation that occurs in TNXI 1F1 mutant, Phe186Ser, is responsible for altered metal-binding property and also strongly enhances the metal requirement of TNXI as seen in a very low Tm of TNXI 1F1 apo-enzyme compared to those of the wild-type TNXI and TNXI V185T relative to their holo-enzymes. With a Tm difference of 289°C (approximately 2-fold higher that of TNXI) between apo- and holo- forms of TNXI 1F1, its extreme thermal stability is even more strongly metal—dependent than the wild-type TNXI. 139 INTRODUCTION Xylose isomerase from hyperthermophilic eubacterium T hermotoga neapolitana (TNXI) is optimally active at a temperature of 95°C or above. Its xylA gene was cloned, sequenced, and expressed in Escherichia cali, which yielded a recombinant XI with catalytic characteristics identical to those of the native enzyme (Vieille et al., 1995). Further study by gel filtration chromatography showed that the recombinant enzyme was both a homodimer and a homotetramer, with the dimer being the more abundant form (Hess et al., 1998). The ratio of dimer to tetramer was approximately 20:1, based on total protein assay data (Bradford, 1976). The two forms had comparable stabilities when they were thermoinactivated at 95°C. Differential scanning calorimetry (DSC) revealed thermal transitions at 99 and 109.5°C for both forms suggesting that the association of the subunits into the tetrameric form may have little impact on the stability and biocatalytic properties of the enzyme (Hess et al., 1998). Typically for X18, two divalent cations (Mg2+, (302+, and Mn2+) per monomer are required for catalytic activity and stability (Hess et al., 1998) . The three metals activate TNXI with C02+ being the best activating metal. Activity of the Mg2+-TNXI and the Mn2+-TNXI are approximately 40 and 16 %, respectively, of the activity observed with the Coz+-TNXI. Also, the stabilization provided by metals to TNXI is metal specific: the Mn2+-TNXI is significantly more stable than the C0931 and Mg2+-TNXI (Vieille et al., 2001). The active site of TNXI has been previously engineered to improve its catalytic efficiency toward glucose (Sriprapundh et al., 2000). The TNXI V185T is the most 140 efficient site-directed mutant with a 3.1 fold increase in its catalytic efficiency toward glucose and comparable kinetic properties on glucose and xylose. To further optimize this enzyme’s potential industrial utility, directed evolution (sequential random mutagenesis and low temperature/low pH activity screening) was successfully applied to the TNXI V185T-encoding gene to obtain enzymes that have high glucose isomerase activity at low temperature and low pH. The best mutant enzyme, TNXI 1F1 (containing V185T, L282P, and F186S mutations), was dramatically more active than TNXI V185T at all temperatures and pHs tested. TNXI 1F1 is also more kinetically stable than TNXI and TNXI V185T. TNXI 1F1 ’5 enhanced stability is thought to be a result of additional H-bond formation between Ser186’s sidechain and the neighboring L229 residue’s mainchain structure. This, in turn, strengthens local conformation and the affinity of E231 co-ordination with the structural metal, hence improving the thermostability of the mutant enzyme. In this study, the thermodynamic stability of TNXI, TNXI V185T and its directed evolution mutant derivatives, TNXI 3A2 and IF] were followed by DSC in both apo- and holo- forms to compare and contrast their unfolding behaviors with regard to metal ions requirement for their thermostability. We show here that increased thermal stability of xylose isomerase is associated with metal binding especially in TNXI 1F1 where the apo-enzyme was 30°C less stable than the holo-enzyme. 141 MATERIALS AND METHODS Protein Purification: TNXI and its mutant derivatives were purified using the procedure of Vieille et al. (1995) followed by an additional ion-exchange chromatography step. Partially purified enzyme was applied to a DEAE-Sepharose column (2.5x15 cm) equilibrated with buffer A (50 mM MOPS pH 7.0, 5 mM MgSO4, 0.5 mM CoClz), and the enzyme was eluted using a 500 ml linear 0—300 mM NaCl gradient in buffer A. The pooled fractions from the DEAE-Sepharose column were concentrated in a stirred ultrafiltration cell (MW cut-off 30 kDa) (Amicon, Beverly, MA) and dialyzed twice against buffer A. Concentrated, homogenous enzyme was dispensed and stored frozen at -70°C. Differential Scanning Calorimetry (DSC): DSC experiments were performed on a Nanocal differential scanning calorimeter (Calorimetry Sciences Corp., Provo, UT) using a scan rate of 1°C/min. Samples were scanned from 25°C to 100°C. To obtain the apo- enzymes (metal-free enzymes), the purified enzymes in buffer A were incubated overnight at 4°C in 50 mM MOPS buffer (pH 7.0) containing 10 mM EDTA. They were then dialyzed twice against 50 mM MOPS buffer (pH 7.0) containing 2 mM EDTA, and they were finally dialyzed twice against 50 mM MOPS buffer (pH 7.0) without EDTA. The apo-enzymes were scanned against 50 mM MOPS @H 7.0). Enzymes containing both Mg2+ and Co2+ were dialyzed against buffer A, then scanned against the dialysis buffer as control. 142 RESULTS TNXI mutants’ melting temperatures (Tm) were determined by DSC in the presence and absence of metals (Table 1). With the holo-enzymes (Figure 1), all scans except TNXI 1F1 revealed two thermal transitions. The TNXI wild-type and TNXI V185T went through thermal transitions at 101°C and then 110°C and 114.5°C, respectively. TNXI 3A2 went through these transitions at significantly lower temperatures (86.6°C and 101°C). Surprisingly, 1F1 showed only one thermal transition, at a slightly lower temperature than those of the wild-type TNXI and TNXI V185T second thermal transitions suggesting that 1F1 lacks the first thermal transition common in the wild—type TNXI, TNXI Vl85T, and TNXI 3A2 mutant. DSC of apo-enzymes (Figure 2) of all TNXIs revealed only one thermal transition at 96.5°C, 96.9°C, 84.4°C, and 784°C for the wild-type, TNXI V185T, TNXI 3A2, and TNXI 1F1, respectively. The result suggested that the additional mutation that occurs in TNXI 1F1 mutant, Phe186Ser, is responsible for altered metal-binding property and also strongly enhances the metal requirement of TNXI as seen in a very low Tm of TNXI 1F1 apo-enzyme compared to those of the wild-type TNXI and TNXI V185T relative to their holo-enzymes. A Tm difference of 289°C between TNXI 1F1 apo- and holo- form was observed. This high Tm difference is approximately 2-fold higher than those observed in TNXI, TNXI V185T, and TNXI 3A2 (135°C, 176°C, and 16.6°C, respectively). 143 100 , q , . 90 . —TNX1 WT i —-—TNXIV185T; 80 ------ TNXI3A2 ,_. . '--—-TNXllFl it 70 _ "a . \ E. 50 E [’1 Q 40 i: ’ I a 2 30 1 w 20 l \\ 10 . ‘ 0 I I l l l ll 70 80 90 100 110 120 130 Temperature (°C) Figure 1: Thermal unfolding of the holo-forms of TNXI and its mutant derivatives in the presence of 5 mM MgSO.; and 0.5 mM CoClz followed by DSC. 144 120 V ~ _ —##‘ i l ——TNXI WT g 100 —TNXIV185T :4 4_TN?