Structure-function relationships in a f3-1,4
Transcription
Structure-function relationships in a f3-1,4
Structure-function relationships in a f3-1,4-glycanase (Cex) from Cellulomonas fimi: identification of catalytic residues by Alasdair Muir MacLeod B.Sc. University of Western Ontario, 1986 M.Sc. Laurentian University, 1989 A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in THE FACULTY OF GRADUATE STUDIES (Genetics Graduate Program) We accept this thesis as conforming to the required St rd THE UNIVERSITY OF BRITISH COLUMBIA September 1994 © Alasdair Muir MacLeod, 1994 presenting In thesis this partial in of fulfillment the requirements for an advanced degree at the University of British Columbia, I agree that the Library shall make it freely available for reference and study. I further agree that permission for extensive copying of this thesis for scholarly purposes may be granted the head by of my department or by his or her representatives. It is understood that copying or publication of this thesis for financial gain shall not be allowed without my written permission. (Signature) Department of 6 o ar The University of British Columbia Vancouver, Canada Date -. 2 44 11 ABSTRACT The objective of this study was to identify and examine the roles of catalytic residues in the active site of a 13-1,4-glycanase (Cex) from the bacterium Cellulomonas fimi. The use of the Gram positive bacterium, Streptomyces lividans, as a host for the expression of the cex gene was explored. The cex gene was successfully expressed in S. lividans from the aph promoter of p1J680-cex. The polypeptide was efficiently exported to the culture medium. Yields of recombinant polypeptide were about 6 mg/L following 40 hours of growth. Cex produced by S. lividans had a molecular mass greater than that produced by E. coli due to glycosylation. Similar to the native enzyme from C. fimi, the glycosylation protected the enzyme from a C. fimi protease, particularly when the enzyme was bound to cellulose. Cex hydrolyses E3-1,4-glycosidic bonds in cellulose with retention of anomeric configuration, releasing 13-cellobiose. On the basis of amino acid sequence alignments, Glu 233 was proposed as the catalytic nucleophile in Cex. The kinetic parameters determined for Glu 233 mutants generated by site-directed mutagenesis were consistent with this interpretation. Similarly, Glu 127 was proposed as the acid/base catalyst in Cex. The kinetic parameters were determined for Glu 127 mutants generated by site-directed mutagenesis, using a range of soluble cellobiosides and glucosides with differing requirements for acid catalysis. The results were consistent with a role of Glu 127 as acid/base catalyst. In the presence of azide, a new product, f3-cellobiosyl-azide was formed with uncharged Glu 127 mutants and catalytic activity was restored to wild-type levels. The results suggest azide occupied a vacant anionic site consistent with the removal of the acid/base catalyst. The approach used in this study could be applied to the identification of the acid/base catalyst in other {3-l,4-glycanases, particularly in the absence of threedimensional structure information. 111 TABLE OF CONTENTS ABSTRACT ii LIST OF TABLES vi LIST OF FIGURES vii LIST OF ABBREVIATIONS ix ACKNOWLEDGMENTS xi 1. INTRODUCTION 1 1.1 Structure-function relationships 1 1.2 Cellulose and cellulases 1 1.3 Cellulomonasfimi f3-1,4 glycanase (Cex) 3 1.4 Catalytic mechanisms for hydrolysis of a 13-1,4 glycosidic linkage 5 1.5 Identification of catalytic residues 6 1.6 Expression of the cex gene 15 1.7 Objectives 17 2. MATERIALS AND METHODS 18 2.1 Buffers, enzymes and media components 18 2.2 Bacterial strains, plasmids and phage 18 2.3 Media and growth conditions 19 2.4 Recombinant DNA techniques 21 2.4.1 Production and isolation of single-stranded DNA 21 2.4.2 Site-directed in vitro mutation of cex 22 2.4.3 Sequencing of DNA 23 2.5 Detection of cex gene expression 2.5.1 MUCase and PNPCase assays 24 24 iv 2.5.2 Binding to Avicel . 2.6 Detection of protein 24 25 2.6.1 SDS-PAGE and Western Blotting 25 2.6.2 Glycoprotein detection 26 2.6.3 N-terminal amino acid sequencing 26 2.6.4 Determination of protein concentration 26 2.7 Production and purification of recombinant Cex 26 2.7.1 Production of Cex in Streptomyces lividans TK64 26 2.7.2 Production of Cex and Cex mutants in E. coli 27 2.7.3 Purification of Cex 27 2.8 Proteolysis of Cex 28 2.8.1 Substrate-bound Cex 28 2.8.2 Cex in solution 29 2.9 Enzyme kinetics 29 2.9.1 Determination of steady-state kinetic parameters 29 2.9.2 Kinetics of inactivation of Cex mutants with 2F-DNPC 30 2.10 Characterization of the products of enzymatic hydrolysis 31 2.10.1 Thin layer chromatography 31 2.10.2 H 1 -NMR spectrometry of the products of enzymatic hydrolysis 32 2.11 Protein Mass determination 32 2.12 Amino acid sequence alignments and database searches 32 3. RESULTS (part 1): expression of the cex gene in Streptomyces lividans 33 3.1 Construction of p1J702-cex and p1J680-cex 33 3.2 Production of Cex by S. lividans TK64[pfJ702-cex I 33 3.3 Production of Cex by S. lividans TK64[p1J680-cex.] 39 V 3.4 Glycosylation of Cex by S. lividans . 4. RESULTS (part 2): identification of catalytic residues in Cex 44 49 4.1 Wild-type Cex: kinetics 49 4.2 The putative catalytic nucleophile in Cex: Glu 233 54 4.2.1 Generation of mutants at position 233 4.2.2 Determination of kinetic parameters for Glu 233 mutants 4.3 Identification of the acid/base catalyst 4.3.1. Generation of mutants at position 127 4.3.2 Kinetic characterization of mutants of Glu 127 55 60 63 63 65 4.3.3 Effects of sodium azide on reaction rates 72 4.3.4 Effects of sodium azide on products of hydrolysis 78 4.4 Asp 123: another conserved acidic residue 5. DISCUSSION 85 91 5.1 Expression of the cex gene in Streptornyces lividans 91 5.2 Catalysis 94 5.2.1 The catalytic nucleophile 94 5.2.2 The acid/base catalyst 95 6. APPENDIX 101 6.1 Determination of kcat and Km: an example 101 6.2 Interpretation of kcat, K 1 and kcat /IKm 104 6.3 Nucleotide and amino acid sequence of Cex 106 7. REFERENCES 108 vi LIST OF TABLES Table 1.1 Family F of 13-1,4-glycanases 2.1 Bacterial strains 19 2.2 Plasmids, phagemid and phage 20 2.3 Oligonucleotides used in site-directed mutagenesis of cex 23 2.4 Wavelengths monitored and extinction coefficients for hydrolysis of aryl cellobiosides and aryl glucosides 30 3.1 N-terminal processing of native and recombinant Cex 43 4.1 Kinetic parameters for hydrolysis of cellobiosides and glucosides by wild-type Cex 53 . 11 4.2 Kinetic parameters for hydrolysis of p-nitrophenyl-13-D cellobioside (PNPC) by Cex and Cex P233 mutants 62 4.3 Mass spectrometry of purified Cex and E127 mutants in the absence and presence of 2,4-DNPC 66 Kinetic parameters for hydrolysis of various substrates by Cex and E127 mutants 68 4.5 HNMR spectra for 13-cellobiosyl-azide and -glucosyl-azide 1 82 4.6 Products of hydrolysis with wild-type and E127 mutants in the presence or absence of sodium azide 84 Kinetic parameters for hydrolysis of cellobiosides by D123A 88 4.4 4.7 vu LIST OF FIGURES 1.1 Cellulose and its hydrolysis 1.2 The exoglucanase/xylanase Cex from Cellulomonns firni 4 1.3 Stereochemical classification of f3-1,4 glycanases 5 1.4 Proposed mechanism for hydrolysis of glycosides by Cex 7 1.5 2F-DNPC: a mechanism-based inactivator of Cex 8 1.6 Cellobiosides with differing requirements for acid catalysis 13 3.1 Construction of p1J702-cex 35 3.2 Construction of p1J680-cex 37 3.3 Production of exoglucanase by S. lividans TK64[p1J702-cexl 38 3.4 Production of exoglucanase by S. lividans TK64[p1J680-cex] 40 3.5 Identification of extracellular Cex in S. lividans [p1J680-cexl cultures 42 3.6 Glycosylation of Cex by S. lividans 45 3.7 Sensitivity of Cex from S. lividans to the protease from C. fimi 48 4.1 Alignment of family F catalytic domains 50 4.2 Construction of pTZI8R-cex 57 4.3 Generation of pUC12-1.lcex(PTIS) encoding Cex mutants 59 4.4 Purification of Cex E233D and E233Q 61 4.5 Purification of Cex E127A, E127G and E127D 64 4.6 Inactivation of Cex EI27A by 2F-DNPC 71 4.7 PNPCase activity of Cex and E127 mutants in the presence of various concentrations of sodium azide 74 . 2 viii 4.8 4.9 4.10 4.11 4.12 Kinetic parameters for hydrolysis of PNPC by Cex E127 in the presence of various concentrations of sodium azide 75 Kinetic parameters for hydrolysis of 2,4-DNPG by Cex E127 in the presence of various concentrations of sodium azide 76 Kinetic parameters for hydrolysis of 2,4-DNPC by Cex E127 in the presence of various concentrations of sodium azide 77 Proposed mechanism for hydrolysis of glycosides by Cex E127A in the presence of azide 79 Hydrolysis of PNPC by wild-type Cex in the presence and absence of azide 80 Hydrolysis of PNPC by Cex E127A and E127G in the presence and absence of azide 81 Time course of hydrolysis of PNPC by Cex E127A in the presence of azide 82 4.15 Purification of Cex D123A 86 4.16 PNPCase activity of Cex D123A in the presence and absence of azide 87 Hydrolysis of PNPC by Cex D123A in the presence and absence of azide 90 4.13 4.14 4.17 ix LIST OF ABBREVIATIONS 2,4-DNPC 2”,4”-dinitrophenyl f3-D-cellobioside 2,4-DNPG 2’,4’-dinitrophenyl f3-D-glucoside 2F-DNPC 2”,4”-dinitrophenyl 2-deoxy-2-fluoro 13-D cellobioside 2F-DNPG 2”,4”-dinitrophenyl 2-deoxy-2-fluoro 13-D glucoside 4-BrPC 4” -bromophenyl f3-D-cellobioside A absorbance aa amino acid aph gene encoding aminoglycoside phosphotransferase Apr ampicillin resistance bp base pair Cex C. fimi exo 3-1,4-glycanase CMC carboxymethylcellulose DMSO dimethyl sulfoxide EDTA ethylenediamine tetra-acetic acid IPTG isopropyl-f3-D-thiogalactoside kb kilobase pairs kcat enzyme turnover number kDa kilodaltons K equilibrium binding constant k inactivation rate constant Km Michaelis cons Iant I<rnr kanamycin resistance lacZpo E. coli lacZ gene promoter and operator LB Luria-Bertani medium me! gene encoding tyrosinase (melanin) x Mr relative molecular mass MUC methylumbelliferyl 13-D-cellobioside NMR nuclear magnetic resonance PEG polyethylene glycol PFU plaque-forming units PNPC paranitrophenyl 13-D-cellobioside PNPG paranitrophenyl -D-glucoside PTIS portable translation initiation site R2YE regeneration yeast extract SDS-PAGE sodium dodecyl sulfate-polyacrylamide gel electrophoresis TE tris-EDTA Thr thiostrepton resistance TLC thin-layer chromatography TYP tryptone, yeast extract, phosphate maximal velocity of enzyme-catalysed reaction xi ACKNOWLEDGMENTS I wish to extend my thanks to my supervisor Dr. Tony Warren for providing continual advice and encouragement throughout this study. I also thank Drs. Doug Kilburn, Robert Miller and Neil Gilkes for their guidance. My appreciation also goes to Dr. Loida Carlson of the UBC Biotechnology lab who helped me with the Streptomyces expression. Dr. Steve Withers, Karen Rupitz, Dr. Thisbe Lindhorst and Dee Tull of the UBC Chemistry Department provided valuable discussions regarding the kinetic studies of Cex. In addition, Dr. Withers provided substrates crucial to the success of this study. Members of the cellulase lab, too numerous to mention, provided an enjoyable, enriching and stimulating environment in which to work. In particular, I’d like to thank Shen Hua, John Coutinho, Emily Kwan and Edgar Ong, former members of the Wesbrook room 206 family, for their mentorship. Patti Miller was indispensable to the running of the lab and for informing me when new flavours appeared at Baskin-Robbins. I am also grateful to Helen Smith for her original sense of humour and for providing the opportunity on many occasions to discuss science while descending Pocalolo or Exhilaration at Whistler/Blackcomb. I thank David for listening to a lot of stuff that may have seemed dry at times. I appreciate the financial support received from PENCE and from NSERC for a postgraduate scholarship. 1 1. Introduction 1.1 Structure-function relationships Modern molecular biological techniques are providing new tools with which to study the structure and function of biocatalysts, leading to new insights into the mechanisms of catalysis. Such information is making possible the engineering of novel biocatalysts with desired properties, including enhanced catalytic efficiency, altered substrate specificity and increased pH stability or thermostability. There is little doubt that a greater understanding of the structure-function relationships in enzymes, including an understanding of the roles that particular amino acid residues play in catalysis, will facilitate the design of new enzymes. 1.2 Cellulose and cellulases The study of structure-function relationships in enzymes that will hydrolyse plant biomass has received considerable attention over the last twenty years partly because of the economic potential of these enzymes in industries such as fuel production and pulp and paper (Yang et a!., 1992). Cellulose is the most abundant carbohydrate available from plant biomass (Beguin and Aubert, 1994). Cellulose molecules are linear polymers of 13-1,4-linked D-glucose residues. Native cellulose from plant cell walls consists of individual cellulose chains, each consisting of up to 10,000 glucose residues, which are arranged in parallel and are hydrogen bonded to each other to form an insoluble crystalline matrix (Blackwell, 1981; Figure 1.1 this study). Within the matrix, there are regions of high crystallinity and regions more amorphous in 2 A B9°9x Crystalline region 4 Amorphous region Adsorption of cellulases Endoglucanases Cellobiohydrolases 4 o.oo 3-glucosidases Figure 1.1. Cellulose and its hydrolysis (Adapted from Beguin and Aubert, 1994): A) Molecular structure of cellulose: 13-1,4 linked D-glucopyranose polymer (only 3 residues shown). B) Synergistic model for cellulose hydrolysis. Shaded glucopyranose residues indicate the reducing ends of the cellulose chains. 3 nature. Considerable effort has been directed towards the understanding of structurefunction relationships in enzymes which hydrolyse this substrate. Many microorganisms, both bacteria and fungi, are capable of hydrolysing cellulose. It is dear that the hydrolysis of cellulose presents a challenge to microorganisms in terms of both the accessibility and the insolubility of the substrate. Although there are a variety of strategies employed in the hydrolysis of cellulose to glucose by microorganisms, most rely on the secretion of a number of cellulolytic enzymes which act synergistically (Beguin and Aubert 1994). The generally accepted model for hydrolysis of cellulose is that three major classes of enzymes are involved (Figure 1.1). Endoglucanases (EC 3.2.1.4) hydrolyse f3-1,4 linkages in the internal regions of a cellulose chain, probably initiating attack at the amorphous regions of the cellulose fibril. This hydrolysis results in the liberation of oligosaccharides of various sizes which can then be acted on by exoglucanases or cellobiohydrolases (EC 3.2.1.91). The exoglucanases or cellobiohydrolases hydrolyze oligosaccharides generally from the non-reducing ends to liberate cellobiose units. Cellobiose can then be hydrolysed into glucose by the f3-glucosidases (EC 3.2.1.21). 1.3 Cellulomonas fimi 3-1,4-glycanase (Cex) Cellulomonas fimi is a Gram positive, coryneform, mesophilic bacterium capable of growth on cellulose. To date, several cellulase genes from this organism have been cloned, sequenced, and expressed in E. coli (Gilkes et a!., 1984a; Wong et a!., 1986; O’Neill et a!., 1986; Moser et a!., 1989; Coutinho et a!., 1991; Meinke et a!., 1991; 1993; 1994). These include four endoglucanases, CenA, CenB, CenC, and CenD, an exoglucanase/xylanase 4 (Cex) and a cellobiohydrolase (CbhA). A seventh cellulose-binding polypeptide (Cbpl2O) is currently being characterized (Shen et a!., 1994). Cex, 443 amino acids long, comprises an N-terminal catalytic domain (315 amino acids) and a C-terminal cellulose binding domain (107 amino acids) separated by a proline-threonine rich linker (Figure 1.2). The domains in Cex retain their respective catalytic and cellulose-binding properties when separated proteolytically (Gilkes et a!., 1988, 1989). Cex contains three disulfide linkages, two within the catalytic domain and a third within the cellulose binding domain (Gilkes, et at., 1991a). When expressed in E. coli, Cex has a molecular mass of 47 kDa. Native Cex from C. fimi is a glycoprotein with a slightly greater molecular mass of 49 kDa. The glycosylation is not necessary for the catalytic function of the enzyme (Langsford et at., 1984). When the enzyme is bound to cellulose, however, the glycosylation affords it protection from a protease present in culture supernatants of C. fimi (Langsford et at., 1987). 42aa 316aa U ss leader catalytic domain lO7aa U ss ss PT linker 20 aa 47.1 kDa I celuose binding domain Figure 1.2. The f3-1,4-glycanase Cex from Cellulomonasfirni aa = amino acid; ss = disulfide bond; PT = proline/threonine Cex was originally classified as a cellulase, however, it has significant activity on xylan, a 3-1,4 linked polymer of D-xylose (Gilkes et at, 1984b). Xylan, although chemically similar to cellulose, is structurally more complex (Settineri et at., 1965). 5 Xylan molecules may be twisted and the backbone substituted with arabinose, glucuronic acid or methyiglucuronic acid. Although the activity of Cex on insoluble cellulose is quite low relative to other cellulases, it has significant activity against a range of soluble cellobiosides, glucosides and xylobiosides such as p-nitrophenyl f3-D cellobioside (PNPC) and 2’,4’-dinitrophenyl f3-D glucoside (2,4-DNPG) and p nitrophenyl [3-D xylobioside (PNPXX) (Gilkes et a!., 1991 a; this study). 1.4 Catalytic mechanisms for hydrolysis of a 13-1,4-glycosidic linkage The hydrolysis of a 13-1,4-glycosidic linkage results in either the retention or inversion of the configuration at the anomeric carbon (CI) (Sinnott, 1990). 13-1,4-. glycanases can thus be classified according to their stereospecificity. Those enzymes for which the products of hydrolysis retain the [3-configuration (i.e. [3—> [3)are termed “retaining” enzymes. In contrast, those which catalyse the hydrolysis of this linkage with inversion of configuration at the anomeric carbon (i.e. f3 —> a) are termed “inverting” enzymes (Figure 1.3). HO } 0 { Retaining HO\.oR . Inverting OH Figure 1.3. Stereochemical classification of [3- 1,4 glycanases. Only one glucose residue is shown for simplicity. The site ofbond cleavage is indicated by . 6 Cex is a “retaining” enzyme. It hydrolyses 3-1,4-g1ycosidic bonds in cellulose and xylan, two residues in from the non-reducing terminus, with release of f3-cellobiose or xylobiose (Withers et at., 1986). Hydrolysis therefore occurs with retention of configuration at the anomeric carbon. It is generally accepted that hydrolysis of a l-1,4glycosidic bond by a retaining enzyme, such as Cex, involves the formation and hydrolysis of a covalent c-D-glycopyranosy1-enzyme intermediate via oxocarbonium ion-like transition states (Koshland et al., 1953; Withers et al., 1988) as shown in Figure 1.4. A number of amino acids must be involved in binding and stabilizing the transition state complexes. Two amino acids, the catalytic nucleophile and the acid/base catalyst play particularly important roles. In the formation of the glycosyl-enzyme intermediate (the glycosylation step) the acid/base catalyst donates a proton to the glycosidic oxygen which facilitates bond cleavage by stabilizing the leaving group. The glycosyl-enzyme intermediate is stabilized by the formation of a covalent bond between the carboxyl group of the catalytic nucleophile and the anomeric carbon (Cl) of the sugar. The second step (the deglycosylation step) involves the hydrolysis of the glycosyl-enzyme intermediate. In this step, the acid/base catalyst acts as a general base by accepting a proton from water. The resulting hydroxyl group then acts to displace the catalytic nucleophile resulting in a product which retains the j3-configuration at the anomeric carbon. 1.5 Identification of catalytic residues. The catalytic nucleophile in a retaining 13-1,4-glycanase, such as Cex, can be “trapped” using a mechanism-based inactivator such as 2 “,4” dinitrophenyl 2-deoxy-2fluoro f3-D cellobioside (2F-DNPC; see Figure 1.5). The basis for such trapping is that the presence of the fluorine on carbon 2 destabilizes the transition states for both the 7 ‘A/B 7’ o O1 Nuc Nuc -ROH ‘A/B 7’ Glycosylation o 0HO glycosyl-enzyme intemdiate Deglycosylation Nuc 0 2 +H A/B oc /\ 7’ o Hoo: O\fO Nuc Nuc Figure 1.4. Mechanism for hydrolysis of glyco sides by Cex. The acid/base catalyst is represented by an “A/W. The catalytic nucleophile is represented by “Nuc” Only one glucose residue is shown for simplicity. Adapted from Koshland et al., 1953. 8 formation and hydrolysis of the glycosyl-enzyme intermediate (refer to Figure 1.4) resulting in the slowing down of both steps of the reaction. With the presence of an activated leaving group on the substrate such as dinitrophenol the requirement for acid catalytic assistance is reduced facilitating an increase in the rate of the first step (glycosylation). 2 NO 2F-DNPC Figure 1.5. 2F-DNPC: a mechanism-based inactivator of Cex The net result is the accumulation or HtrappingI of the glycosyl-enzyme intermediate (Withers et al.,1990). The covalently-bound catalytic nucleophile can then be identified in proteolytic digests of the protein, using mass spectrometry to compare the masses of the peptides with those released from the native (non-inactivated) enzyme. Using the mechanism based inactivator 2 ‘,4 dinitrophenyl 2-deoxy-2-fluoro 3-D glucoside (2FDNPG), the catalytic nucleophile in an Agrobacteriurn 13-glucosidase (Abg), a retaining glucanase from another family of enzymes, was identified as G1u358 (Withers and Street, 1988; Withers et al, 1990). No reliable strategies have been devised for the unequivocal labeling of the acid/base catalyst in a retaining glycanase. Group specific reagents such as carbodiimides can be used to label carboxylic acid residues with unusually high pKa 13- 9 values; however, these frequently label at multiple sites. Affinity labels such as glycosyl epoxides have also proven unreliable in this regard (Legler, 1990). X-ray crystallographic analysis of enzyme or enzyme/inhibitor complexes often reveals residues that are suitably positioned to act as catalytic residues. Hen egg-white lysozyme (HEWL), a retaining 13-glycanase, was the first enzyme for which the threedimensional structure was determined by X-ray crystallography (Blake et al., 1965). Two acidic residues, Asp 52 and Glu 35 were determined to be correctly positioned in the active site to act as the stabilizing anion/nucleophile and the acid/base catalyst respectively. Although the mechanism is still a subject of investigation, it was essentially derived de novo from examination of the three-dimensional structure. However, it is estimated that three-dimensional structural information is available for only 4% of those proteins for which the amino acid sequence is known (Rost et al., 1993). Although some proteins are clearly related in amino acid sequence to proteins of known three-dimensional structure, for 5 of every 6 new genes sequenced, there is no homologous structure in current data banks (Rost et al., 1993). The catalytic domain of Cex has been crystallized (Bedarkar et al., 1992) but the structure has yet to be solved. To date, the genes encoding more than 170 cellulases and xylanases from various microorganisms have been cloned and sequenced. Currently, these enzymes can be grouped into about a dozen families (A to L) on the basis of amino acid sequence similarities in their catalytic domains (Henrissat et al., 1989). To date, three-dimensional structures have been published for only a few of these enzymes. Many of these families contain enzymes from bacteria and fungi, and some contain both endo- and exo -f3-1,4glycanases. Endoglucanases and exoglucanases clearly have active sites of somewhat different topologies. However, members of a given family have been shown to exhibit the same stereospecificity (Gebler et a!., 1992b). Therefore, the enzymes within a particular family hydrolyse 13-1,4-glycosidic bonds by the same mechanism (i.e. with 10 retention or inversion) suggesting that their catalytic residues are likely to be conserved. Cex belongs to family F of 3-1,4-glycanases (Henrissat et at., 1989). In 1989, twelve members could be assigned to this family on the basis of amino acid sequence alignments. Currently, twenty enzymes can be assigned to the family, sixteen of bacterial origin and four of fungal origin (Table 1.1). Family F-members are primarily categorized as 3-1,4-xylanases although activity against PNPC or methylumbelliferyl - cellobioside (MUC) has been reported for several of them (Grepinet et at., 1988; Luthi et at., 1990; Lin et at., 1991; Shareck et at., 1991; Haas et at., 1992). In addition, low activity against carboxymethylcellulose (CMC) has been reported in some instances (Luthi et at., 1990). Clearly, family F enzymes have a mixed specificity for both xylan and cellulose in contrast to the low molecular weight family G xylanases, which are not active against cellulose. In addition to Cex, which was the first family F enzyme to be crystallized (Bedarkar et at, 1992), two more enzymes in the family have been crystallized: XynA from Pseudomonas fluorescens (Pickersgill et at., 1993) and, very recently, Clostridium thermocellum XynZ (Souchon et at., 1994). To date, however, no three-dimensional information has been published for any member of this family. Furthermore, prior to the initiation of this study, no catalytic residues had been identified in any member of this family. However, in families such as family F, which contains a large number of enzymes, amino acid sequence alignments can serve to pinpoint potential catalytic residues in the absence of three-dimensional structure information. Acidic residues, particularly aspartates and glutamates, are the most likely to act as catalytic nudeophiles and acid/base catalysts (Sinnott, 1990; Zvelebil and Sternberg, 1988). Several such residues are conserved throughout family F (see Results, Figure 4.1). Site-directed in vitro mutagenesis has proven to be a valuable tool in probing the structure-function relationships of enzymes and has contributed greatly to an understanding of the role of specific residues in catalysis. Mutation of “critical” residues 11 previously targeted by chemical modification alone has often proved the residues not to have the roles proposed for them (Schimmel 1990). Even in cases where the threedimensional structure is known, site-directed mutagenesis experiments have been very revealing in confirming or contradicting existing beliefs derived from such information regarding the roles amino acids play in catalysis. For example, studies involving the Table 1.1 Family F of 3-1,4-g1ycanases’ Enzyme Organism XynA XynA XynA XynB XynA Ce1B ORF4 Cex Xyn XynX XynZ Xyn Xyn XynA XynB XynA XynA Xy1A Xyn XynA Aspergillus kawachii Bacillus sp. strain C-125 Butyrivibrio fibrisolvens Butyrivibriofibrisolvens H17c Caldocellum saccharolyticurn Caldocellum saccharolyticurn Caldocellum saccharolyticum *Cellulomonasfjmi Clostridium stercorarium (str F9) Clostridium thermocellum *Clostridium thermocellum Cryptococcus albidus Penicillium chrysogenum *pseudomonas fluorescens Pseudomonasfluorescens Ruminococcusfiavefaciens Streptomyces lividans Thermoanaerobacter saccharolyticum (B6A-RI) Thermoascus aurantiacus Thermophilic bacterial sp rt8.84 type # amino acids total cat Acc# f b b b b b b b b b b f f b b b b 328 368 378 635 312 1011 312 443 387 1087 809 311 353 585 555 1092 436 328 368 350 347 312 347 312 315 387 ? 315 311 353 345 272 332 436 D14847 P07528 P23551 X61495 M34459 A43802 M34459 L11080 D12504 M67438 M22624 JS0734 S31307 Xl 5429 P23030 P29126 M64551 b f b 1127 p.sq 685 ? ? ? M97882 P23360 L18965 Acc# = GenBank or swissprot database accession number; cat = catalytic (family F) domain; p.sq. = partial (protein) sequence; type: f = fungal, b = bacterial. ? = unknown; * = enzyme (or catalytic domain) has been crystallized; 1 As of Mar, 1994. 12 replacement of the putative catalytic nucleophile in T4 lysozyme (Asp2O) with Cys (Hardy and Poteete, 1991) or with Ala (Rennell et al., 1991) revealed the mutant retained 80% of its activity, casting into doubt the previous assignment of this residue as the catalytic nucleophile. The determination of Michaelis-Menten parameters for hydrolysis of various substrates by mutants of Cex generated by site-directed in vitro mutagenesis can provide insight into the roles that various amino acids play in catalysis. Wild-type Cex will hydrolyse aryl glucosides and aryl cellobiosides and such as p-nitrophenyl -D cellobioside (PNPC), 2”, 4”-dinitrophenyl f3-D-cellobioside (2,4-DNPC) and 4”bromophenyl 3-D-cellobioside (4-BrPC). These well-defined soluble substrates contain aglycon moieties with different leaving group abilities (Figure 1.6). Those of high pKa, therefore poor leaving group ability, will have a greater requirement for acid catalytic assistance for hydrolysis of the 3-1,4-g1ucosidic linkage than those of low pKa. In effect, hydrolysis of those substrates with a high pKa will depend on the presence of a functional acid/base catalyst. In contrast, those of low pKa should have a much reduced requirement for the presence of a functional acid catalyst for departure. A detailed examination of the kinetic parameters of mutants of Cex with these substrates, therefore, will aid in the identification of the acid/base catalyst. kcat values will yield information on how particular mutations effect the overall rates of hydrolysis of these substrates. Based on knowledge of the mechanism of Cex, examination of the Km and kcat/Km values will yield information on how particular mutations effect the overall rates of the glycosylation or deglycosylation steps. The generally-accepted double displacement mechanism for Cex (see Figure 1.4) can be expressed as follows: 13 SUBSTRATE pKa of aglycon unit 2 NO OH 3.96 HOOZ© HO HO 2,4-DNPC 7.18 PNPC Br OH 9.34 4-BRPC Figure 1.6. Cellobioside substrates with differing requirennts for acid catalysis. The site of bond cleavage is indicated by ?tt 14 >glycosylation deglycosylation 3 k E.S E+S E-G -ROH where, E S E.S E-G P = = = = = E+P ÷H20 concentration of unbound enzyme concentration of unbound substrate concentration of enzyme-substrate complex concentration of glycosyl-enzyme intermediate concentration of unbound product The value of K the apparent dissociation constant for all the enzyme-bound species, can be expressed as: Km = [El [SI or [all enzyme bound species I [El [SI [E.S] + [E-GI [2] It can be seen from equations [1] and [21 that mutations which reduce the rate of deglycosylation will result in an accumulation of the glycosyl-enzyme intermediate [E-G] and therefore result in a reduction in Km. Similarly, a reduction in the rate of glycosylation would be expected to increase the Km. These assumptions are valid provided deglycosylation is the rate-limiting step as shown in the Appendix. The effect of a mutation on the rate of the glycosylation step can also be estimated by examination of kcat/Km which presumably reflects the rate of the formation of the glycosyl-enzyme intermediate from the free enzyme and substrate. Another measure of the effect of a mutation on the glycosylation step of the reaction can be made by examination of the 15 rate of the inactivation of the mutant relative to the wild-type using the mechanismbased inactivator 2F-DNPC. Essentially, the reaction proceeds only through the glycosylation step as discussed earlier. That is, k 3 <<k , resulting in the accumulation 2 of a stable glycosyl-enzyme intermediate. A more detailed interpretation of kcat, Km and kcat/Km in terms of the rate constants shown in equation [1], and an explanation of the assumptions made above is presented in a separate section (see Appendix). In addition to analysing the kinetic parameters with the substrates as described above, the ability to restore activity to a mutant generated by site-directed mutagenesis could yield information into the role the particular amino acid residue plays in catalysis. Cupples et at. (1990) reported that the anionic nucleophile sodium azide enhanced the rate of hydrolysis of o-nitrophenyl r3-D glucoside (ONPG) by E461 mutants of E. coli - galactosidase, another retaining glycosidase. It has subsequently been shown by use of a mechanism-based inactivator that the nucleophile in E. coli f3-galactosidase is E537. It is suggested that a probable role for E461 is as the general acid/base catalyst in this enzyme (Gebler et a!., 1992a). Since [3-galactosidase hydrolyses f3-1,4-glycosidic linkages by the same (retaining) mechanism as Cex (Gebler et a!., 1992b) it is likely that sodium azide, or other anionic nucleophiles could enhance the rate of such a mutant in Cex. 1.6 Expression of the cex gene A necessary first step for the study of structure-function relationships of proteins is to have a system in place which facilitates high level production and isolation of the protein product under investigation. The cex gene has been successfully subcloned and expressed in E. coli strains from the E. coli lacZ promoter using pUC-based vectors. In E. coli, Cex is produced as a soluble polypeptide, exported to the periplasm, correctly processed and disulfide bonded (Gilkes et a!., 1.991a). The protein can be purified from 16 the periplasm or from whole-cell extracts of E. coli by affinity chromatography on cellulose (Gilkes et a!., 1991a; this study). Typical yields of purified wild-type Cex obtained from overnight batch cultures of E. coli JM1O1 [pUC12-1.lcex (PTIS)] are about 5-20 mg/L of culture. E. coli currently remains the microbial system of choice for the production of heterologous proteins from cloned genes. It would be useful however, to employ a host strain which would secrete the protein directly into the culture medium. Gram-positive bacteria have a single cell membrane. Consequently, proteins which are exported across the cell membrane are secreted directly into the culture medium. Streptomyces species are Gram-positive Actinomycetes which are well noted for their production of antibiotics and extracellular enzymes (Brawner et a!., 1991). Some naturally-occurring extracellular enzymes secreted by Streptomyces include polysaccharidases such as cellulases (Nakai et al., 1988) and xylanases (Shareck et a!., 1991). Of the approximately 500 species of Streptornyces known to date, the best characterized is S. coelicolor. As a host for the production and secretion of heterologous polypeptides, however, S. lividans is the most widely used (Hopwood et a!., 1985; Ghangas and Wilson, 1989; Steiert et a!., 1989; Taguchi et a!., 1989; Anne and Van Mellaert, 1993). Unlike other Streptornyces species, S. lividans is useful as a host for the expression of foreign DNA because it apparently lacks an endonuclease restriction system (Kieser and Hopwood, 1991). In addition, cloning vectors and protocols for efficient protoplast transformation (> i0 7 transformants/.ig DNA) have been fairly well established (Hopwood et a!., 1985). S. lividans may be particularly well suited as a host for the expression of C. fimi DNA. C. fimi is distantly related to Streptomyces lividans (Stackebrant and Woese, 1981). The genomes of C. firni and S. lividans are extremely rich in G+C, each with an average of about 75%. Hence, both organisms have an extremely biased codon usage of G or C in the third position. Based on the above, S. lividans appeared to be an attractive host for expression of the C. fimi cex gene. 17 1.7 Objectives The overall objective of this study was to contribute to a better understanding of the structure-function relationships in the exoglucanase/xylanase (Cex) from Cellulomonas fimi. The approach was, firstly, to explore the use of Streptomyces lividans as a host for the expression of the C. fimi cex gene. The main part of the study involved the characterization of mutants of Cex generated by site-directed in vitro mutagenesis. Residues were targeted on the basis of amino acid sequence alignments of family F 1- 1,4-glycanases and on knowledge of the types of residues likely to be involved in hydrolysis of a 13-1,4-glycosidic linkage by a retaining enzyme. Detailed kinetic analyses were carried out on various substrates in an attempt to identify both the catalytic nucleophile and the acid/base catalyst in Cex. Following the initiation of this study, the nucleophile in Cex was identified as Glu 233 by use of the mechanism-based inactivator 2F-DNPC as mentioned in section 1.5. (Tull et a!., 1991). Therefore, the major thrust of this study became identification of the acid/base catalyst. 18 2. Materials and Methods 2.1 Buffers, enzymes and media components. Buffers and solutions used in this study were prepared as described by Sambrook et a!. (1989). All buffer chemicals were obtained from Sigma. Restriction endonucleases, polymerases, ligase and nucleotides were from Pharmacia or New England Biolabs (NEB). SequenaseTM was from United States Biochemical (USB). CF-I cellulose was from Whatman. Avicel (type PH-i 01) was from FMC International. Substrates 2”,4”-dinitrophenyl 13-D-cellobioside (2,4-DNPC), 2’,4’-dinitrophenyl -D glucoside (2,4-DNPG), 4” -bromophenyl 13-D-cellobioside (4-BrPC), and 2 “,4” dinitrophenyl 2-deoxy-2-fluoro 13-D cellobioside (2F-DNPC) were a gift from Dr. Steve Withers, Dept. of Chemistry, UBC. 13-cellobiosyl-azide and f3-glucosyl-azide were synthesized by Dr. Thisbe Lirtdhorst, Dept. of Chemistry, UBC. Cellobiose, p nitrophenyl (3 -D-glucoside (PNPG) and p-nitrophenyl [3 -D-cellobioside (PNPC) were obtained from Sigma. All media components were obtained from Difco. 2.2 Bacterial strains, plasmids, and phage. The bacterial strains, plasmids and phage used in the study are described in Tables 2.1 and 2.2. Bacterial stocks were maintained at -70°C in LB medium containing 10% DMSO. Plasmid DNA was stored in TE buffer or water at -20°C. Phage were stored in TYP medium at 4°C. 19 Table 2.1. Bacterial strains Bacterial strain Genotype Reference or source E. coli JMIO1 supE thi z(1ac-proAB)[F’ traD36 proAB lacIZAMl5] Yanisch-Perron et a!., 1985 E. coli RZ1032 HfrKL16 P0/45 [lysA(61-62)j dutl ungl thu relAl Zbd-279::TnlO supE44 Kunkel et a!., 1987 S. lividans TK64 pro-2, str-6 Hopwood et a!., 1985 S. lividans TK24 str-6 Hopwood et a!., 1985 2.3 Media and growth conditions Luria-Bertani (LB) medium was described previously (Sambrook et at., 1989). TYP medium contained the following (per litre): 16 g tryptone, 16 g yeast extract, 5 g NaC1, 2.5 g 4 HPO adjusted to pH 7.0. Small-scale cultures of E. coli containing pUC 2 K , or pTZ-based vectors were grown in shake flasks at 225 rpm at 30°C or 37°C in LB or TYP supplemented with ampicillin (Amp) at 100 !lg/mL. Generally, small-scale S. lividans cultures were grown in LB medium in baffled shake flasks at 30°C and 250 rpm. Flask volumes were limited to 10% of capacity. With pIJ702 or pIJ68O vectors, thiostrepton (Th) was included at 5 j.tg/mL. S. lividans transformants were selected on regeneration yeast extract agar (R2YE, Hopwood et at., 1985) overlaid with 0.3% nutrient agar supplemented with thiostrepton at 50 jg/mL. Solid media contained 1.5% agar. 20 Table 2.2. Plasmids, phagemids and phage Plasmid, phagemid or phage Relevant characteristics pUC12-1.lcex contains 2.5 kb C. fimi DNA; expresses O’Neill et at., 1986b cex from lac promoter; Apr pUCI2-1.lcex(PTIS) contains 1.8 kb C. fimi DNA; expresses O’Neill et at., 1986a cex from inc promoter; contains portable translation initiation site (PTIS); Apr pTZ18R contains Fl on; Apr Mead et at., 1986; Pharmacia pTZ18R-cex as above; contains 1.8 kb C. fimi DNA This study Streptomyces expression Hopwood et at., 1985 p1J702 vector; Th Reference or source p1J702-cex as above, contains entire pUC12-1.lcex This study plasmid; expresses cex from mel promoter p1J680 Streptomyces expression vector; Thr Hopwood et at., 1985 p1J680-cex as above, expresses cex from aph promoter; contains entire pUC12-1.lcex plasmid M13K07 helper phage for preparation of single- Vieira and Messing, 1987 stranded DNA; Kmr This study; MacLeod et at., 1992 21 Growth of E. coli was monitored spectrophotometrically by A600. Growth of Streptomyces was monitored as follows: Mycelium dry weight was determined by washing mycelial pellets in distilled water, followed by filtration through Millipore type HA filters (Bedford, Maryland). The cell mass was dried for about 4 days at 37°C. 2.4 Recombinant DNA techniques Most recombinant DNA techniques were carried out essentially as described by Sambrook et a!. (1989). Double-stranded plasmid DNA was isolated from E. coli by the small-scale alkaline lysis method (Sambrook et al., 1989). Plasmid DNA was isolated from Streptomyces, by the boiling method essentially as described by Kendall and Cullum (1984). Restriction endonuclease digestions and ligations were carried out according to the directions of the manufacturer& in the buffers provided. Electrophoresis of DNA generally employed 0.8% agarose gels in TBE buffer (Sambrook et al., 1989). DNA fragments excised from agarose gels were purified using the Gene Clean II Kit (BiolOl, La Jolla, Ca) according to the manufacturer’s directions. Streptomyces protoplasts were prepared and transformed as described by Hopwood et a!. (1985). Preparation and transformation of competent E. coil strains was performed as described by Hanahan (1983). 2.4.1 Production and isolation of single-stranded DNA. Single-stranded template DNA was obtained as follows: Single, overnight colonies of E. coli JM1O1 containing pTZ18R-based constructs were inoculated into 2mLs of TYP medium containing 100 tg/mL ampicillin and 10 PFU/mL M13K07 helper phage. Following 1 h incubation at 35°C, kanamycin was added to 70 j.tg/mL. 22 Cultures were grown overnight at 35°C and 225 rpm. Cells were removed by centrifugation in a microfuge for 5 mm at room temperature. The phagemid were precipitated at 4°C with 1.7 M ammonium acetate and 12% (w/v) PEG-6000. Single-stranded DNA was isolated from the phagemid as follows (D. Trimbur, personal communication). After centrifugation for 10 minutes at 4°C the supernatant was removed and the phage pellet was resuspended in 20 iL of 10 mM Tris-HC1, 1 mM EDTA pH 8.0 (TB). 200 pL of 4 M NaC1O 4 were added and the mixture was incubated for 5 mm at room temperature. Glass fibre filter disks (GF/C, Whatman) were placed in the bottom of a microtitre plate in which the wells had been pierced with a hot needle. The DNA was bound to the glass fibre by vacuum filtration through the microtitre plate. The bound DNA was washed thoroughly with 70% ethanol and air dried for 5 minutes. The filter disks were transferred to 0.5-mL Eppendorf tubes pierced at the bottom, which were in turn placed inside 1.5-mL Eppendorf tubes. The DNA was eluted from the filter disks with 20 iL of H 0 and captured by centrifugation 2 for 30 s. 2.4.2 Site-directed in vitro mutation of cex Site-directed in vitro mutation was performed essentially as described by Kunkel et al., 1987. Single-stranded, uracil-containing DNA templates (pTZ18R-cex) were produced as described in section 2.4.1 with the following modifications: E. coil RZ1032 was used as the host strain and uridine was included in the medium at 1 j.tg/mL. Synthetic oligodeoxyribonucleotide primers used for mutagenesis were prepared (Atkinson and Smith, 1984) by the UBC Nucleic acid and Protein synthesis unit (NAPS) with an Applied Biosystems 380A DNA synthesizer. Oligonucleotides were purified by precipitation with n-butanol as described by Sawadogo and Van Dyke (1991). The oligonucleotides (25mers) employed in this study are described in Table 2.3. 23 2.4.3 Sequencing of DNA Sequencing of plasmid DNA used the modified dideoxy chain termination method described previously (Tabor and Richardson, 1987) with the following modifications: in the primer extension reaction, T7 DNA polymerase (Sequenase) was used, the reaction temperature was increased to 43°C and 7-deazaGTP was substituted for dGTP. Table 2.3. Oligonucleotides used in site-directed mutagenesis of cex Oligonucleotide sequence* Mutation in Cex 5’P-G CGC ATC ACC GAC CTC GAC ATC CGC-Y. E233D 5’P-G CGC ATC ACC CAG CTC GAC ATC CGC-3’. E233Q 5P-AC GTC GTC AAC GAC GCG TTC CCC GA-3’ E127D 5’P-AC GTC GTC AAC CCG GCG TTC GCC GA-3’ E127A 5’P-AC GTC GTC AAC GGC GCG TTC CCC GA-3 E127G 5’P-TC GCG TCG TGG GCC GTC GTC AAC GAG-3’ D123A *P indicates the oligonucleotide is 5’ phosphorylated 24 2.5 Detection of cex gene expression 2.5.1 MUCase and PNPCase assays E. coli or S. lividans transformants expressing exoglucanase activity were detected as follows. Cells were plated onto LB agar containing the appropriate antibiotic and 100 jiM methylumbelliferyl 13-D- cellobioside (MUC). Following growth overnight at 30°C, colonies expressing exoglucanase activity fluoresced under UV light (365nm). This assay was also used as an initial screen for Cex mutants. Following sitedirected mutagenesis, non-fluorescing colonies, indicating loss of exoglucanase activity, were isolated and plasmid DNA was prepared for sequencing. Cex activity in crude-cell extracts or culture supernatants could be estimated by using the following assay (Gilkes et al., 1984b): Culture supernatant or diluted cell extract (0.5 mL) was added to 0.5 mL of 12.5 mM PNPC to give a final PNPC concentration of about 10 x Km as determined previously for Cex (Gilkes et a!., 1984b). The mixture was incubated at 37°C for about 15 mm or until a yellow colour appeared. The reaction was stopped by the addition of 1 M 3 CO (0.5 mL) and the absorbance 2 Na at 400 nm was determined. One unit of activity was defined as that which released 1 jimol PNP per minute (Gilkes et a!., 1984b). 2.5.2 Binding to Avicel When the activity of the enzyme was unknown (for example, with the Cex mutants), the levels of gene expression could be approximated from crude-cell extracts or culture supernatants as follows: Approximately 100 tL of cell extract (described in section 2.7.2) or 0.5 mL of culture supernatant was added to 20 mg of Avicel (microcrystalline cellulose) in a final volume of 1 mL 50 mM phosphate buffer, pH 7.0. The protein was adsorbed to the Avicel by mixing for 1 h at 4 °C. The Avicel was 25 centrifuged for 30 s, washed once with 500 pL IM NaC1 in phosphate buffer, then once with phosphate buffer. The Avicel was collected by centrifugation, then boiled for 2 mm in SDS loading buffer (Laemmli, 1970). The bound polypeptides were analyzed by SDS-PAGE (section 2.6.1). An estimation of mutant protein yield could be made by comparison with a known quantity of wild-type Cex. 2.6 Detection of protein 2.6.1 SDS-PAGE and Western blotting Proteins were resolved by 0.1% sodium dodecyl sulfate-10% polyacrylamide gel electrophoresis (SDS-PAGE) as described previously (Laemmli, 1970; Schägger and Von Jagow, 1987) using a Bio-Rad MiniPROTEANTM apparatus. Molecular weight standards (Sigma) were as follows (kDa): Myosin, 212; f3-galactosidase, 130; phosphorylase B, 97.4; bovine serum albumin, 68; catalase, 57; glutamate dehydrogenase, 53; alcohol dehydrogenase, 45; ovalbumin, 41; glyceraldehyde-3phosphate dehydrogenase, 36; carbonic anhydrase, 29; bovine trypsin inhibitor, 20; cytochrome C, 12.4. Protein bands were visualized by staining with Coomassie blue (Meril, 1990). Western blots were performed as described previously (Towbin et at., 1979). Western blots were probed with rabbit anti-Cex serum (1/4000 dilution) using goat anti-rabbit IgG-alkaline phosphatase conjugate (BRL) as the secondary antibody (Gilkes et at. , 1988) at 1/7000 dilution. The detection reagents were 5-bromo-4-chloro3-indolyl-phosphate and nitroblue tetrazolium dye (Sigma). Prestained molecular weight standards were from BRL or Bio-Rad. 26 2.6.2 Glycoprotein detection Glycoproteins in SDS-PAGE gels were detected with the periodic acid-Schiff reagent (Zacharius et a!., 1969). Glycoproteins were detected following western blotting by probing the blots with concanavalin A-horseradish peroxidase (ConA-HRP; Seikagaku). The blots were developed with HRP development reagent (Bio-Rad). 2.6.3 N-terminal amino acid sequencing N-terminal amino acid sequences were determined at the University of Victoria’s protein sequencing unit by automated Edman degradation, using an Applied Biosystems 470A gas-phase sequenator. 2.6.4 Determination of protein concentration Protein concentrations were determined by dye binding (Bradford, 1976) using the Bio-Rad protein kit, with bovine serum albumin as the standard. The concentration of purified Cex was determined spectrophotometrically from A280. The extinction coefficient of Cex 1 mg/mL 2 (E = 1.61) was determined experimentally by the method of Scopes (1974). 2.7 Production and purification of recombinant Cex 2.7.1 Production of Cex in Streptomyces lividans TK64 S. lividans TK64 (250 .iL from a frozen stock culture) containing the cex gene encoded on p1J702-cex or p1J680-cex was inoculated to 10 mL LB medium supplemented with 5 jig /mL thiostrepton. Cultures were incubated for 3 days at 30°C, then added to 100 mL of the same medium. After growth overnight at 30°C and 225 rpm, the culture was added to 10 L of the same medium in a Chemap FZ3000 fermenter. After growth for 45 h at 300 C, 600 rpm and 5 L/min aeration, the mycelium was removed by 27 passage of the culture through a Sharples centrifuge. Cex was purified from the culture supernatant as described in section 2.7.3. 2.7.2 Production of Cex and Cex mutants in E. coli E. coil containing the plasmid pUC12-Llcex(PTIS) encoding wild-type Cex or mutants of Cex were grown in 2L-cultures in shake flasks or in a 60 L Chemap FZ3000 fermenter. Cultures were grown in LB supplemented with ampicillin at 100 .ig/mL. At A600 = 1.0, IPTG was added to 0.1 mM and growth was allowed to continue overnight. The cells were collected by centrifugation at 10 000 x g for 10 mm at 4 °C. The cell pellet was resuspended in approximately 1/100 of the original volume in 50 mM phosphate buffer pH 7.0 (phosphate buffer) at 4 °C. Cells were ruptured by passage twice though a French pressure cell. PMSF was added to 0.5 mM, Pepstatin A to 1 iM and EDTA to 1 mM. The crude-cell extract was clarified by centrifugation at 40 000 x g for 30 mm at 4°C. Streptomycin sulfate was added to 1.5% and the mixture stirred at 4 °C overnight. The extract was centrifuged again at 40 000 x g for 30 mm at 4°C. Cex was purified from the clarified cell extract as described below. 2.7.3 Purification of Cex Culture supernatant from Streptomyces or clarified cell extract from E. coil was stirred with CF-I cellulose (Sigma) (about ig CF-1/mg Cex) in phosphate buffer for 3 h at 4 °C. The cellulose was recovered by vacuum filtration on a glass fibre filter disk (GF/C, Whatman), washed twice with 500 mL I M NaCl in phosphate buffer and twice with 500 mL phosphate buffer. The cellulose was packed into a column (5 cm X 25 cm) and attached to an FPLC system (Pharmacia). Washing was continued by passing a column volume of phosphate buffer through the column at a flow rate of 1 mL/min In the case of Cex produced in E. coli, adsorbed polypeptides were eluted 28 with distilled water at a flow rate of 1. mL/min. In the case of Cex produced in Streptomyces, adsorbed polypeptides were eluted with a linear gradient of 0-8.0 M guanidinium hydrochloride (Gdm HC1) in phosphate buffer, total volume 500 mL at a flow rate of 1 mL/min. In both cases, the absorbance of the eluate was measured continuously at 280 nm and peak fractions were assayed for Cex activity with PNPC (section 2.5.1) The appropriate fractions were pooled and centrifuged at 40 000 x g for 30 mm at 4°C to remove any cellofines. The supernatant was concentrated to greater than 1 mg protein/mL by diafiltration through an Amicon PM1O membrane. For Cex from Streptomyces, the Gdm HC1 was replaced with phosphate buffer during the diafiltration step such that the final concentration of Gdm HC1 was calculated to be less than 0.5 jiM. The purified polypeptides were analyzed by SDS PAGE (section 2.6.1). Final protein concentration was determined by A280 as described in section 2.6.4. Purified protein solutions were passed through a 0.22 jim filter (Millipore) and then stored at 4°C. 2.8 Proteolysis of Cex from S. lividans and E. coli 2.8.1 Substrate-bound Cex C. fimi protease, generously provided by Shen Hua, was prepared as described previously (Gilkes et a!., 1988). Purified Cex from S. lividans or from E. coli was adsorbed to 5 mg Avicel as described in section 2.5.2 then resuspended in 25 jiL 20 mM Tris.HC1, pH 7.5. After adding 0.5 units* of C. firni protease, the suspensions were incubated at 37°C for 24 h. The Avicel was recovered by centrifugation and bound polypeptides were analyzed by SDS-PAGE as in section 2.5.2. The supernatants, containing polypeptides released from the Avicel by protease treatment, were also analysed by SDS-PAGE. * (1 unit is defined as that which will release 1.0 0D /h @ 585 37°C, pH 7.5 from cowhide powder azure). 29 2.8.2 Cex in solution Purified Cex from E. coli, S. lividans and C. fimi, was incubated with C. fimi protease at 37°C for 24 h in 20 mM Tris.HC1, pH 7.5. The samples were analyzed by SDS-PAGE as above. 2.9 Enzyme kinetics 2.9.1 Determination of steady-state kinetic parameters Michaelis-Menten parameters for aryl cellobiosides and aryl glucosides were determined by continuous spectrophotometric measurement of the release of p nitrophenol, dinitrophenol or p-bromophenol using a Hitachi U 2000 spectrophotometer with a temperature-controlled cell holder. Reactions were carried out in 50 mM phosphate buffer (pH 7.0), 37°C and contained BSA at I mg/mL where indicated. Reaction mixtures were pre incubated within the water-circulated cell holder at the appropriate temperature for a period of 10 mm prior to addition of enzyme. Where indicated, sodium azide was included in reactions at the concentrations shown. For each substrate to be tested, the appropriate concentrations of substrate to be used was estimated by monitoring the initial rates of hydrolysis at three different substrate concentrations and then estimating the Km value from a Hanes plot. Initial rates of enzyme-catalysed hydrolysis for a particular substrate were measured at 6 to 10 different substrate concentrations ranging from about 1/7 x Km to 7 x K where practical. Incubation times ranged from 5 minutes to 18 hours (in the case of 4-BrPC). Values for Km and kcat were determined from the initial rate of hydrolysis ) vs. substrate concentration, by non-linear regression analysis using the computer 0 (V program Grafit 2.0 (Erithacus Software Ltd., Staines, U.K.). An example calculation of Km and kcat is presented in the Appendix. 30 The substrates used in this study, along with the wavelengths monitored and extinction coefficients for the corresponding aglycon units, are shown in Table 2.4. Table 2.4. Wavelengths monitored and extinction coefficients for hydrolysis of aryl cellobiosides and aryl glucosides. Substrate Aglycon unit Wavelength monitored (nm) Extinction coefficient (AE) .cm at given (M ) 1 wavelength, pH 7.0 PNPC PNP 400 7280 PNPG PNP 400 7280 2,4-DNPC DNP 400 10900 2,4-DNPG DNP 400 10900 4-BrPC PBrP 288 1 120 2.9.2 Kinetics of Inactivation of Cex mutants with 2F-DNPC. Inactivation of Cex mutants with 2F-DNPC was performed essentially as described previously for inactivation of wild-type Cex by 2F-DNPG (Tull et a!., 1991). The enzyme was incubated at 37°C in 50 mM phosphate buffer (pH 7.0) containing 2FDNPC at the concentrations indicated. At various times, 10 iiL aliquots of the inactivation mixture (as described above) was transferred into a solution of 2,4-DNPC (100 tM) and sodium azide (60 mM) in 50 mM phosphate buffer (pH 7.0; 37°C; final volume 500 j.iL). Enzyme activity was determined by continuous spectrophotometric measurement of the release of dinitrophenol using a Hitachi U 2000 spectrophotometer 31 with a temperature-controlled cell holder. For each inactivator concentration, residual enzyme activity (v/v ) at various time points was calculated by comparison with the 0 activity of the control sample which contained no inactivator. From the slope of a plot of in [V/Vol for each inactivator concentration, pseudo-first order rate constants (kobs) were calculated. The inactivation rate constant (kj) and the equilibrium binding constant (1(i) can be determined from a double-reciprocal plot (Lineweaver-Burk) of kobs vs. inactivator concentration. This curve is presented in the results section for display purposes and was not used for calculations. The inactivation rate constant (ki) and the equilibrium binding constant (Ki) were determined from kobs vs. inactivator concentration by non-linear regression analysis using the computer program Grafit. 2.10 Characterization of the products of enzymatic hydrolysis Products were characterized from reactions generally containing 100 j.Lg/mL mutant Cex or 1 j.tg/mL wild-type Cex and 6 mM substrate (as indicated) in phosphate buffer, incubated overnight at 37°C. Sodium azide was included where indicated. 2.10.1 Thin layer chromatography Thin layer chromatography was performed on 0.2 mm silica gel aluminum plates (#60 F254; E. Merck). Approximately 2 !IL of each reaction mixture or standard was applied to the base of the TLC plate. The plate was transferred to an Erlenmeyer flask and dried under vacuum for about 5 mm. The chromatography was performed with a mixture of ethyl acetate/methanol/water (7:2:1 v/v/v) and allowed to air dry for 5 mm. The chromatographs were dipped in a solution of 10% 4 S0 in methanol 2 H and heated with a Red Devil heat gun until the reaction products were visible. f3cellobiosyl-azide, f3-glucosyl-azide, cellobiose and glucose were the standards used. 32 2.10.2 1 H-NMR spectrometry of the products of enzymatic hydrolysis Enzymatic reactions were carried out as described above. Prior to H-NMR, the enzyme was removed by filtering the solution through a 10 kDa cutoff polysulfone membrane (ultrafree-MC; Millipore). Samples were prepared for 1 H-NMR analysis by Dr. Thisbe Lindhorst, Department of Chemistry, UBC, by repeatedly dissolving in 2 D 0 , freeze drying several times, and finally dissolving in 2 D 0 . Spectra were recorded with a Bruker 400 MHz spectrometer at the NMR Facility, Department of Chemistry, UBC, with the assistance of Dr. Thisbe Lindhorst. 2.11 Protein Molecular Mass determination Protein molecular mass was predicted from amino acid sequence data using the MacProMass computer program (Beckman Research Institute, Duarte, CA). Actual protein molecular mass was determined by Ion spray mass spectrometry using a PESCIEX API III ion spray LC/MS system by Dr. Shichang Miao, Department of Chemistry, UBC. Where indicated, the enzyme was incubated with substrate (2,4DNPC) for 1 mm at 37 °C prior to mass spectrometric analysis. 2.11 Amino acid sequence alignments and database searches Amino acid sequences were aligned on an IBM-compatible computer using PCGENETM (Intelligenetics, Mountainview, CA). Amino acid sequence information for family-F cellulases and xylanases was obtained through electronic mail ([email protected]) from the GenBank and Swiss-Prot databases at the National Center of Biotechnology Information (NCBI), National Library of Medicine, NIH, in Bethesda, MD. 33 3. Results (part 1): Expression of the cex gene in S. lividans 3.1 Construction of p1J702-cex and p1J680-cex. Plasmids for expression of cex in Streptomyces were constructed as outlined in Figures 3.1 and 3.2, respectively. In p1J702-cex, the cex gene and its promoter are situated downstream from the tyrosinase or “melanin” (mel) promoter of p1J7O2. In p1J680-cex, the cex gene and its promoter are situated downstream from the S. fradiae aminoglycoside phosphotransferase (a ph) promoter of p1J680. A construct was also obtained in which the C. fimi DNA was inserted in the reverse orientation (p1J702-xec, not shown). The constructs p1J702-cex and p1J680-cex, contained pUC12-1.lcex in its entirety permitting replication in both Streptomyces and E. coil. E. coli JM1O1 was transformed with the ligation mixture and clones were selected on LB medium supplemented with ampicillin. The desired constructs were isolated from E. coil by small-scale alkaline lysis (Sambrook et a!., 1989) and then used to transform S. lividans TK64 protoplasts as described by Hopwood et a!. (1985). Streptomyces transformants were selected on R2YE solid medium supplemented with thiostrepton and MUC. Transformants hydrolysing MUC (expressing exoglucanase activity) fluoresced under UV light. The presence of p1J702-cex or p1J680-cex was confirmed by analysing plasmid DNA (isolated as described in section 2.4) by agarose gel electrophoresis. 3.2 Production of Cex by S. lividans TK64 [p1J702-cex] As indicated in Figure 3.3, exoglucanase (PNPCase) activity was secreted into the culture supernatant by S. lividans TK64 containing p1J702-cex. Exoglucanase activity in the culture supernatant peaked at 0.008 units/mL culture following about 45 hours of growth. This represented less than 1 mg of active Cex per litre of culture, based on the 34 Figure 3.1. Construction of p1J702-cex. p1J702 (Hopwood et a!., 1985) was isolated from S. lividans TK24 by the method of Kendall and Cullum (1984). pUC12-1.lcex, containing the exoglucanase (cex) gene from Cellulornonasfirni, was described previously (O’Neill et a!., 1986a). Plasmid p1J702-cex (11.1 kb) was constructed by ligating Barn HI cut pUC121.lcex (5.3 kb) into the Bgl II site of p1J702 (5.8 kb). This resulted in the destruction of both sites (*) 35 BamHl 89111 Barn Hi BgI II T4 DNA ligase 36 Figure 3.2. Construction of p1J680-cex. p1J680 (Hopwood et al., 1985) was isolated from S. lividans TK24 by the method of Kendall and Cullum (1984). Plasmid p1J680-cex (10.6 kb) was constructed by ligating Barn HI cut pUC12-1.lcex(5.3 kb) into the Barn HI site of p1J680 (5.3 kb). 37 BamHI BamHI BamHI T4 DNA ligase 38 0.010 -J E D > 4-, > 4-, C) c ci U) 0.004 C) D x w 0.002 0.000 20 40 60 80 100 120 Time (h) Figure 3.3. Production of exoglucanase by S. lividans TK64[p1J702-Cexl. S. lividans TK64 containing plasmids p1J702-cex (0), p1J702-xec () or p1J702(D) was grown in 100 mL of LB supplemented with thiostrepton (5 jig /mL) in a 1-L baffled shake flask at 30°C, 225 rpm. The initial cultures had an A600 = 0.01. Samples of 5 mL were withdrawn at approximately 3-h intervals, and culture supernatants were collected by centrifugation for 7 mm at 3000 x g. Exoglucanase activity was determined from culture supernatants as described in section 2.5.1. 39 specific activity of the wild-type enzyme. Production of exoglucanase (PNPCase) activity in S. lividans TK64 [p1J702-xec], in which the C. firni fragment (including the cex gene and its promoter) was inserted in the reverse orientation, was only about 10% of that containing p1J702-cex. Only trace amounts of exoglucanase activity could be detected from culture supernatants of S. lividans TK64 containing p1J702 without an insert. 3.3 Production of Cex by S. lividans TK64 [p1J680-cexl Exoglucanase (PNPCase) activity was secreted into the culture medium by S. lividans TK64 [p1J680-cexl as indicated in Figure 3.4. Maximum activity in the culture supernatant (0.06 units/mL) was reached after about 40 h of growth. Based on activity, this represents about 5.5 mg of Cex per litre of culture, an 8-fold increase over that produced by S. lividans TK64 [p1J702-cex]. As a fraction of total protein present in the culture medium, the activity peaked at 3.3 units/mg (about 14% of total secreted protein) following 35 h of growth. This value subsequently declined because of additional protein appearing in the culture medium during the late log and stationary phases of growth. As shown in Figure 3.5A, p1J680-cex encoded a 49 kDa polypeptide (p ) which 49 was the major polypeptide in the culture supernatant. The proportion of p49 in the culture supernatant increased with time, consistent with an increase in exoglucanase activity with time as indicated in Figure 3.4. The identification of p49 as Cex was confirmed by its reaction with rabbit anti-Cex as indicated in Figure 3.5B. The 49 kDa polypeptide could be purified from the culture supernatant by adsorption to Avicel (Figure 3.5A, lanes 6 and 7) indicating the presence of a functional cellulose-binding 0.060 0.070 0.000 10 0.010 0.020 0.030 0.040 20 30 50 Time (h) 40 60 70 80 0 90 v 30 a) —S C 0.0 n I-v 2.0 ‘)ñ J.u 4.0 _j . > ca I— 0 -I -E 0 + a) —S Figure 3.4. Production of exoglucanase by S. lividans TK64 [p1J680-cexJ. S. lividans TK64 [p1J680-cexl was grown in 100 mL of LB supplemented with 5 .tg Th/mL in a 1-L baffled shake flask at 30°C, 225 rpm. The initial culture had an A600 = 0.01. Samples of 5 mL were withdrawn at approximately 3-h intervals, and culture supernatants were collected by centrifugation for 7 mm at 3000 X g. Exoglucanase activities were determined from culture supernatants as described in section 2.5.1. Protein concentrations were determined by dye-binding as described in section 2.6.4. Mycelium dry weight measurements were determined as described in section 2.3. Cl) a) ct C.) > •.. 0.050 0 C 41 Figure. 3.5. Identification of extracellular Cex in S. lividans [p1J680-cex] cultures. S. lividans cultures were grown as described in Figure 3.4. Samples of culture were removed at various time intervals, the mycelium removed by centrifugation and 500 jiL of the supernatants made 10 % in trichioroacetic acid. The precipitated proteins were collected by centrifugation and analysed by SDS-PAGE. A further 500 jiL sample of the supernatant was shaken with 12.5 mg Avicel (microcrystalline cellulose) for 1 h at 4°C. Polypeptides which adsorbed to the Avicel were analyzed by SDS-PAGE as described in section 2.5.2. Panel A: Gel stained with Coomassie blue. Panel B: Western blot (samples prepared as above) probed with rabbit anti-Cex serum (see section 2.6.1). Lanes: 1, molecular weight standards as indicated (kDa); lanes 2 and 8, 2 jig and 0.5 p.g Cex purified from E. coli (section 2.7.3); lanes 3 and 9, supernatant from S. lividans TK64 after 4 h growth; lanes 4 and 10, supernatant from S. lividans TK64 [p1J680-cexl after 20 h growth; lanes 5 and 11, supernatant from S. lividans TK64 [p1J680-cexl after 40 h growth; lanes 6 and 12, unbound polypeptides after treatment of the 40 h supernatant from S. lividans TK64 [p1J680-cex] with Avicel; lanes 7 and 13, bound polypeptides after treatment of the 40 h supernatant from S. lividans TK64 [p1J680-cex] with Avicel. (‘4 CD to I hUT 1J 1’IH C4JO o LOC)LO,—CD CDF0C) CDLO C) 0 0) 43 domain. The other major polypeptide (p42) present in the culture supernant (lanes 4 and 5), reacted with rabbit anti-Cex (lanes 10 and 11), but did not adsorb to Avicel (lane 6). Therefore, this polypeptide is probably a degradation product of Cex which is lacking a functional cellulose-binding domain. Cex produced from S. lividans had an apparent molecular mass 2 kDa more than that of the polypeptide produced from E. coli. Determination of the N-terminal amino acid sequence of the polypeptide from S. lividans gave two sequences in approximately equal proportions (section 2.6.3), Ala-Thr-Thr-Leu-Lys (ATTLK) and Gln-Ala-Ala-Thr Thr (QAATT), indicative of leader peptide processing at two adjacent sites, i.e., between A(-1) and A(+1) and between A(-3) and Q(-2) respectively (Table 3.1). In E. coli and in C. fimi, Cex is processed at a single site between A(-1) and A(+1) (O’Neill et a!. 1986c). , The additional amino acids were insufficient to account for the difference in the apparent molecular masses. Table 3.1. N-terminal processing of native and recombinant Cex Source of Cex Site(s) of processing : 1 -3 -2 -1 ÷1 +2 +3 +4 ÷5 2 E.coliJMlOl 2 C.fimi S. lividans TK64 A+A T T L K. A+A T T L K. A+A T T L K A+QAATTLK I Determined by N-terminal (5-cycle) amino acid sequencing. A (+1) represents the N terminal amino acid in Cex from C. firni. 2 The N-terminal sequence was determined previously (O’Neill et a!., 1986c). 44 3.4. Glycosylation of Cex by Streptomyces lividans It was reported previously that the difference in size between Cex from C. fimi and that from E. coli is a consequence of glycosylation of the native enzyme (Langsford et al., 1984; Gilkes et a!., 1988). The enzyme produced in E. coli is not glycosylated. Although there were no reports of the glycosylation of a heterologous polypeptide by Streptomyces prior to this study, Streptornyces was known to produce glycoproteins (Mihoc and Kluepfel, 1990). It appeared reasonable that the difference in size between the E. coli and S. lividans-produced enzymes were the result of glycosylation of the latter. Cex produced by S. lividans [p1J680-cexl and E. coli JMIOI [pUCI2-1.lcex] was purified to apparent homogeneity by affinity chromatography on cellulose (Fig. 3.6) as described in Materials and Methods (section 2.7.3). Purified Cex from C. firni was generously provided by Emily Kwan. The specific activity of the purified enzyme from Streptomyces was 11.2 U/mg. The specific activities of Cex purified from E. coli and C. fimi were similar: 12.0 U/mg and 13.1 U/mg, respectively. As shown in Figure 3.6A, the recombinant polypeptide from S. lividans appeared to be the same molecular mass as the native, glycosylated polypeptide from C. fimi (49 kDa) as judged by SDS-PAGE. It was slightly greater than that from E. coli (47 kDa). As revealed by Schiff’s staining (Figure 3.6B) and by reaction with Concanavalin-HRP (Figure 3.6C), Cex produced by S. lividans was also glycosylated. It was determined previously that the glycosyl groups on Cex from C. fimi protect it from a protease present in C. fimi culture supernatants if Cex is adsorbed to cellulose, and to a lesser extent when the enzyme is in solution (Langsford et a!., 1987; Gilkes, N.R., unpublished). In contrast, non-glycosylated Cex from E. coli is susceptible to proteolysis whether or not the enzyme is bound to cellulose. This yields a major 2 3. 4 567 B 1 8 Figure 3.6. Glycosylation of Cex by S. lividans. Cex produced by S. lividans [p1J680-cex], E. coli JM1O1 [pUC121.lcex] and C. fimi were purified to apparent homogeneity by affinity chromatography on cellulose (section 2.7.3). The purified polypeptides were analyzed by SDS-PAGE. Panel A: gel stained with Coomassie blue. Panel , gel stained by the periodic acid-Schiff method to detect glycosylation. Panel , Western blot probed with concanavalin A-horseradish peroxidase to detect glycosylation. Lane 1: size standards as indicated. Lanes 2, 5 and 8,2 j.tg Cex purified from E. coli; lanes 3, 6 and 9, 2 .tg Cex purified from C. fimi; lanes 4, 7 and 10, 2 p.g Cex purified from S. lividans. 41’ 36’ 29’ 97.4 S. 68.0%.. 57.5 53..- 1 kDa 212\ A Ui 46 proteolysis product of 35 kDa corresponding to the catalytic domain of Cex (Langsford et al., 1987; Gilkes et a!., 1991a). In order to determine whether the glycosyl groups on the enzyme from Streptomyces were functionally similar in this regard to those of the native enzyme, its susceptibility to the C. fimi protease was analysed. As shown in Figure 3.7A and 3.7B, the Streptomyces-produced enzyme behaved identically to the native enzyme. Unlike the enzyme produced in E. coli, it was not cleaved when adsorbed to cellulose (Figure 3.7A). The Streptomyces-produced enzyme was hydrolyzed when in solution (Figure 3.7B), although more slowly than non-glycosylated Cex from E. coli. Cex from S. lividans gave a major proteolysis product of 42 kDa, similar to that of Cex from C. fimi. 47 Figure 3.7. Sensitivity of Cex from S. lividans to the protease from C. fimi. Panel A: Cex adsorbed to cellulose (Avicel). Purified Cex (20 tg) from S. lividans and from E. coli was each adsorbed to 5 mg Avicel, then resuspended in 20 mM Tris.HC1, pH 7.5. The suspension was incubated with 0.5 units of C. firni protease at 37°C for 24 h. Polypeptides bound to the Avicel were analysed by SDS-PAGE. The supernatants, containing polypeptides released from the Avicel by protease treatment, were also analysed by SDS-PAGE. Lane 1, molecular weight standards; lanes 2-5, Cex from S. lividans; lanes 6-9, Cex from E. coli; lanes 2 and 6, bound polypeptides left after protease treatment; lanes 3 and 7, polypeptides released from the Avicel by protease treatment; lanes 4 and 8, bound polypeptides without protease treatment; lanes 5 and 9, unadsorbed polypeptides without protease treatment. Panel : Cex treated in solution with C. fimi protease. Purified Cex (10 .tg) from E. coli, S. lividans and C. fimi, was incubated with 0.5 units of C. fimi protease at 37°C for 24 h in 35 .tL 20 mM Tris.HC1, pH 7.5. 10 iL of the S. lividans and E. coli samples and all of the C. fimi sample was analyzed by 0.1% SDS-10% PAGE. Lanes 10 and 17, molecular weight standards; lanes 11 and 12, Cex from S. lividans; lanes 13 and 14, Cex from E. coli; lanes 15 and 16, Cex from C.firni. Lanes 11, 13 and 15, with protease treatment; lanes 12, 14 and 16 without protease treatment. 48 A kDa 1 2 3 4 5 6 7 8 9 16 17 130 97.4 — 29’ kDa B 130 97.4 10 12 13 14 — Z — 53 11 kDa 130 — A. = — -53 .- — — — .— 36 / 29” 15 — —97.4 68.O —575 41 36 49 4. Results (part 2): Identification of catalytic residues in Cex 3-1,4-glycanases can be grouped into at least 12 families on the basis of similar amino acid sequences (Henrissat et at., 1989; Beguin, 1990; Gilkes et at., 1991b; Henrissat, 1991, Shen et a!., 1994). Cex belongs to family F of 3-1,4-g1ycanases (Henrissat et a!., 1989). Searches of the GenBank database revealed at least 20 enzymes that can now be assigned to the family (see Introduction, Table 1.1) Some of these enzymes contain more than one catalytic domain; some contain carbohydrate binding domains unrelated to that of Cex. To date, three-dimensional information is not available for any enzymes in family-F. In order to pinpoint potential catalytic residues, the enzymes were aligned to reveal conserved amino acid residues, The alignment was made with family-F catalytic domains related to the catalytic domain of Cex. As shown in Figure 4.1, there are several conserved acidic residues. Three aspartates and three glutamates are conserved in all members of the family, corresponding to Glu 43, Asp 123, Glu 127, Asp 170, Glu 233, and Asp 277 in Cex. In addition, Asp 235 is conserved in all but one member of the family. 4.1 Wild-type Cex: kinetics Cex was purified from cultures of E. coli JM1OI LpUC12-1.1cex] as described in sections 2.7.2 and 2.7.3. The purified protein obtained by affinity chromatography on cellulose was >95% homogeneous when analyzed by SDS-PAGE (not shown). Approximately 30 mg of purified Cex was obtained from a 2-L culture. The kinetic parameters were determined for hydrolysis of various cellobiosides and glucosides (Table 4.1). The kcat values for p-nitrophenyl 13-D- cellobioside (PNPC), and 2”,4”-dinitrophenyl -D-ce1lobioside (2,4-DNPC) and t ,2 4’-dinitrophenyl 13-D- “---“ “_“ Figure 4.1. Alignment of Family F catalytic domains. Only a portion of each enzyme is indicated. The putative catalytic nucleophile in Cex is indicated by an “n”. The putative acid/base catalyst is indicated by an “a/b”. Conserved acidic residues are indicated in bold-face. means the amino acid sequence is contiguous on either side of the gap. means the end of the sequence or that sequence information is incomplete. Accession numbers (Genbank or SWISS-PROT) are indicated in parenthesis. a) C. fimi Cex (L11080); b) S. lividans XynA (M64551); c) C. thermocellum XynZ (M22624, P10478); d) A. kawachii XynA (D14847); e) T. aurantiacus XYN [partial sequence only] (P23360); f) P. chrysogenum Xyn [partial sequence only] (S31307); g) R. flavefaciens XynA (P29126); h) B. fibrisolvens XynA (P23551); i) B. fibrisolvens XynB (X61495, S55274); j) C. saccharolyticum ORF4 (M34459); k) Thermophilic bacterial sp rt8.84 XynA (L18965); 1) C. saccharolyticum Ce1B (A43802, X13602); m) C. thermocellum XynX (M67438); n) Thermoanaerobacter saccharolyticum (strain B6A-RI) Xy1A (M97882); o) Bacillus sp strain C-125 XynA (D00087, P07528); p) C. stercorarium strain F9 (D12504); q) C. saccharolyticum XynA (M34459); r) P. fluorescens subsp-cellulosa XynB (P23030); s) C. albidus XynA (JS0734); t) P. fluorescens subsp-cellulosa XynA (X15429). t r s p q c m n o k 1 j b d e f g h i a t r p q m n o k 1 j f g h i e d c b a - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - + + - - - 181 VASWDVVNEAFADG_DGPP QDSAFQQKLGNG YIETAFRAARAA_DPTAXLCINDYNVEGINAKSN IVQWDVVNEAFADGSSGAR RDSNLQNVIGQD YLDYAFRYAREA_DPDALLFYNDYNIEDLG_PKSN IYAWDWNEIF NEDGSLRDSVFYKVIGDD YVRIAFETARAA_DPNAKLYINDYNLDSASYPKLA AFAWDVVNEAF NITGRL EVAAASRTDPNAKLYINDYNLDSARYPKTQ VMKNHITTVMKQYKGK LYAWDWNEIF EEDGTLRDSVFSRVLGED FVRIAFETAREA_DPEAKLYINDYNLDSATSAKLQ RLESMIKNTFAALKSQYPNLD_VYSYDVCNELFLNNGGGMRGAD NSNWVKIYGDDSFVINAFKYARQY APAGCKLYLNDYNEYI P_AKTN RLEFYVKSVNGHFYSGKTGST_LVYWDVCNETL HAp NSGWEAVYGSNKTNAVYVKKAFNYAYQVLEQYKLTNSVKLFYNDYNTYMEVN RLESYIHGVLDFVQTNYPGI_IYAWDVVNE_IVDEGA_FRK SIWTETVGED FFIKAFEFARKY AAPEVSLFYNDYETAQP_WKRD RLESYIKQVIEFCQKNYPGV_VYCWDVVNEAILDDGS_WREI NNNWYTINKEK YVEKAFYYARKY AKKDVALFYNDYNVFL P_AKRE RLKKHIQTVVGRYKGK VYAWDWNEAIDE NQPDGYRRSDWYNILGPE YIEKAFIWAHEAD_PKAKLFYNDYSTEDP_Y_KRE RLKQYIYDVVGRYKGK VYAWDWNEAIDE NQPDSYRRSTWYEICRSGND_WIEVAFTRARAAD_PSAELCYNDYNVENWTWAKTQ VMKNHITTVMTHYKGK IVEWDVANECMDDSGNG LRSSIWRGPE YIEKAFIWAHEAD_PNAICLFYNDYNTEIS_K_KRD RLKTHITTVLDHFKTKYGAQNPIIQWDWNEVLD DNGSLRNSKWLQI IGPD YIEKAFEYAHEAD_PSMKLFINDYNIENNGV_KTQ RLKTHITTVLDHFKTKYGSQNPIIGWDVVNEVLD DNGNLRNSKWLQIIGPD YIEKAFEYAHEAD_PSNKLFINDYNIENNGV_KTQ RMENHIKTVVERYK DDVTSWDVVNEVID DGGGLRESEWYQITGTD YIKVAFETARKYG_GEEAELYINDYNTEVP_SJCRD RLENYIRAVVLRYK NDPGGMRNS PWYQITGTE DDIKSWDVVNEVIEP YIEVAFRATREAG_GSDIKLYINDYNTDDP_V_KRD RLREHIKTLCERYK DVVYAWDVVNEAVED KTEKLLRESNWRKIIGDD YIKIAFEIAREYAG_DAXLFYNDYNNEMP_Y_KLE -QWIRDYCARYPDT NIDVVNEAVPGHQPAGYAQRAFGNNWIQRV FQLARQYC-- PNSILILNDYNNIR- -WQHN VLKNHIDNVIGRYKDD LAYFDIVNE PL YIETALRYAHE_VAPKIvIKLCINDYNIETVN_AKSQ NENGTYIKSNVWYNVGLES DFARHIDTVAAJ-{FAGO VKSWDWNEALFDSADDPDGRGSANGYRQSVFYRQFGGP EYDEAFRRAPRA_DPTAELYYNDFNTEE_NGAKTT 103 AMVNHVTKVADHFEGK 4IDHINGVMàHYKGK VJxIKNHITTVMQHYKGK 123 127 (a/b) 32 102 IADSEFNLVVAENANKWD ATEPSQN_SFS FGAGDRVASYAADTGI<ELYGHTLVWHS_QLPDWA KNLN_G_SAFES IAGREFNMVTAENEMKID ATEPQRG_QFN FSSADRVYNWAVQNGIKQVRGHTLAWIIS_QQPGWM QSLS_G_RPLQ_ ILQREFSMVVCENEMKFD ALQPRQN_VFD FSKGDQLLAFAERNGMQMRGHTLIWHN_QNPSWL TNGNWNRDSLLA VIKADFGALTPENSMKWD ATEPSRG_QFS FSGSDYLVNFAQSNNKLIRGHTLVWFIS_QLPSWV QAIT_DKNTLIE IIQADFGQVTPENRMKWD ATE PSQG_NFN FAGADYLVNWAQQNGKLIYGHTLVWWS_QLPPWV VSIT_DI<_____ IIKANFGQLSPENSMKWD ATEPSQG_QFS FAGSDYFVEFAETNGKLIRGHTLVWHS_QLPSWV SSIT_D1CTTLTD FLKI{HYNSITPENELKPES ILDQGACQQKGNNVNTQ ISLSRAAQTLKFCEQNGIALRGHTFVWYS_QTPDWFFRENFSQNG_AYVS KDIMNQ DEAKRLGYYIPSNYIKERWPK IDFRTVDEAVKICYENGLKMRGHTLVWHS_QTPTWLFRENYSGNG_RFVN TATMDA LLAEQFNSFTCENDMKPMYYLDREANK_KDPEKYNLS PALTFENAIPYLEFAKDNKIANRGHTLVWHN_QTPKWFFCERYNENF_PMAD RETILA MKQQYLLDYEATh_ASK_NGMPVCKFDSCIPALQFCKENGIKMRGI-IVLVWHN_QTPEWFFHKDYDVSK_PLVD AATNAR VIKRHFNSITPENEMKPESL QPYEG_GFS FS IADEYVDFCKKDNISLRGHTLVWHQQTPSWFFTN_PETGEKL TNSEKDKEILLD MVLK}-IFNSITAENEMKPESLL_AGQTSTGLSYR FSTADAFVDFASTNKIGIRGHTLVWHN_QTPDWFFKD_S_NGQRL S KDALLA LTAKHFNNLVAENANKPESL QPTEG_NFT FDNADRIVDYAIAHNNKMRGHTLLWHN_QVPDWFFQD_P SDPTK_PASRDLLLQ LTAKHFNNLVAENANKPESL QPTEG_NFT FDNADKIVDYAIAHNMKMRGHTLLWHN_QVPDWFFQD_P SDPSK_SASRDLLLQ ILKHHYNSLVAENAMKPESL QPREG_EWN WEGADKIVEFARKHNMELRFHTLVWHS_QVPEWFFID_EDGNRMVDETDPDKREANKQLLLE LYKKHVNMLVAENAMKPASL QPTEG_NFQ WADADRIVQF(ENGMELRFRTLVWHN_QTPTGFSLQKEGKPMVEETDPQKREENRKLLLQ ILLKHFNSLTPENNKFENI HPEEQ_RYN FEEVARIKEFAIKNDMKLRGHTFVWHN_QTPGWVFLDKNGE EASKELVIE ---RYWNQITPBNESKWGSV EGTRNVYNWAPLDRIYAYARQNNIPVKAHTFVW_GAQSPSWL ILESQFDAITPENEMKWE VVEPTEGNFD FTGTDKIVAEAKKTGSLLRGHNICWDS_QLTPAYV TSIT_DPTKLKK IVRAEFNQITAENIMKM SYMYSGS_NFS FTNSDRLVSWAAQNGQTVHGHALVWH PSYQLPNWA SD_S_N_ANFRQ s t m n o p q r k 1 j I h g f e d c b a t m n o p q r s k 1 j h 1 g f e d c b a - - - - - - - - - - - - -________ QVDDYYT QKARYKE SVSACLG_NDLCPGVSIWQFADPTSW IVQAYLEVVPPGRRGGITVWGIADPDSWjYTHQNLPDWPLLFFNDNLQPKPAYQGVVEALSG - - - -- - - - -- - - 314 VVQACMQVT_RCQGVTVWGITDKYSWVPDVFPGEGAALVWDASThJ(KPAYAAVMEAFGA VTNVCLAVSRCLGITVWGVRDSDSW RSEQTPLLFNNDGSKKAAYTAVLDALNG LMKICLA_NPNCNTFVMWGFTDKYTWIPGTFPGYGNPLIYDSNYNPKPAYNAIKEALMG VVEACLQQPKCIGITVWGVADPDSW TSTDYVD VVNACLQQPKCVGITVWGVADPDSW QADLYEKIF LANQNSAQIPAVTIWGTQDTVSWRSS K QNP LLFSAGYQPKPAY LNNYAYRLF KNI_NAAKKNGGNISCITWWGPSDAETWIRN EKP LIWSNIGVAKPAY LATRYQEFF QTYL_DAKKSGKANITSVTFWNLLDENSWLSG_FRRETSYP LVFKGKCEAKEAYYAVLKA-QADRYYEMM YLWDKNCNPKPCFYSFLQAKLLLKEDTDNGGPCNITCVTVFGICDDYPLYKN_FK2CM QAQKLKAIFDVLKKYRNVV TSVTFWGLKDDYSWLRG DMPLLSDKDYQPKFAFWSLID--QSQKYKEIFTMLKKYKNVV KSVTFWGLKDDYSWLRS_FYGKN DWPLL_FFEDYSAKPAYWAVIEQARLYKQLFDLFKAEKQYI TAVVFWGVSDDVTWLS KPNAPLL_FDSKLQAKPAYWAIAD--QARLYEQLFDLFKAEKQYI TAVVFWGVSDDVTWLS KPNAPLL_FDSKLQAKPAFWAVVD--QADRYDQLFELYEELAADI SSVTFWGIADNHTWLDGRAREYNNGVGIDAPFV_FDHNYRVKPAYWRIIDQAKRYQELFDALKENKDIV SAVVFWGISDKYSWLNG_FPVKRTN APLL_FDRNFMPKPAFWAIVDQAKVYEDVFAVFREYKDVI TSVTLWGISDRHTWKDN$PVKQRKDWPLL_FDVNGKPKEALYRI 250 QAADYKK PASTYAN QANNYKE SSTDYVE 182 249 SLY_DLVKDFKARGVPLDCVGF QSHLIVG_QVPGDFRQNLQRFADL_GVDVRITELDIRMRTPSDATK LAT -AMY_NMVRDFI{QRGVPIDCVGF QSHFNSG_SP_YNSNFRTTLQNFAAL_GVDVAITELDIQ GA -AVFNMIKSMI<ERGVPIDGVGF QCHFINGMSPEYLASIDQNIKRYAEI_GVIVSFTEIDIRI PQSEN PATAFQV -GMVSHVKKWIEAGIPIDGIG SQTH_LSGGAG ISGALNALAGA_GTKEIAVTELDI AGA -AIVNRVKQWCAAGVPIIGIG NQTARAA -GMV_SHVKKWIAAGVPIDGIG SQTH_LGAG AGAAASGALNALASA_GTEEVAVTELDI AGA -DIY_NMANKLI{QL_GYIDGIGM QSH_LATNYP DANTYETALKKFLST_GLEVQITELDITC TNS AE -DV IKLV NYINQGKKVCAGVGNQSH_LGTGFP SVDYYTNALNSFLRA_GFEVQITELDITN KGD YD -FILEKVLGPLIDK_KLIDGMGM QSJ-1_LLMDHP DISEYRTALEMYGST_GLQIHITELDMH NADPSEESMHA -AIY_NLAQKLKEK_GLIDGLGL QPT_VGLNYPELDSDDIDSFKTTLETFAKL_GLQIHITELNFEI KGDESNRTP_ENLKK -FIY_KLIKNLKAKGVPVHGVGL QCI-I_ISLDWPD VSEIEETVKLFSRI PGLEIHFTEIDISIAKNMTDDDAYN RYLLVQ -FIY_NMVKNLKSKGIPIHGIGM QCHINVNWPS VSEIENSI_LFSSIP_GIEIHITELDMSL YNYGSSENYSTPPQDLLQK -AMY_DLVKKLKSEGVPISGIGM QMH_INI_NSN IDNIKASIEKLASL_GVEIQVTELDMNN_flG NVSNEALLK -AMY_DLVKI<LKSEGVPIDGIGM QMH_INI_NSN IDNIKASIEKLASL_GVEIQVTELDMNM_NG NISNEALLIK -DLY_NLVKDLLEQGVPIDGVGH QSH.JQIGWPS IEDTRASFEKFTSL_GLDNQVTELDMSL_YGWPPTGAYT SYDDIPAELLQA -ILYELVKNLLEKGVPIDGVGH QTH_IDIYNPP VERIIESIKKFAGL_GLDNIITELDMSIYSWNDRSDYG DSIPDYILTL -IKTY_KVLKELLERGTPIDGIGI QAH_WNIWDIKNLVSNLKKAIEVYASL_GLEIHITELDISV_FEFEDKRT_D LFEPTP_EMLEL -EFIALAKAQGNYIDAVGL QAHELIKGMTAAQVXTAIDNIWNQVGKPIYISEYDIGDTNQVQLQNFQAHFPVF-AMA_IKVAAGLLAKGAPLHCIGMFKNAKRRSSGLLIRTASSGLESHFIGGSTPKDIPAAMNLFSD_QGLEVPMTELDVRIP VNGNDM_PANATVAKE -ALV_NLVQRLLNNGVPIDGVG FQMHVMND_YPS IANIRQANQKIVALSPTLKIKITELDVRLNNPYDGNSSNDYTNRNDC_AVSCAGLDR + 233 (n) 01 53 Table 4.1. Kinetic parameters for hydrolysis of cellobiosides and glucosides by wild-type Cex 1 Enzyme Substrate ) 1 kcat (miir Km (mM) kcat/Km ) 1 1 .mM (min Wild-type DNPG DNPC PNPC PNPG 860 419 677 1.4 1.87 0.06 0.53 8.33 460 6983 1278 0.17 Parameters were determined at 37°C in 50 mM phosphate buffer, pH 7.0. 1 54 glucoside (2,4-DNPG) were within the same order of magnitude at 677 min , 419 miiv 1 1 and 860 mm 4 respectively. The ‘<cat value for p-nitrophenyl 13-D-glucoside (PNPG), 1.4 , was much lower. The Km values were lowest for the cellobiosides, 0.06 mM for 4 mm 2,4-DNPC and 0.53 mM for PNPC, whereas those for the glucosides were much higher: 1.87 mM for DNPG and 8.33 mM for PNPG. The kcat/Km values were 3 to 4 orders of magnitude greater for the cellobiosides than they were for the glucosides. The hydrolysis of 13-1,4-glycans by a retaining enzyme involves a two-step process: the formation (glycosylation) and hydrolysis (deglycosylation) of an x-D glycopyranosyl-enzyme intermediate (see Introduction, Figure 1.4). The rate of the deglycosylation step in the reaction will be identical for substrates containing the same sugar residue(s) since the aglycon unit is no longer present. It was previously determined (Tull and Withers, 1994) that the rate-limiting step is glycosylation for the glucosides and deglycosylation for the cellobiosides. The results of this study are consistent with the findings of Tull and Withers (1994). The cellobiosides, with aglycon units of quite different leaving group ability, were hydrolysed at essentially the same rates suggesting that deglycosylation is rate-limiting. In contrast, the leaving group ability of the aglycon unit on the glucosides had a dramatic impact on the rate of hydrolysis, suggesting that glycosylation is rate-limiting for the glucosides. 4.2 The putative catalytic nucleophile in Cex: Glu 233 The nucleophile in a retaining 3-1,4-glycanase can be identified using an activated 2-deoxy-2-fluoro glycoside mechanism-based inactivator which functions by forming a highly stabilized intermediate in which the sugar is covalently bonded to the catalytic nucleophile. Using the inactivator 2-deoxy-2-fluoro J3-D glucoside (2F-DNPG), G1u358 was identified as the catalytic nucleophile in an Agrobcicterium 13-glucosidase 55 (Abg) (Withers and Street, 1988; Withers et al, 1990) a retaining f3-glucanase from another family of enzymes. When the active site nucleophile was targeted by sitedirected in vitro mutagenesis, Abg retained measurable activity upon mutation of G1u358 to Asp, but virtually no activity upon mutation to Asn or Gin. Glu358 is part of a conserved ITE motif found in many retaining glycosidases. Similarly, in Cex, Glu 233 is part of a highly conserved ITE motif within the F-family of f3-1,4-glycanases (Figure 4.1). 4.2.1 Generation of mutants at position 233 Glu 233 was targeted by site-directed mutagenesis to determine whether it had properties consistent with those of a catalytic nucleophile. Site-directed in vitro mutagenesis was performed in order to substitute Glu 233 with Asp and Gin using the oligonucleotides as listed in Materials and Methods (Table 2.3). The phagemid pTZ18Rcex was constructed as outlined in Figure 4.2. Following mutagenesis, pTZ18R-cex was transformed to E. coliJMlOl. Initial screening for Cex mutants was on LB agar supplemented with 100 ig/mL ampicillin and 100 j.iM MUC. Approximately 50% of the colonies were non-fluorescing, indicating loss of exoglucanase activity. Mutation anywhere within the codon for Glu 233 resulted in the destruction of a Sca I restriction site. Restriction analysis of plasmid DNA isolated by small-scale alkaline lysis revealed approximately 80% of the non-fluorescing colonies had lost the ScaT restriction site. DNA sequencing of an 800 base pair Barn H1-Pstl fragment of several of these clones revealed they contained only the mutation of interest. The 800 bp sequenced mutant cassettes (the Barn H1-Pstl fragments) were subcloned from pTZI8R-cex into pUC121.lcex(PTIS), replacing the equivalent wild-type fragment of DNA as shown in Figure 4.3. pUC12-1.lcex(PTIS) was transformed to E. coli JM1O1 for expression of the mutant proteins. 56 Figure 4.2. Construction of pTZ18R-cex. Plasmids pTZI8R (Mead et al., 1986) and pUC12-1.lcex(PTIS) (ONei11 et a!., 1986a) were digested with BamHI and HindIII. The 1.8 kb BamHI-Hin dill fragment containing the cex gene was isolated and ligated with the 2.9 kb BamHI-HindIII fragment of pTZ18R to give plasmid pTZ18Rcex (4.7 kb). 57 tli.fidll1 ,Pstl BamHl BamHl Iaczpo IacZpo on Fl on cex pTZ18R U Ap Pstl V pUC12-1.lCex (PTIS) R 4.5kb R IacZ Hindill Pstl Sail Digest each with BamHI/HindIlI Isolate 1 .8 kb fragment by GeneCleanTM ligate DNA BamHl IacZpo oil Pstl cex pTZ18R-Cex 4.7kb lion IacZ Sail Pstl Hindlil R Ap 58 Figure 4.3. Generation of pUC12-I.Icex(PTIS) encoding Cex mutants. In vitro mutagenesis was performed as described Materials and Methods (section 2.4.2). For each mutant, an 800 bp mutant cassette in pTZI8R-cex was sequenced to confirm that only the desired mutation was present. The 800 bp fragment was isolated following digestion with Barn HI and Pstl. pUC12-I.lcex(PTIS) was also digested with Barn Hi and Psti and the 3.7 kb fragment was isolated. The 800 bp mutant cassette was ligated with the 3.7 kb fragment to replace the equivalent fragment in pUC12-i.lcex(PTIS). The positions of the mutation(s) generated in this study are indicated. in vitro mutagenesis 59 BamHl BamHl IacZpo E127 IacZpo on oil E233’ cex cex R pTZ18R-Cex 4.7 kb pUC12-tiCex Pstl Ap Ap 4.5 kb ii on IacZ IacZ Sail Pstl Hindlll Hindlil Sail Digest each construct with BamHI/Pstl Isolate 0.8 kb sequenced mutant cassette by GeneCleanTM Isolate 3.7 kb fragment by GeneCleanTM * ligate DNA BaniHi Dl24 El27 !acZpo on E233 Pstl cex pUC12-1.lCex (PTIS) 4.5kb R Ap IacZ Hindu Sail 60 The two mutants obtained, encoding Cex E233D and E233Q, were expressed at approximately the same levels as the wild type gene. The mutant proteins were purified by affinity chromatography on cellulose. When analysed by SDS-PAGE (Figure 4.4), the mutant polypeptides had been purified to electrophoretic homogeneity and had the same apparent molecular mass as wild-type Cex. No differences in behaviour from the wild-type were observed during the purification procedure. 4.2.2 Determination of kinetic parameters for Glu 233 mutants The kinetic parameters for hydrolysis of PNPC by mutants at position 233 are presented in Table 4.2. In comparison with the wild-type enzyme, E233D had only 1/4000 the activity. The Km value for this substrate rose only slightly upon mutation to Asp, from 0.53 mM to 1.2 mM suggesting that mutation of this residue likely effects both the glycosylation and deglycosylation steps of the reaction. If the mutation preferentially effected one step of the reaction, the concentration of the glycosyl-enzyme intermediate would change, which would likely be reflected in the Km value. No activity could be detected upon mutation of Glu to Gin. The kinetic parameters for the Glu 233 mutants in Cex were consistent with those found for mutants of the catalytic nucleophile in Abg (Withers and Street, 1988; Withers et al, 1990). Simultaneously, it was shown that Cex, like Abg, could be inactivated with 2F-DNPG leading to the accumulation of the covalent enzyme intermediate in which the sugar was esterified to Glu 233 (Tull et al., 1991). Labeling of this same residue was also been observed recently with 2’ ‘,4’ ‘-dinitrophenyl 2-deoxy-2-fluoro-13-cellobioside (2F-DNPC) (Tull and Withers, 1994) providing further evidence that Glu 233 is the catalytic nucleophile in Cex. 61 kDa 1 2 3 4 5 6 78 130 97.4 68.0 57.5— 53 -, 45— 41 / 36 / 29 E233D E233Q Figure 4.4. Purification of Cex E233D and E233Q. Cex P233D and E233Q were purified from 2-L cultures of E. coli containing pUC12-1.lcex E233D (PTIS) and pUC12-1.Icex E233Q (PTIS) as described in Materials and Methods (sections 2.7.2 and 2.7.3). The purification was followed by SDS-PAGE: Lane 1, molecular weight standards with sizes as indicated; lane 2, 3 tL (9.4 mg/mL) crude extract; lane 3, same volume extract following streptomycin sulfate precipitation; lane 4, same volume extract following 40,000 x g centrifugation; lanes 5 and 6: following cellulose affinity chromatography, E233D 2 ig and 10 .tg; lanes 7 and 8: E233Q 2 g and 10 pg respectively. 62 Table 4.2. Kinetic parameters for hydrolysis of p-nitrophenyl-I3-D cellobioside (PNPC) by Cex and Cex E233 mutants. 1 Enzyme 2 V max 1 .mg (.tmo1.min ) 1 Cex wild-type 14.4 Km (mM) kcat ) 4 (mm 677 3 o- E233D 3.6 x E233Q no activity detected 0.53 .168 1.2 — kcat/Km 4 .mM min ) 4 1278 1.97 — Parameters were determined at 37°C in 50 mM phosphate buffer, pH 7.0. 2 1 j.imol PNP released per minute per rng of purified enzyme. 63 4.3 Identification of the acid/base catalyst In addition to Glu 233, there are six other conserved acidic residues within the Ffamily of 3-glycanases which could be candidates for the acid/base catalyst in Cex (Figure 4.1). Of these, Glu 127 in Cex is part of a highly conserved region, WDVVNEA. Within this region, the short consensus sequence NEX (where X is a small hydrophobic residue) occurs in a large number of retaining glycanases in other sequence-related families. This type of motif has been suggested previously (Baird et al., 1990) to play an important role based upon alignments of sequences of a number of glycanases. It appeared therefore that Glu 127 might be a suitable candidate for the acid/base catalyst in Cex. 4.3.1 Generation of mutants at position 127 In vitro mutagenesis was performed in order to substitute Glu 127 with Asp, Ala and Gly using the oligonucleotides listed in Materials and Methods (Table 2.3). Approximately 50% of the colonies were non-fluorescent, indicating loss of exoglucanase activity. DNA sequencing of the 800 bp Barn H1-Pst I fragment of several of these clones revealed that about 80% of them contained only the mutation(s) of interest. Subcloning was performed as described for the Glu 233mutants. The three mutants obtained, E127A, E127G and E127D were expressed at approximately the same levels as the wild type. The purified proteins obtained by affinity chromatography on cellulose had the same apparent molecular mass as wild-type Cex and were electrophoretically homogeneous as judged by SDS-PAGE (Figure 4.5.) No differences in behaviour were observed during the purification procedure. Purified wild-type Cex and Cex E127A and E127G were subjected to mass spectrometry to verify that the actual molecular mass agreed with that predicted from 64 kDa 1 2 3 4 5 6 7 8 130 97.4 68.0 57.5 —I-. 41 36 29 E127A E127D E127G Figure 4.5. Purification of Cex E127A, E127G and P127D. Cex E127A was purified from a 20-L culture of E. coli containing pUC12-1.lcex E127A (PTIS). Cex E127D and E127G were purified from 2-L cultures of E. coil containing pUCI2-1.lcex E127D (PTIS) and pUCI2-1.lcex E127G (PTIS) respectively as described in Materials and Methods (sections 2.7.2 and 2.7.3). Purified protein was analyzed by SDS-PAGE: Lane 1, molecular weight standards with sizes as indicated; lane 2, 1 iL crude extract (0.lmLs of original culture volume); lanes 3 and 4, following cellulose affinity chromatography, 2 ig and 20 tg (by A280); lanes 5 and 6: E127D following cellulose affinity chromatography, 2 tg and 10 ‘g; lanes 7 and 8, E127G following cellulose affinity chromatography, 2 ig and 20 pg respectively. 65 the amino acid sequence. As shown in Table 4.3, the difference in mass between the mutants and the wild type were consistent with that predicted from the amino acid composition, within experimental error. When the mutants were reacted with 2,4DNPC prior to mass spectrometry, the masses increased by an amount equivalent to the mass of cellobiose (less an OH) indicating the presence of the cellobiosyl-enzyme intermediate as expected based on the mechanism proposed for the wild-type enzyme (see Introduction, Figure 1.4) 4.3.2 Kinetic characterization of mutants of Glu 127 The catalytic mechanism of retaining glycosidases such as Cex involves a two step process; formation and hydrolysis of a glycosyl-enzyme intermediate (see Introduction, Figure 1.4). It is likely that a single amino acid residue (the acid/base catalyst) is responsible for both proton transfer steps. The first step is acid catalysis, in which a proton is transferred from the acid/base catalyst to the glycosidic oxygen, facilitating bond cleavage through stabilization of the leaving group. In the second step, the same residue functions as a general base catalyst, removing a proton from water in a concerted process as the water attacks the anomeric center of the glycosyl-enzyme intermediate. It is clear, therefore, that mutation of the acid/base catalyst will affect the rates of both steps. The extent to which each step is affected however, is not necessarily equivalent. The effect of modifying the acid/base catalyst on the rate of the second step, deglycosylation, will necessarily be identical for all substrates containing the same sugar residues. However, the effects on the first step (glycosylation) will depend upon the leaving group ability of the aglycon (refer to Figtires 1.4 and 1.6). Those of high pKa, therefore poor leaving group ability, should be affected most following mutation of the acid/base catalyst, while those of low pKa, which need little or no protonic assistance for departure, should be affected very little. 316.1 ± 13.2 320.7 ± 11.4 47 370.2 ± 6.6 47 363.7 ± 6.0 47 064.43 47 050.40 47 054.8 ± 6.6 47 043.1 ± 5.4 E127G The Mr of 2,4 DNPC is 508.4. The Mr of cellobiose (less OH) is 325.3 2 Substrate was added to the enzyme sample prior to mass spec analysis. n.d. means not determined 1 Predicted Mr is based on the aa sequence and is calculated as the average isotopic (MH+) mass n.d. E127A n.d. 47 122.47 47 119.5 ± 3.8 Difference (b-a) Wild-type Cex Reacted with 2 (b) ’ 1 2,4-DNPC Predicted Mr 1 Determined Mr (a) Mass spectrometry of purified Cex and E127 mutants in the absence and presence of 2,4-DNPC Protein Table 4.3. 67 For each mutant at position 127, kinetic parameters were determined for various cellobiosides and glucosides with differing requirements for acid catalysis (Table 4.4). Three different cellobioside substrates were studied: PNPC (pKa of phenolic leaving group = 7.18), 2,4-DNPC (pKa of phenolic leaving group = 3.96) and 4-BrPC (pKa of phenolic leaving group = 9.34). Kinetics were also determined for 2,4-DNPG (pKa of phenolic leaving group = 3.96). The effect of mutation of Glu 127 to Ala, Glu or Asp on kcat values for PNPC, 2,4-DNPC and 2,4-DNPG was to decrease them approximately 200 to 400-fold. There were also dramatic decreases in Km for these substrates, ranging from 20-fold for PNPC to 600-fold for DNPG. With the mutants, the Km values for 2,4DNPC were less than 1 p.M, a drop of about 200-fold relative to the wild-type enzyme. The kcat/Km values however, for PNPC and 2,4-DNPC remained relatively unchanged relative to the wild-type. Based on the data in Table 4.4 , it seems likely that deglycosylation is the ratedetermining step for hydrolysis of PNPC and 2,4-DNPC for the Glu 127 mutants, as it is for the wild-type enzyme. First, essentially the same kcat values were seen for the two substrates with quite different leaving-group ability. Second, the Km values dropped dramatically. The drop in Km suggests an accumulation of the glycosyl-enzyme intermediate as would be expected if the deglycosylation step were rate-limiting. Similarly, the rate-determining step for DNPG appears to be deglycosylation. For this substrate, the rate-determining step seems to have switched from glycosylation in the case of the wild-type enzyme, to deglycosylation with the mutant. It is apparent, therefore, that since deglycosylation is rate-limiting in the presence of a good leaving group (i.e. pNP or DNP), the decrease in kcat observed following substution of Glu 127 is due to slowing of the deglycosylation (base catalysis) step. 68 Table 4.4. Kinetic parameters for hydrolysis of various substrates by Cex and E127 mutants’ Enzyme Substrate ) 4 kcat (mm Km (mM) kcat/Km mM (min ) 1 Native Cex DNPG2 2 DNPC PNPC2 4-BrPC 860 419 677 255 1.87 0.06 0.53 2.0 460 6983 1278 128 E127A DNPG DNPC PNPC 4-BrPC 2.7 2.4 2.3 0.04 0.0032 0.0003 0.025 1.9 843 7742 92 0.02 E127G DNPG DNPC PNPC 2.2 2.1 2.0 0.0035 0.0004 0.029 628 5526 69 E127D DNPC PNPC 1.6 3.3 0.0017 1.74 941 2 1 P arameters were determined at 37°C in 50 mM phosphate buffer, pH 7.0. Data from Table 4.1. 2 69 Further evidence of the slowing of the deglycosylation step and the resulting accumulation of the glycosyl-enzyme intermediate can be seen from examination of the mass spectrometry data (Table 4.3). As mentioned previously, reaction of the 127A or 127G mutants with 2,4-DNPC prior to mass spectrometry resulted in an increase in molecular mass corresponding to an accumulation of the glycosyl-enzyme intermediate. Such an increase in mass can not be observed with the wild-type enzyme presumably because the enzyme-glycosyl intermediate is rapidly hydrolysed (S. Withers, personal communication). Again, these results suggest that with substrates with a good leaving group (which presumably do not require acid-catalytic assistance for hydrolysis), the reductions in activity observed are likely due to a slowing of the deglycosylation step, where base catalysis occurs, and not due to a slowing of the rate of glycosylation. One measure of the effect of this mutation on the glycosylation step or acid catalysis step for these substrates (i.e. the formation of the glycosyl-enzyme intermediate) can be obtained by examination of kcat/Km values, as mentioned in the introduction. Essentially no reduction in kcat/Km was observed with 2,4-DNPC, while a 10-20 fold reduction was observed for PNPC. The effect of the mutation upon the glycosylation step can also be estimated by studying the kinetics of inactivation with 2FDNPC. As discussed earlier, the inactivation of Cex with 2F-DNPC involves simply the formation of the glycosyl-enzyme intermediate. The results of inactivation of E127A with 2F-DNPC are shown in Figure 4.6. It is apparent that this mutant was inactivated very quickly. The rate of inactivation of E127A by 2F-DNPC (k/K was at least as fast as that of the wild type enzyme (k/Kt = = 1.6 1 min m M’) ) 1 0.61 minmM determined previously (Tull and Withers, 1994). Again, these results suggest that mutation of Glu 127 to Ala did not slow down the rate of the glycosylation (acid catalysis) step in the reaction when a substrate with a good leaving group was hydrolysed. 70 Figure 4.6. Inactivation of Cex E127A by 2F.-DNPC. Panel A: Residual DNPCase activity vs. time for E127A incubated with 2F-DNPC at the following concentrations: 0.031 mM (V), 0.045 mM (), 0.090 mM (EJ) and 0.136 mM (0). 0.015 mM Residual enzyme activity was monitored as described in section 2.9.2. Panel B: double reciprocal plot of k b (derived from panel A) vs. 2F-DNPC concentration. 0 Panel C: The inactivation rate constant (k) and the inactivation binding constant ) were derived from the data in panel B. 1 (K (a), 71 A 0.0 1.0 0 > -2.0 > -3.0 -4.0 10 0 30 20 40 Time (mm) B 60 50 C E 40 Co 30 -o 0 20 10 0 0 10 20 30 40 1/[2F-DNPC] 50 60 70 (1/mM) C Inactivation rate and binding constants for E127A with 2F-DNPC 1 (min k ) 1 1 (mM) K kj/Kj (minl.mm1) 0.44 ± 0.18 0.28 ± 0.16 1.6 ± 0.34 72 E127A was also assayed with 4-BrPC (pKa of phenolic leaving group = 9.34), a substrate with a greater requirement for acid-catalytic assistance. As shown in Table 4.4, the kcat value was lowered 6400-fold in comparison with the wild-type enzyme, whereas the Km value remained relatively unchanged. The kcat/Km value was also reduced 6400-fold relative to the wild-type enzyme with this substrate. The ratedetermining step for this substrate with wild type enzyme is thought to be formation of the glycosyl-enzyme. The fact that the Km value for the mutant was similar to that of the wild type would suggest that the glycosylation step remained rate-limiting with the mutant. The 6400-fold rate reduction therefore represents the effect upon the glycosylation step (acid catalysis) for this substrate, a much greater rate-reduction than that seen on the glycosylation step for 2,4-DNPC or PNPC. Hydrolysis of 4-BrPC would be expected to be more dependent on the acid catalyst than hydrolysis of PNPC, 2,4DNPC and 2,4-DNPG. The marked reduction in kcat for hydrolysis of 4-BrPC and the much smaller reductions for the hydrolysis of PNPC, 2,4-DNPC and 2,4-DNPG by E127A are consistent with Glu 127 being the acid/base catalyst in Cex. 4.3.3 Effects of sodium azide on reaction rates With mutants of Glu 127, the presence of a good leaving group on substrates such as 2,4-DNPC or PNPC reduced the need for acid catalytic assistance in the formation of the glycosyl-enzyme intermediate. In effect, the presence of a good leaving group on substrates such as 2,4-DNPC or PNPC compensated for the missing acid catalyst function of the mutants. The drop in activity seen with these substrates reflects primarily the effect of the mutation on the deglycosylation step, which would normally be facilitated by base catalysis. An attempt was made to compensate for the missing base catalyst function. With the wild-type enzyme, hydrolysis of the glycosyl-enzyme intermediate involves the transfer of a proton from water to the acid/base catalyst. The 73 resultant hydroxyl (-OH) species then acts as a nucleophile to displace the catalytic nucleophile Glu 233, resulting in a product showing retention of configuration at the anomeric carbon. The mutation of Glu 127 to Ala or Gly presumably generates a cavity in the active site. Anions of high nucleophilicity might bind at this site and react in place of water without any need for general base catalysis. Various competitive nucleophiles including acetate, azide, cyanide, formate, imidazole, and thiocyanate were tested for their ability to enhance the rate of hydrolysis of PNPC by the Glu 127 mutants. Of the nucleophiles tested, only sodium azide was found to effect substantially the rate of hydrolysis of PNPC. The rate-enhancement was maximal with about 100 mM azide. As indicated in Figure 4.7, the effect was specific to the alanine and glycine mutants. No rate-enhancement was observed with either E127D or with the wild-type enzyme. This was probably because the glutamate or aspartate carboxylate prohibited the access of the azide nucleophile, either by steric hindrance, or more likely, through electrostatic repulsion. The kinetics of the rate enhancement were studied in detail with the E127A mutant and several different substrates. The kinetic parameters for hydrolysis of PNPC, 2,4-DNPC and 2,4-DNPG were analyzed at various concentrations of sodium azide. Maximal increases in kcat were about 8-fold for PNPC to 10-fold for DNPG in the presence of 60 mM azide as shown in Figures 4.8 and 4.9, respectively. A dramatic increase in kcat of over 200-fold was seen with 2,4-DNPC at up to 2 M azide (Figure 4.10), surpassing the rate of the wild-type enzyme. The most likely explanation for this is that as the azide concentration increased, the deglycosylation rate continued to increase until it surpassed the glycosylation rate. The limiting rate observed in each case presumably reflected the rate of the glycosylation step for each substrate. With 2,4DNPC, glycosylation is facilitated by the good leaving-group ability of 2,4-DNPC, as discussed earlier. In contrast, azide did not increase the keat for PNPG (not shown). Q 74 E127A E127G 0.25 0.4’ 0.20E E 0.15 0) a D .S 0.3 0.2 0.10 > > U C° 0.05 0.00- 0.1- 0 0 i.i.i.i- 50 0.0 I 150 250 350 450 Azide (mM) i-i.i.i. 50 E127D Wild—type 0.20- I I 150 250 350 450 Azide (mM) 20 0.15- 15 E 0, 0 0.10 10 >‘ >‘ II > > 005 n nn. 0 1111’ 50 150 250 Azide (mM) 350 450 5, 0’ 50 150 250 350 450 Azide (mM) Figure 4.7. PNPCase activity of Cex and E127 mutants in the presence of various concentrations of sodium azide. Reactions were carried out at 37°C in 50 mM phosphate buffer, pH 7.0 with 6 mM (>10 x Km) PNPC for about 15 mm. The reaction was stopped by the addition of 3 CO and the absorbance at 400nm was 2 Na read. Activity is expressed in units 1mg enzyme where 1 Unit =1 tmo1 PNP released/mm. 75 20 0 0.30 600 15 0.20 C E 400 0 C 10 .4-’ (U 0 0.10 2 E E 200 5 0 0 10 20 30 40 50 60 0.00 0 Azide concentration (mM) Figure 4.8. Kinetic parameters for hydrolysis of PNPC by Cex El. 27A in the presence of various concentrations of sodium azide. Parameters were determined at 37°C in 50 mM phosphate buffer, pH 7.0. 2 76 30 30 20 o 15 20 C - -— 2 o 10 5 0 A 0 - 0 10 - 20 . 30 I 40 . I 50 . • 60 0 0 Azide concentration (mM) Figure 4.9. Kinetic parameters for hydrolysis of 2,4-DNPG by Cex E127A in the presence of various concentrations of sodium azide. Parameters were determined at 37°C in 50 mM phosphate buffer, pH 7.0. 77 600 0.06 400 0.04 600 0 ‘I LI 400 E (U 0 200 0.02 E E E 200 E (U 0 0 0.00 0 500 1000 1500 0 2000 Azide concentration (mM) Figure 4.10. Kinetics of hydrolysis of 2,4-DNPC by Cex E127A in the presence of various concentrations of sodium azide. Parameters were determined at 37°C in 50 mM phosphate buffer, pH 7.0. The dotted line represents the activity on this substrate (kcat) for the wild-type enzyme. 78 This finding is entirely reasonable since the rate-determining step for this substrate was shown previously to be glycosylation; thus, increasing the rate of the deglycosylation step should not affect the steady state rate. Dramatic increases in Km values with increasing azide concentration were observed for each substrate, essentially paralleling the effects on kcat (Figures 4.8, 4.9 and 4.10). The net result of this was that values of kcat/Km remained essentially constant at differing azide concentrations. The increases in Km were likely a consequence of the decreased extent of accumulation of the glycosyl-enzyme intermediate as the rate of the deglycosylation step increased. This finding is consistent with azide acting primarily on the deglycosylation step, that is, on the base catalysis step and not on glycosylation. 4.3.4 Effects of sodium azide on products of hydrolysis If Glu 127 is the acid/base catalyst, mutation to Ala or Gly should generate a cavity close to the f3-face of the substrate. If azide activates the mutants by the mechanism proposed above (see Figure 4.11), the reaction product should be the corresponding 3-glycosy1-azide. In order to test this hypothesis, the products of hydrolysis in the absence and presence of azide were analysed by thin layer chromatography (TLC). Analysis of reaction mixtures containing PNPC revealed that wild-type Cex produced only the expected nitrophenol (visible under UV light) plus cellobiose, both in the absence and presence of sodium azide as shown in Figure 4.12. With both E127A and E127G the reaction products differed in the absence and presence of azide (Figure 4.13). In the absence of azide, only cellobiose plus nitrophenol was observed in each case. About 50% of the starting material remained due to the reduced rate of hydrolysis by the 79 I E127A E127A OH OH HO’° HO0 OR HO Qo HO 33 E233 - Glycoson ROH E127A OH HO0 HO gIycosy1-enzyn I Q0 intemdiate 33 De1Ycoo,,/) + N ...E127A I E127A I I OH HO-° I I I I I OH HO-L..->’ HOào 33 I 33 I Figure 4.11. Proposed nchanism for hydrolysis of glycosides by Cex E127A in the presere of azide. Only 1 glucose residue is shown for simplicity. 80 Wild-type Cex pNPC cellobiosyl-azide cellobiose s- + Figure 4.12. Hydrolysis of PNPC by wild type Cex in the presence and absence of azide. Reactions ran for 2 h at 37°C in 50 mM phosphate buffer, pH 7.0 with 1 p.g/mL Cex, 6 mM PNPC and 60 mM azide where indicated. Thin layer chromatography was performed as described in section 2.10.1. Standards lane (5): 2 i.LL each of 6 mM PNPC, cellobiose and cellobiosyl-azide. Reaction lanes (-) without azide, (+) with azide, 2 jiL each. 81 E127A El 27G pNPC cellobiosyl-a zide cellobiose S - + S- + Figure 4.13. Hydrolysis of PNPC by Cex E127A and E127G in the presence and absence of azide. Reactions ran overnight at 37°C in 50 mM phosphate buffer, pH 7.0 with 100 .tg/mL mutant Cex, 6 mM PNPC and 60 mlvi azide where indicated. Thin layer chromatography was performed as described in section 2.10.1. Standards lane (S): 2 jiL each of 6 mM PNPC, cellobiose and cellobiosyl-azide. Reaction lanes: (-) without azide, (+) with azide, 2 tL each. 82 mutants. In the presence of azide however, a different sugar product was formed. The new compound co-migrated on TLC with 3-cellobiosyl-azide. The kinetics of hydrolysis with the alanine mutant indicated that cellobiosyl-azide was in fact the only product produced, as shown in Figure 4.14. With reduced concentrations of azide, both cellobiosyl-azide and cellobiose could be detected as products (not shown). The stereochemistry of the new product was determined by H 1 -NMR spectrometry. The analysis revealed an identical spectrum to that obtained from the chemically synthesized f3-cellobiosyl-azide (Table 4.5), providing confirmation of the f3- stereochemistry. With 2,4-DNPG, the new product formed was 3-glucosyl-azide as judged by TLC (not shown) and 1 H-NMR (Table 4.5). Table 4.5. H 1 -NMR spectraa for -cellobiosyl-azide and f3-glucosyl-azide: -cellobiosyl-azide, 1 H-NMR (400 MHz, D 0): 8 4.47 (d, Ji, 2 2 Hz, H-i); 3.89 (m, 3 H); 3.65 (m, 4 H); 3.35 (m, 6 H). = 7.8 3-glucosyl-azide, 1 H-NMR (400 MHz, D 0): 8 —4.7 (d, Ji, 2 2 = 9.0 Hz, H-i); 3.89 (dd, 1 H, J 12.4, 6 12.4, J 65 2.2 Hz, H-6); 3.71 (dd, 1 H, 5.6 H-6’); Hz, (m, 3.50 2 H, H-3, 3.40 H, 5); 1 9.2, 9.8 (t, ’, 6 J 5 413 J 45 J Hz, H-4); 3.24 (t, 1 H, J 21 9.0, J 3 8.9 Hz, H-2). , 2 aprovided by T. Lindhorst. Identical results were observed with both the alanine and glycine mutants. The results are summarized in Table 4.6. The fact that only f3-cellobiosyl-azide or 3glucosyl-azide was detected confirmed azide attacked only from the top face of the substrate. This is consistent with a location of the acid/base catalyst on this same face and confirms that azide activates the mutants by the mechanism proposed. 0 0.75 1.5 2.25 3.0 3.75 4.5 5.25 Std - —cellobiose —cello bios y I a z ide —pNPC Figure 4.14. Time course of hydrolysis of PNPC by Cex E127A in the presence of azide. Reactions were carried out at 37°C in 50 mM phosphate buffer, pH 7.0 with 200 .tg.mL 1 mutant Cex, 6 mM PNPC and 60 mlvi azide. Thin layer chromatography was performed as described in section 2.10.1. Standards lane (Std): 2 jiL each of 6 mM PNPC, cellobiose and cellobiosyl-azide. Reaction lanes: 2 iL samples were withdrawn at various intervals as indicated. Time (h): Cex El 27A 00 84 Table 4.6. Products of hydrolysis with wild-type and E127 mutants in the presence or absence of sodium azide Enzyme Substrate Products 1 (no azide) (with azide) Wild-type PNPC cellobiose cellobiose E127D PNPC cellobiose cellobiose E127A PNPC DNPG PNPG cellobiose glucose glucose j3-cellobiosyl-azide 13-glucosyl-azide glucosyl-azide E127G PNPC DNPG cellobiose glucose f3-cellobiosyl-azide 3-glucosy1-azide Products were determined by thin layer chromatography. Reaction mixtures were incubated overnight at 37°C in 50 mM phosphate buffer, pH 7.0 with 100 j.ig/mL mutant Cex or 1 Lg/mL Cex, 6 mM substrate and 60 mM azide where indicated. Stereochemistry was determined by 1 H-NMR. Products also include p-nitrophenol or dinitrophenol. 85 4.4 Asp 123: another conserved acidic residue As a control, similar experiments involving sodium azide were performed with mutants of another residue conserved within the F-family, Asp 123. Asp 123 is also part of the WDVVNEA motif and is in proximity to Glu 127. Site-directed in vitro mutagenesis was carried out in order to substitue Asp 123 with Ala as described for mutants at position 233 and 127. The protein was purified by affinity chromatography on cellulose as described for the previous mutants. The purified protein is shown in Figure 4.15. The rate of hydrolysis of PNPC by D123A was found to be affected both by azide and by thiocyanate as shown in Figure 4.16. The rate-enhancement observed with thiocyanate was about half that observed with azide. For either anion, the rate-enhancement was maximal at about 500 mM. The kinetic parameters for hydrolysis of PNPC and 2,4-DNPC by D123A are presented in Table 4.7. The kcat value for 2,4-DNPC was similar to that of the wild-type enzyme, whereas that of PNPC was reduced by about 1500-fold. Initially, the small rate-reduction for 2,4DNPC and the much larger rate reduction for PNPC appeared to be consistent with a role of Asp 123 as acid/base catalyst. The values for both substrates, however, increased 10-fold and 33-fold for 2,4-DNPC and PNPC respectively. The increase in Km is inconsistent with mutation at Asp 123 slowing down preferentially the deglycosylation step of the reaction. This is in contrast to what was observed with mutants at position 127, where Km values for these substrates were drastically reduced. The kinetic parameters for hydrolysis of PNPC were determined also in the presence and absence of azide. In the presence of 500 mM azide, the rate of hydrolysis of PNPC rose about 10-fold, whereas the Km value dropped slightly. The drop in Km is inconsistent with azide acting to increase the rate of the 86 1 kDa 2 3 200 116 97.4— 66 — — 45 31 21.5— D123A Figure 4.15. Purification of Cex D123A. Cex D123A was purified from a 2-L culture of E. coli pUC12-1.lcexDl23A (PTIS) as described in Materials and Methods (sections 2.7.2 and 2.7.3). The purification was analysed by SDS-PAGE: Lane 1, molecular weight standards with sizes as indicated; lanes 2 and 3, following cellulose affinity chromatography, 2 ig and 10 jig respectively. 87 8.0 o 7.0 6.0 x C 0 0 0 0 D 5.0 E 0 0 0 azide thiocyanate D 4.0 D D 3.0 > 0 2.0 0 1.0 I 0.0 0 500 1000 — I 1500 — — I 2000 Anion concentration (mM) Figure 4.16. PNPCase activity of Cex D123A in the presence of azide or thiocyanate. Reactions were carried out with 30 .ig/mL enzyme at 37°C in 50 mM phosphate buffer, pH 7.0 with 5 mM PNPC and anion concentrations, NaN 3 (0) and KSCN ( D) as indicated. Hydrolysis of PNPC was monitored continuously at 400 nm over a 10 mm. period. Activity is expressed as initial rates of hydrolysis (Vo). 88 Table 4.7. Kinetic parameters for hydrolysis of cellobiosides by D123A 1 Enzyme Substrate ) 1 kcat (miir Km (mM) kcat/Km 1 mM (min ) 1 Wild-type DNPC PNPC 419 677 0.06 0.53 6983 1278 D123A DNPC PNPC PNPC (O.5Mazide) 325 0.43 0.53 18 613 0.02 5.0 11 0.47 Parameters were determined at 37°C in 50 mM phosphate buffer, pH 7.0. 1 89 deglycosylation step. Azide did not enhance the rate of hydrolysis of 2,4DNPC (not shown). TLC analysis of the products of hydrolysis of PNPC in the presence and absence of azide revealed that only cellobiose was produced (Figure 4.17). Furthermore, cellobiosyl-azide was found not to be a substrate. The lack of cellobiosyl-azide as a product clearly distinguishes the behaviour of Asp 123 from that of Glu 127. Possible roles for Asp 123 and the mechanism of azide and thiocyanate activation will be discussed. 90 D123A pNPC cellobiosyl-azide cellobiose S - + Figure 4.17. Hydrolysis of PNPC by Cex D123A in the presence and absence of azide. Reactions mixtures were incubated overnight at 37°C in 50 mM phosphate buffer, pH 7.0 with 500 ig/mL Cex D123A, 6 mM PNPC and 500 mM azide where indicated. Thin layer chromatography was performed as described in section 2.10.1. Standards lane (S): 2 p.L each of 6 mM PNPC, cellobiose and cellobiosyl-azide. with azide, 2 jiL each. Reaction lanes (-) without azide, (+) with azide. 91 5. Discussion 5.1 Expression of the cex gene in Streptomyces lividans The cex gene was subcloned and successfully expressed in S. lividans TK64 from both the S. antibioticus melanin (mel) promoter of p1J702 and from the S. fradiae aminoglycoside phosphotransferase (aph) promoter of p1J680. Cex was produced as a soluble polypeptide, and retained full catalytic activity. Intracellular activity could not be detected and Cex was efficiently secreted into the culture supernatant. Processing of the leader peptide of Cex by Streptornyces resulted in equal proportions of the polypeptide having either the natural amino-terminus or an amino-terminus with an additional 2 amino acids. Studies of 40 secreted Streptomyces polypeptides have shown that the -1 and +1 residues, relative to the processing site, are usually small neutral amino acids, specifically alanines and serines (Brawner et al., 1991). Furthermore, alanine is very often found at the -3 position. The normal (C. firni) processing site (AQA t ATTLK) appeared to be well suited, therefore, for correct processing in Streptomyces. However, amino-terminal heterogeneity resulting from signal peptide processing is not uncommon in foreign proteins expressed in Streptomyces (Brawner et al., 1991). There was no indication, however, that the additional two amino-terminal residues in Cex affected the activity of the enzyme. Maximum exoglucanase activity in the culture supernatant of S. lividans TK64 [p1J680-cexl was reached following 40 h of growth, representing about 5.5 mg of Cex per liter, based on the specific activity of the purified protein. The fragment of C. fimi DNA in p1J680-cex and p1J702-cex contained the cex gene and its promoter. When the fragment was inserted in the reverse orientation in p11702, the production of Cex was reduced by about 90%. One interpretation of this result is that the cex promoter was not 92 efficiently recognized by S. lividans. It is of note that although S. lividans will recognize many eubacterial promoters (Gusek and Kinsella, 1992; M. Bibb, personal communication) Streptomyces promoters usually lack the typical -35 and -10 regions found upstream of the cex gene and other eubacterial genes. Removal of the upstream (740 bp) non-coding region of the cex gene and its promoter, and fusion of the sequence encoding the leader peptide of Cex and the Cex cellulose binding domain to the aph structural gene in p1J680 increased expression slightly (Ong et al., 1994). The increase in expression might be a reflection of increased translational efficiency. In this case, translation was initiated from the aph ribosome-binding site, and not the cex ribosome binding site. The ribosome-binding site in cex (AGGAGG) is consistent with that of the consensus ribosome-binding site deduced from 40 Streptomyces genes ((A/G)GGAGG) (Anne and Van Mellaert, 1993) although it may less suitably located (4 nucleotides in cex vs. 5 to 12 in Streptomyces genes relative to the initiation codon) for efficient initiation of translation. The level of expression from the aph promoter in p1J680-cex was about 6-fold higher than that from the mel promoter of p1J702-cex. Other investigators have observed the aph promoter to be superior to the mel promoter when used to express the same fragment of foreign DNA. (L. Carlson, personal communication). The S. fradiae aph promoter was the first promoter employed to direct the expression of a heterologous gene in Streptomyces on the basis that the promoter reportedly directs the production of its gene product to 10% of the total soluble cell protein (Brawner et at., 1991). It is clear, however, that more study must be done in order to understand how the signals for transcription initiation, translation and protein localization in Streptomyces effect the efficiency of heterologous gene expression. Streptomyces is clearly a potentially very useful host for the high level expression of heterologous genes. The level of expression 93 of the cex gene in Streptomyces , however, was not sufficiently high to offer an advantage over E. coli as a host for the expression of this gene. Native Cex from C. fimi is a glycoprotein. Interestingly, S. lividans also produced a glycosylated Cex. Based on the apparent molecular mass of the recombinant protein, the extent of glycosylation was very similar to that of Cex from C. firni. In addition, the glycosyl groups on the Streptomyces-produced Cex were functionally similar to those of C. fimi. The glycosylation afforded protection against proteolysis by C. fimi protease particularly when the enzyme was bound to cellulose, and to a lesser extent when in solution. This might be the result of greater steric hindrance of the protease when the enzyme was immobilized on the cellulose surface. Previous work has shown that in the case of non-glycosylated Cex from E. coli, proteolysis by either the C. fimi protease or papain, occurs within and at the ends the PT linker region effectively separating the catalytic domain and the cellulose-binding domain (Gilkes et al., 1991a). The linker region may be susceptible to the C. fimi protease or papain because it presumably adopts an extended conformation. It has subsequently been shown that in S. lividans produced Cex, the PT linker region is in fact the site of glycosylation (Ong et a!., 1994). S. lividans is known to produce native glycoproteins, for example, a glucosidase (Mihoc and Kluepfel, 1990) and a xylanase (Kluepfel et a!., 1990). The glycosylation of Cex observed in this study is significant in that it was the first report of a heterologous glycoprotein produced by this organism (MacLeod et al., 1992). Recently, a cellobiohydrolase from Microbispora bispora has also been shown to be glycosylated by Streptomyces (Hu et al., 1993). Hosts which secrete and glycosylate heterologous polypeptides expressed from cloned genes are potentially very useful both for basic science and industrial applications. Although E. coli, is a commonly used host for the expression of heterologous genes, it does not glycosylate polypeptides. C. fimi is a potentially useful host for the expression of genes encoding glycoproteins, but suitable 94 vectors have yet to be developed. Sacchnrornyces cerevisiae, another commonly used host, does produce glycoproteins. However, S. cerevisine may hyperglycosylate them, giving them undesirable characteristics (Curry et al., 1988). The use of mammalian cells for the expression of cloned genes often yields polypeptides with the correct, or at least acceptable, patterns and levels of glycosylation (Goochee et a!., 1991). However, animal cell culture processes are complex and expensive. It is clear, therefore, that the ability of Streptomyces to glycosylate heterologous polypeptides is a very useful feature of this bacterium. 5.2 Catalysis 5.2.1 The catalytic nucleophile Glutamic acid 233 in Cex, the putative catalytic nucleophile, was targeted by sitedirected mutagenesis on the basis of amino acid sequence alignments. Glu 233 in Cex is part of a conserved ITE motif within family F cellulases and xylanases. Glu 358 in Abg from A robacterium, also part of a conserved ITE motif within its family, was identified as the catalytic nucleophile in this retaining enzyme by inactivation with the mechanism-based inactivator 2F-DNPG (Withers and Street, 1988; Withers et a!, 1990). The kinetic parameters found in this study for mutants of Glu 233 in Cex were consistent with those found for similar mutants of the catalytic nucleophile in Abg (Withers et al., 1992). It is clear from these studies that the shortening of the carboxylate side chain of the catalytic nucleophile by as little as lÀ (a Glu --> Asp mutation) has a dramatic impact on the activity of the enzyme. Shortly after initiation of this study, the labeling of Glu 233 with the inactivator 2F-DNPG (Tull et a!., 1991), and more recently with 2F-DNPC (Tull and Withers, 1994) provided unequivocal evidence that Glu 233 was the catalytic nucleophile in Cex. 95 5.2.2 The acid/base catalyst Glu 127 was proposed as the acid/base catalyst in Cex on the basis of amino acid sequence alignments. The kinetic parameters determined for mutants at position 127 provide firm support for the assignment of Glu 127 as the acid/base catalyst in Cex. The kcat/Km values for substrates with various leaving group abilities provide an estimation of the effect of the mutation on the glycosylation step (where acid catalysis occurs). With DNPC, which presumably does not require acid-catalytic assistance for initial bond cleavage, the Glu 127 mutants exhibited kcat/Km values similar to that of the wild-type enzyme. With PNPC, the leaving group of which has an intermediate pKa, the mutants exhibited somewhat reduced kcat/Km values relative to the wild-type enzyme. When 4-BrPC was tested with the E127A mutant, a substrate which presumably requires acid-catalytic assistance for initial bond cleavage, the kcat/Km value was dramatically reduced relative to the wild-type enzyme. Therefore, the rate constants for glycosylation were affected relatively little for substrates not requiring acid catalysis, and dramatically for substrates requiring acid catalysis. An accumulation of the glycosyl-enzyme intermediate would be expected if the rate of the deglycosylation step was reduced to a greater extent than that of the glycosylation step, and should also result in decreased Km values. Relative to the wild-type enzyme, the decreases in Km for the substrates (about 200-fold for DNPC, 20-fold for PNPC and unchanged with 4-BrPC) are consistent with this interpretation. In other words, with a good leaving group (i.e. DNP) the rate of the glycosylation step remained unchanged relative to the wild-type. Consistent with this was the finding that E127A was in fact inactivated with 2F-DNPC at a rate at least as fast as that of the wild-type enzyme. The reductions in kcat values observed for DNPC and PNPC with the Glu 127 mutants (200400 fold) therefore reflect primarily the effect upon the deglycosylation step of the reaction, due to the removal of general base catalytic assistance. 96 Sodium azide significantly enhanced the rate of hydrolysis of PNPC, DNPC and DNPG by uncharged Glu 127 mutants. With increasing azide concentration, both kcat and Km values rose concurrently, the net result of which was that kcat/Km values remained essentially constant across the range. Therefore, the kinetic parameters for the mutants in the presence of sodium azide were consistent with azide acting to increase the rate of the deglycosylation (base-catalysis) step. The proposal that azide occupied a vacant anionic site created by removal of the acid/base catalyst, and acted instead of water (OH-) as a nucleophile in the hydrolysis of the glycosyl-enzyme intermediate was confirmed by the detection of the 13-linked cellobiosyl-azide (13-glucosyl-azide with DNPG as substrate). The fact that only f3-linked adducts were detected confirms that azide attacked only from the 13 face of the substrate, a result consistent with a location of the acid/base catalyst (Glu 127) on this same face. Interestingly, the rate-enhancement with azide was specific to the Ala and Gly mutants. A rate enhancement was not seen with the wild-type enzyme or with the E127D mutant. Presumably the glutamate or aspartate carboxylates prohibited the access of azide through electrostatic repulsion. The lack of a rate enhancement with E127D was unlikely due to a steric effect in light of the finding of Cupples et al. (1990) that azide could in fact enhance the rate of an E461Q mutant in 13-galactosidase, an uncharged, but equally bulky substitution. It is likely, therefore, that an E127Q or E127N mutation in Cex would have a similar effect. Although several nucleophiles other than azide were tested for their effects on the E127A and E127G mutants in the current study, only azide was found to significantly increase rates of hydrolysis and form adducts detectable by TLC. It’s likely, however, that other nucleophiles should be able to function in this manner. Huber and Chivers (1993) recently reported that similar 13-galactosyl adducts were produced by uncharged E461 mutants of f3-galactosidase in the presence of nucleophiles 97 including azide, acetate and imidazole to name a few. It was further suggested that with certain nucleophiles, f3-galactosyl adducts such as 13-galactosyl-acetate were formed but are unstable. The techniques employed in this study may be applicable to other glycosidases belonging to sequence-related families (Henrissat, 1991). The approach would involve generation of alanine or glycine mutants of conserved glutamic and aspartic acids. The mutants would then be screened for the generation of new products of hydrolysis in the presence of azide and other anionic nucleophiles. Positive mutants should then be subjected to the following kinetic analysis. kcat and Km values should be measured for a pair of substrates, one of which requires acid catalysis and one of which does not. Values of kcat/Km for these two substrates should differ greatly. For the activated substrate, the kcatlKm value should be much higher than that for the other. Values of kcat and Km for highly activated substrates with these mutants should increase dramatically with increasing azide concentration, and level off at a value dependent on the leaving group ability of the aglycon. This sort of approach could facilitate identification of the acid/base catalyst in other glycosidases. The acid/base catalyst in a family G xylanase from Bacillus circulans (Glu 172) was deduced from examination of the crystal structure (Wakarchuk et al., 1994). Uncharged mutants at this position tested for similar azide effects in light of this investigation resulted in the generation of 13xylobiosyl-azide from p-nitrophenyl 13-xylobioside (PNPXX). In addition, preliminary studies with acid/base catalyst mutants of Agrobacterium f3-glucosidase have yielded similar results (S. Withers, personal communication). The generation of the new f3-linked product in this study is clearly a significant result. A rate-enhancement in the presence of the nucleophile in which a conserved acidic residue has been mutated does not constitute a positive screen for an acid/base catalyst. First, in the case of a retaining enzyme, the substrate used must be rate- 98 limiting for the deglycosylation step with the mutant in question. For example, enhancement of the rate of hydrolysis with PNPG was not observed with the mutants in the presence of azide, although glucosyl-azide could be detected by TLC. This result is not surprising given that the rate-limiting step for this substrate with the wild-type enzyme is glycosylation (Tull and Withers, 1994). Activation of the deglycosylation step by azide, therefore, would not be expected to affect the steady-state rate of hydrolysis of PNPG. Second, anions clearly will enhance the steady-state rates of mutants at other positions. In this study, both azide and thiocyanate enhanced the rate of mutants of Asp 123, a conserved acidic residue in close proximity to Glu 127. In this case however, cellobiose was the only product in the absence or presence of azide. No concomitant production of cellobiosyl-azide or cellobiosyl-thiocyanate was observed. In this case azide or thiocyanate might bind at the site of the missing side chain restoring local charge requirements. Asp 123, however, clearly plays a role in facilitating catalysis. Conceivably, this residue might be involved in binding the substrate, or play a role in maintaining the correct environment (i.e. pKa’s) at the acid/base catalyst or the catalytic nucleophile possibly through a hydrogen bonding network. It is of note that in B. circulans xylanase, there is reportedly a conserved acidic residue found within the active site, in addition to the acid/base catalyst and the catalytic nucleophile (Wakarchuk et al., 1994). An inverting glycosidase, such as CenA from C. fimi, also hydrolyses f3-1,4glycosidic linkages. In this type of enzyme, however, the acid catalyst and the base catalyst are assumed to be different amino acid residues (Sinnott 1990). Conceivably, a similar approach to that employed in this study could be used to aid in the identification of the base-catalyst in an inverting glycosidase. Since azide was shown to increase the rate of the deglycosylation step in Cex (where base-catalysis occurs), anions such as azide should also enhance the rate of an uncharged base-catalyst mutant in an 99 inverting enzyme. In contrast to a retaining enzyme, the new product formed would be linked in an c’. configuration. The E127A mutant of Cex is potentially useful for several reasons. Firstly, the enzymatic production of 13-cellobiosyl-azide provides an alternative to the synthetic route (T. Lindhorst, personal communication); the mutant, as discussed, may allow the transfer of other anions to the sugar. Secondly, this mutant may be particularly suited for co-crystallization with a substrate such as DNPC since it tends to accumulate the glycosyl-en.zyme intermediate. The catalytic domain of Cex was crystallized in the absence of substrate (Bedarkar et a!., 1992). Examination of the enzyme/substrate complex should provide valuable information about the topology of the active site relative to its substrate and should confirm the presence of the covalently bound glycosyl-enzyme intermediate. Although the wild-type enzyme could be crystallized in the presence of the inactivator 2F-DNPC, DNPC presumably represents a more natural substrate. Lastly, enzymes which will transglycosylate to form oligosaccharides are potentially commercially valuable. Cex has the ability to carry out a weak transglycosylation reaction which essentially competes with the hydrolysis of the glycosyl-enzyme intermediate (Tull and Withers, 1994). Since the deglycosylation step of Glu 127 mutants of Cex with the appropriate substrate is slowed down, this mutant could serve as a starting point for the engineering of an enzyme capable of synthesizing oligosaccharides. In summary, based on amino acid sequence alignments and on knowledge of the types of residues likely to be involved in catalysis by a retaining glycanase, mutants have helped to identify the two key catalytic residues in Cex in the absence of three dimensional structural information. Preliminary X-ray crystallographic data analyzed following the preparation of this thesis in fact confirm the presence of Glu 233 and Glu 127 within the active site cleft of the enzyme (White et a!,, 1994). The two residues are 100 oriented such that their side-chains face each other with an appropriate separation. Based on the known structures of other retaining glycanases (Campbell et a!., 1993; Wakarchuk et a!., 1994) these residues are suitably positioned to act as the catalytic nudeophile and acid/base catalyst respectively. The Asp 123 side chain is not similarly exposed to the active site (White et a!., 1994). The approach used in identifying the acid/base catalyst in Cex in the absence of three dimensional information may in fact be applicable to other glycosidases, both retaining and inverting. This study will clearly serve to facilitate an understanding of the roles of active site residues in Cex, pending the resolution of the crystal structure. 101 6. Appendix 6.1 Determination of kcat and Km: an example: The example below shows the data obtained (as outlined in section 2.9) for the hydrolysis of 2,4-DNPG catalysed by Cex E127A (at 4 tg/mL) in the presence of 3 mM sodium azide: The initial rate of production of DNP 0 (v ) , measured by AA400/min, was converted to nmol DNP released/mm according to the following relationship: =A cL where: For DNP, at 2. 0 v = = E A c L 400 nm, pH 7.0, = nmol DNP released/mm extinction coefficient absorbance concentration (M) cell path length (cm) = = = = 10 900 1 .cm therefore (for a 1 cm path cell): M , AA400/min 0.01 09 = The following data were obtained [curve 11: 1.4 1.2 E 1 0.8 S [curve 1] 0.6 0 > 0.4 0.2 0 0 20 40 DNPG (M) 60 102 Assuming a steady state, the velocity of an enzyme-catalysed reaction can be expressed by the Michaelis-Menten equation: 0 V Vpx [S] Km + IS] = [3] where v 0 is the velocity (measured as the initial rate of production of DNP in this case), Vmax is the maximal velocity, Km is the Michaelis constant and [SI is the concentration of free substrate (DNPG). The rate constants and Km can be extracted graphically from a linearly-transformed plot of the data shown above. A common linearized form of the Michaelis-Menten equation is its reciprocal (the Lineweaver-Burk equation). When both sides of the Lineweaver-Burk equation are multiplied by [SI this yields: [5] [S] Vmax + V 0 V [4] Plotting [SI/v 0 vs. [SI results in a linear curve (a Hanes plot; curve 2), the slope of which is 1 /Vmax, the y-intercept = Km/Vmax and the x-intercept = -Km. 50 .E E ¶1 40 30 20 [curve 2] 10 -20-10 = 0 10 20 30 40 S (jiM DNPG) 50 0.9992 60 103 Therefore, from curve [2]: Km and = 9.57.i.M = 1.57 1 nmol.miir (for 4 pg enzyme) = 0.4 1 (for 1 mg enzyme) jimol.min The maximal velocity Vm is directly proportional to the total enzyme concentration [ET] such that Vm therefore kcat = = = = kcat [ET] Vmax [ET] 0.4 j.tmol.m1n 1 0.02122 i.tmol (where 1 mg enzyme = 2.122 x 10 2 iimol) 18.85 min 1 Values of kcat and Km reported in this thesis were calculated directly by non linear regression by fitting the data to the Michaelis-Menten equation (as shown in curve [11) using the computer program Grafit 2.0 (Erithacus Software Ltd., Staines, U.K.). In this case the values of keat and Km were determined to be 18.73 m1n 1 and 9.35 .tM respectively, in close agreement with the values obtained above. 104 6.2 Interpretation of kcat, Km and kcat/Km: glycosylation 1 k E+S deglycosylation 2 k ES 3 k E-G E+P 1 k-ROH +H20 Equation [1] is representative of the mechanism of Cex as shown in Figure 4.1. kcat, Km and kcat/Km can be expressed in terms of the rate constants of the individual steps as indicated in equation [1] (Fersht, 1985; Schowen, 1978; Sinnot, 1990). kcat = Km = kcat Km = 3 2 k k +k 2 3 [2] 1 ( 3 k k- +k ) 2 ( 1 k 2 ) 3 i-k k [3] 2 k 1 +k 1 k2 [4] When there is a rapid, reversible association step (formation of the E.S complex) and hydrolysis of the glycosyl-enzyme intermediate is rate limiting, as in the case of hydrolysis of cellobiosides by Cex (Tull and Withers, 1994), the kinetic relationships would be k-p>>k 2 and 3 >>k When these conditions are applied to equations [2], 2 k . and [4] above, the kinetic constants can be reduced to: kcat 3 k [51 13] 105 Km = = kcpt Km = = 1 k 3 k 2 1 k where k-i 1 k = Ks jKs) 2 k [6] 1 k 2 1 k k2 Ks [7] When there is a rapid, reversible association step and formation of the glycosyl enzyme intermediate is rate-limiting, as in the case of hydrolysis of glucosides by Cex (Tull and Withers, 1994) the kinetic relationships would be k1 this case, equations [2], >> k and k 3 >> . In 2 k [31 and [41 can be reduced to: kcat = Km = kcpt Km = = 2 k [81 k 1 1 k Ks [9] 1 k 2 k-i 2 k Ks [101 It can be seen from the above examples that kcat is a reflection of the rate of the rate determining step. It can also be seen from equations [7] and [101 that kcat/Km reflects the rate of the formation of the glycosyl-enzyme intermediate from the free enzyme and substrate. This applies whether glycosylation or deglycosylation is rate- 106 limiting. From equation [61, the Km value is proportional to k 2 / 3 k when deglycosylation is rate-limiting. This is consistent with the interpretation of Km as outlined in the introduction. 6.3 Nucleotide and amino acid sequence of mature (processed) Cex 1 1/1 GCG ACC ala thr 61/21 CCC AAC pro asn 121/41 CTC CCC val ala 181/61 CCC CCC gly ala 241/81 ACG CTC thr leu 301/101 GAG ACC glu ser 361/121 TCC TCC ser trp 421/141 TTC CAC phe gin 481/161 CAC CCC asp pro 541/181 AAC TCC asn ser 601/201 CCC TTC gly phe 661/221 CCC TTC arg phe 721/241 CCC TCC pro ser 781/261 TCC ATC cys met 841/281 TCC CTC trp vai 31/11 ACC CTC AAG GAG CCC GCC GAC CCC GCC CCC thr leu lys glu ala ala asp gly ala gly 91/31 CCC CTC TCC GAG GCG CAG TAC AAC GCG ATC arg leu ser glu ala gin tyr lys ala lie 151/51 GAG AAC CCC ATC AAC TCG GAC CCC ACC GAG glu asn ala met lys trp asp ala thr glu 211/71 CGC CAC CGC CTC CCC ACC TAC CCC CCC CAC gly asp arg val ala ser tyr ala ala asp 271/91 GTA TGC CAC TCG CAG CTC CCC CAC TCC CCC val trp his ser gln leu pro asp trp ala 331/111 CCC ATG CTC AAC CAC CTC ACC AAG CTC CCC ala met val asn his val thr lys val ala 391/131 CAC CTC CTC AAC GAG CCC TTC CCC CAC CCC asp val val asn glu ala phe ala asp gly 451/151 CAC AAC CTC CCC AAC CCC TAC ATC GAG ACC gin lys leu gly asn gly tyr ile glu thr 511/171 ACC CCC AAC CTC TCC ATC AAC CAC TAC AAC thr ala lys leu cys ile asn asp tyr asn 571/191 CTC TAC CAC CTC CTC AAC CAC TTC AAC CCC leu tyr asp leu val lys asp phe lys ala 631/211 CAG TCC CAC CTC ATC CTC CCC CAC CTC CCC gin ser his leu ile val gly gin val pro 691/231 CCC CAC CTC CCC CTC CAC CTC CCC ATC ACC ala asp leu gly val asp val arg ile thr 751/251 CAC CCC ACC AAC CTC CCC ACC CAC CCC CCC asp ala thr lys leu ala thr gin ala ala 811/271 CAC CTC ACC CCC TCC CAC CCC CTC ACC CTC gin val thr arg cys gin gly val thr val 871/291 CCC CAC CTC TTC CCC CCC GAG CCC CCC CCC pro asp val phe pro gly glu gly ala ala CCC GAC TTC GGC TTC CCC CTC GAC arg asp phe gly phe ala leu asp GCC GAC ACC GAG TTC AAC CTC GTC ala asp ser glu phe asn leu val CCC TCG CAC AAC AGC TTC TCC TTC pro ser gin asn ser phe ser phe ACC GCC AAC GAG CTG TAC GGC CAC thr gly lys glu leu tyr gly his AAC AAC CTC AAC CCC TCC CCC TTC lys asn leu asn gly ser ala phe GAC CAC TTC CAC CCC AAC CTC CCC asp his phe glu gly lys vai ala CCC CCC CCC CCC CAC GAC TCC CCC gly gly arg arg gln asp ser ala CCC TTC CCC CCC CCA CCT CCC CCC ala phe arg ala ala arg ala ala CTC CAC CCC ATC AAC CCC AAC ACC val glu gly ile asn ala lys ser CCC CCC CTC CCC CTC GAC TCC CTC arg gly val pro leu asp cys val CCC CAC TTC CCC CAC AAC CTC CAC gly asp phe arg gin asn ieu gin GAG CTC CAC ATC CCC ATC CCC ACC glu leu asp ile arg met arg thr CAC TAC AAC AAC GTC CTC CAC CCC asp tyr lys lys val vai gin ala TCC CCC ATC ACC CAC AAC TAC TCC trp gly ile thr asp lys tyr ser CTC CTG TGG CAC CCC ACC TAC CCC leu val trp asp ala ser tyr ala 107 901/301 AAG AAG CCC lys lys pro 961/321 ACC ACG CCC thr thr pro 1021/341 TGC CAG GTG cys gin val 1081/361 AAG AAC ACG iys asn thr 1141/381 CAG CAG GTC gin gin vai 1201/401 CCC AAC CCC arg asn ala 1261/421 GGC TCC CAC gly ser his 1321/441 ACG GTC CCC thr vai giy 931/311 CCC TAC CCC CCC GTG ATC GAG CCC TTC CCC ala tyr ala ala val met glu ala phe gly 991/331 ACC CCG ACC CCC ACC ACG CCC ACG CCC ACC thr pro thr pro thr thr pro thr pro thr 1051/351 CTC TGG CCC CTC AAC CAG TGG AAC ACC CCC ieu trp gly val asn gin trp asn thr gly 1111/371 TCC TCC GCT CCC CTC GAC CCC TGC ACC CTC ser ser ala pro val asp gly trp thr leu 1171/391 ACC CAC CCC TGC AGC TCC ACC GTC ACC CAC thr gin ala trp ser ser thr val thr gin 1231/411 CCC TGG AAC CCC TCC ATC CCC CCC CCC CCC pro trp asn gly ser iie pro ala gly gly 1291/431 ACC CCC ACC AAC CCC CCC CCC ACC CCC TTC thr gly thr asn ala ala pro thr ala phe TCA STOP GCG AGC CCC ACG CCC ACG CCC ala ser pro thr pro thr pro CCG ACG TCC CCT CCC CCC CCC pro thr ser gly pro ala gly TTC ACC CCC AAC GTC ACC CTC phe thr ala asn val thr val ACC TTC ACC TTC CCC TCC CCC thr phe ser phe pro ser gly TCC CCC TCC CCC CTC ACC GTC ser gly ser ala val thr val ACC CCC CAG TTC CCC TTC AAC thr ala gin phe gly phe asn TCC CTC AAC CCC ACC CCC TCC ser leu asn gly thr pro cys (enuf is enuf!) ‘Adapted from O Neill et al. (1986b). Corrections to the original gene sequence have t been deposited to GenBank (New accession # L11080). Numbering refers to the nucleotide/amino acid number from the mature N-terminus. Glu 127 and Glu 233 are indicated in bold type. 108 7. References Anne, J. and Van Mellaert, L. (1993). Streptornyces lividans as host for heterologous protein production. FEMS Microbiol. Lett. 114: 121-128. Atkinson, T. and Smith, M. (1984). Solid phase synthesis of oligodeoxyribonucleotides by the phosphite triester method. In Oligonucleotide synthesis: a practical approach. N.J. Gait (ed.) IRL Press, Oxford. pp. 35-81. Baird, S., Hefford, M.A., Johnson, D.A., Sung, W.L., Yaguchi, M., and Seligy, V. (1990). The Glu residue in the conserved Asn-Glu-Pro sequence of two highly divergent endo f3-1,4-glucanases is essential for enzymatic activity. Biochem. Biophys. Res. Commun. 169: 1035-1039. Bedarkar, S., Gilkes, N.R., Kilburn, D.G., Kwan, E., Rose, D.R., Miller Jr., R.C., Warren, R.A.J., and Withers S.G. (1992). Crystallization and preliminary X-ray diffraction analysis of the catalytic domain of Cex, an exo-3-1,4-glucanase and 3-l,4-xylanase from the bacterium Cellulornonas fimi. J. Mol. Biol. 228: 693. Beguin, P. and Aubert, J-P. (1994). The biological degradation of cellulose. FEMS Microbiol. Rev 13: 25-58 Blackwell, J. (1981). The structure of cellulose and chitin. In Biomolecular Structure, Conformation, Function and Evolution. R. Srinivasan (ed.) Pergamon, New York. Vol.1 Diffraction and Related Studies. pp. 523-535. Blake, C.C. F., Koenig, D.F., Mair, G.A., North, A.C.T., Phillips, D.C. and Sarma, V. R. (1965). Nature 206: 757-783. Bradford, M. (1976). A rapid and sensitive method for the quantitation of microgram quantities of proteins utilizing the principle of protein-dye binding. Anal. Biochem. 72: 248-254. Brawner, M., Poste, G., Rosenberg, M. and Westpheling, 1. (1991). Streptomyces: a host for heterologous gene expression. Current Opinion Biotechnol. 2: 674-681. Campbell, R., Rose, D., Wakarchuk, W., To, R., Sung, W. and Yagachi, M. (1993) In Proceedings of the second TRICEL symposium on Trichoderma reesei cellulases and other hydrolases (Suominin, P., and Reinikainen, T., ED.) Helsinki: Foundation for Biotechnical and Industrial Fermentation Research, Espoo, Finland. pp. 63-72 Coutinho, J.B., Moser, B., Kilburn, D.G., Warren, R.A.J. and Miller, R.C., Jr. (1991). Nucleotide sequence of the endoglucanase C gene (cenC) of Cellulomonas fimi, its high- 109 level expression in Escherichia coli, and characterization of its products. Mol. Microbiol. 5: 1221-1233. Cupples, C.G., Miller, J.H., and Hubert, R.E. (1990). Determination of the roles of Glu 461 in 3-galactosidase (Escherichia coli) using site-specific mutagenesis. I. Biol. Chem. 265: 5512. Curry, C., Gilkes, N., O’Neill G., Miller Jr., R.C. and Skipper, N. (1988). Expression and secretion of a Cellulomonas fimi exoglucanase in Saccharornyces cerevisiae, Appl. Environ. Microbiol. 54:476-484. Fersht, A. 1985. Enzyme Structure and Mechanism. 2nd Ed. W. H. Freeman and Co., New York. Gebler, J.C., Aebersold, R., and Withers, S.G. (1992a). Glu-537, not Glu-461 is the nucleophile in the active site of (lacZ) 13-galactosidase from Escherichia coli. J. Biol Chem 267(16): 11126-11130. Gebler, J., Gilkes, N.R., Claeyssens, M., Wilson, D.B., Beguin, P., Wakarchuk, W.W., Kilburn, D.G., Miller, R.C., Jr., Warren, R.A.J., and Withers, S.C. (1992b). Stereoselective hydrolysis catalysed by related 3-1,4-g1ucanases and 13-1,4-xylanases. J. Biol. Chem. 267: 12559. Ghangas G.S. and Wilson, D.B. (1988). Cloning of Thermornonospora fusca endoglucanase E2 gene in Streptornyces lividans: affinity purification and functional domains of the cloned gene product. Appi. Environ. Microbiol. 54: 2521-2526. Ghangas G.S., Hu, Y.J. and Wilson, D.B. (1989). Cloning of a Thermomonospora fusca xylanase gene and its expression in Escherichia coli and Streptomyces lividans. J. Bacteriol. 171: 2963-2969. Gilkes N.R., Kilburn, D.C., Langsford, M.L., Miller Jr., R.C., Wakarchuk, W.W., Warren, R.A.J., Whittle, D.J. and Wong, W.K.R. (1984a). Isolation and characterization of Escherichia coli clones expressing cellulase genes from Cellulornonas firni. J. Cen. Microbiol. 130: 1377-1384. Gilkes, Langsford, M.L., Kilburn, D.G., Miller Jr., R.C. and Warren, R.A.J. (1984b). Mode of action and substrate specificities of cellulases from cloned bacterial genes. Biol. Chem. 259: 10455-10459. J. Gilkes, N.R., Warren, R.A.J., Miller Jr., R.C. and Kilburn, D.C. (1988). Precise excision of the cellulose binding domains from two Cellulomonas firni cellulases by a homologous protease and the effect on catalysis. J. Biol. Chem. 263: 10401-10407. 110 Gilkes, N.R., Kilburn, D.G., Miller, R.C., Jr. and Warren, R.A.J. (1989). Structural and functional analysis of a bacterial cellulase by proteolysis. J. Biol. Chem. 264: 1780217808. Gilkes, N.R., Claeyssens, M., Aebersold, R., Henrissat, B., Meinke, A., Morrison, H.D., Kilburn, D.G., Warren, R.A.J. and Miller, R.C., Jr. (1991a). Structural and functional relationships in two families of 13-1,4-glycanases. Eur. J. Biochem. 202: 367-377. Gilkes, N.R., Henrissat, B., Kilburn, D.G., Miller, R.C., Jr., and Warren, R.A.J. (1991b). Domains in microbial 13,1,4-glycanases: sequence conservation, function and enzyme families. Microbiol. Rev. 55(2): 303-315. Goochee, C.F., Gramer, M.J., Andersen, D.C., Bahr J.B. and Rasmussen J.R. (1991). The oligosaccharides of glycoproteins: biopro cess factors affecting oligosaccharide structure and their effect on glycoprotein properties. Bio/Technology 9: 1347-1355. Grépinet, 0., Chebrou, M.-C. and Béguin, P. (1988). Nucleotide sequence and deletion analysis of the xylanase gene (xynZ) of Clostricliurn thermocellum. J. Bacteriol. 170: 4582-4588. Gusek T.W. and Kinsella, J.E. (1992). Review of the Streptornyces lividans/ Vector p1J702 system for gene cloning. Critical Rev. Microbiol. 18(4): 247-260. Haas, H., Herfurth, E., Stoffler, G. and Redi, B. (1992). Purification, characterization and partial amino acid sequences of a xylanase produced by Pen iciflium chrysogenum. Biochim. Biophys. Acta 1117: 279-286. Hanahan, D. (1983). Studies on transformation of Escherichia coli with plasmids. Biol. 166: 557-580. J. Mol. Hardy, L.W. and Poteete, A.R. (1991). Reexamination of the role of Asp2O in catalysis by bacteriophage T4 lysozyme. Biochemistry 30: 9457. Henrissat B. (1991). A classification of glycosyl hydrolases based on amino acid sequence similarities. Biochem J. 280: 309. Henrissat B., Claeyssens, M., Tomme P., Lemesle L. and Mornon, J.-P. (1989). Cellulase families revealed by hydrophobic cluster analysis. Gene 81(1): 83-95 Hopwood, D.A., Bibb, M.J., Chater, K.F., Kieser, T., Bruton, C.J., Kieser, H.M., Lydiate, D.J., Smith, C.P., Ward, J.M. and Schrempf, H. (1985). Genetic Manipulations of Streptomyces.: A Laboratory Manual. The John Innis Foundation, Norwich. 111 Hu, P., Chase, T. Jr. and Eveleigh, D.E. (1993). Cloning of a Microbispora bispora cellobiohydrolase gene in Streptomyces lividans. Appi. Microbiol. Biotechnol. 38: 631- 637. Huber, R.E. and Chivers, P.T. (1993). 3-ga1actosidases of Escherichia coli with substitutions for Glu-461 can be activated by nucleophiles and can form 13-D-galactosyl adducts. Carbohy. Res. 250: 9-18. Kendall K. and Cullum J. (1984). Cloning and expression of an extracellular-agarase gene from Streptomyces coelicolor A3(2) in Streptornyces lividans. Gene 29: 315-321 Kieser, T. and Hopwood, D.A. 1991. Genetic manipulations of Streptomyces: integrating vectors and gene replacement. Methods Enzymol. 204: 430-458. Kluepfel D., Vats-Mehta, S., Aumont, F., Shareck, F. and Morsoli R. (1990). Purification and characterization of a new xylanase (xylanase B) produced by Streptornyces lividans 66. Biochem. J. 267: 45-50. Koshland, D.E. (1953). Stereochemistry and the mechanism of enzymatic reactions. Biol. Rev. Camb. Philos. Soc. 28: 416-436. Kristensen T., Voss, H., and Ansorge, W. (1987). A simple and rapid preparation of M13 sequencing templates for manual and automated dideoxy sequencing. Nucleic Acids Res. 15: 5507. Kunkel, A.T., J.D. Roberts and R.A. Zakour. 1987. Rapid and efficient site-specific mutagenesis without phenotypic selection. Methods Enzymol. 154: 367-382. Laemmli, U.K. (1970). Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 227: 680-685. Langsford, M.L., Gilkes, N.R., Wakarchuk, W.W., Kilburn, D.G., Miller Jr., R.C. and Warren R.A.J. (1984). The cellulase system of Cellulomonas firni. J. Gen. Microbiol. 130: 1367-1376 Langsford, M.L., Gilkes, N.R., Singh, B., Moser, B., Miller Jr., R.C., Warren, R.A.J. and Kilburn, D.G. (1987). Glycosylation of bacterial cellulases prevents proteolytic cleavage between functional domains. FEBS Lett. 225: 163-167. Legler, G. (1990). Glycoside hydrolases: mechanistic information from studies with reversible and irreversible inhibitors. Adv. Carbohydr. Chem. Biochem 48: 362 112 Lin, L.-L. and Thomson, J.A. (1991). Cloning, sequencing and expression of a gene encoding a 73 kDa xylanase enzyme from the rumen anaerobe B utyrivibrio fibrisolvens H17c. Mo!. Gen. Genet. 228: 55-61. Luthi, E., Love D.R., McAnulty, J., Wallace, C., Caughey, P.A., Saul, D. and Berquist, P.L. (1991). Cloning, sequence analysis and expression of genes encoding xylan degrading enzymes from the thermophile “Caldocellum saccharolyticum”. Appi. Environ. Microbiol. 57: 694-700. MacLeod, A.M., Gilkes, N.R., Escote-Carison, L., Warren, R.A.J., Kilburn D.G. and R.C. Miller Jr. (1992). Streptomyces lividans glycosylates an exoglucanase (Cex) from Cellulomonas fimi. Gene 121: 143-147. MacLeod, A.M., Lindhorst, T., Withers, S.G. and Warren, R.A.J. (1994). The acid/base catalyst in the exoglucanase/xylanase (Cex) from Cellulornonas fimi is glutamic acid 127: evidence from detailed kinetic studies of mutants. Biochemistry 33: 6371-6376. Mannarelli, B.M., Evans, S. and Lee, D. (1990). Cloning, sequencing and expression of a xylanase gene from the anaerobic ruminal bacterium Butyrivibrio fibrisolvens. J. Bacteriol. 172:4247-4254. Mead, D.A. Szczesha-Skorupa, E., and Kemper, B. (1986). Single-stranded DNA ‘blue’ T7 promoter plasmids: a versatile tandem promoter system for cloning and protein engineering. Protein Eng. 1: 67. Meinke, A., Braun, C., Gilkes, N.R., Kilburn, D.G., Miller, R.C., Jr. and Warren, R.A.J. (1991). Unusual sequence organization in CenB, an inverting endoglucanase from Cellulomonas fimi. J. Bacteriol. 173: 308-314. Meinke, A., Gilkes, N.R., Kilburn, D.G., Miller, R.C., Jr. and Warren, R.A.J. (1993). Cellulose-binding polypeptides from Cellulomonas firni: endoglucanase D (CenD), a family A 13-1,4-glucanase. J. Bacteriol. 175(7): 1910-1918. Meinke, A., Gilkes, N.R., Kwan, E., Kilburn, D.G., Warren, R.A.J. and Miller, R.C., Jr. (1994). Cellobiohydrolase A (CbhA) from the cellulolytic bacterium Cellulornonas fimi is a f3-1,4-exocellobiohydrolase analogous to Trichoderma reesei CBH II. Mol. Microbiol. 12(3) 413-22. Mihoc, A. and Kluepfel, D. (1990). Purification and characterization of a f3-glucosidase from Streptomyces lividans 66. Can. J. Microbiol. 36: 53-56. Moser, B., Gilkes, N.R., Kilburn, D.G., Warren, R.A.J. and Miller Jr., R.C. (1989). Purification and characterization of endoglucanase C of Cellulomonas fimi, cloning of the 113 gene, and analysis of in vivo transcripts of the gene. Appi. Environ. Microbiol. 55: 24802487. Nakai, R., Horinouchi, S. and Beppu, T. (1988). Cloning and nucleotide sequence of a cellulase gene, casA, from an alkalophilic Streptornyces strain. Gene 65: 229-238. O’Neill G.P., Kilburn, D.G., Warren, R.A.J. and Miller Jr., R.C. (1986a). Overproduction from a cellulase gene with a high guanosine-plus-cytosine content in Escherichia coli. Appi. Environ. Microbiol. 52: 737-743. O’Neill G.P., Goh, S.H., Warren R.A.J., Kilburn, D.G. and Miller Jr., R.C. (1986b). Structure of the gene encoding the exoglucanase of Cellulornonas fimi. Gene 44: 325-330. O’Neill G.P., Warren, R.A.J., Kilburn, D.G. and Miller Jr., RC. (1986c). Secretion of a Cellulomonasfimi exoglucanase by E. coli. Gene 44: 331-336. Ong, E., Kilburn, D.G., Miller Jr., R.C. and Warren R.A.J. (1994). Streptomyces lividans glycosylates the linker region of a 13-1,4-glycanase from Cellulornonas fimi. J. Bacteriol. 176(4): 999-1008. Pickersgill, R.W. Jenkins, J.A., Scott, M., Connerton, I., Hazlewood, G.P. and Gilbert H.J. (1993). Crystallization and preliminary X-ray analysis of the catalytic domain of Xylanase A from Pseudomonas fluorescens subspecies cellulosa. J. Mol. Biol. 229(1): 246-8. Rennell, D., Bouvier, S.E., Hardy, L.W. and Poteete, A.R. (1991). Systematic mutation of T4 lysozyme. J. Mol. Biol. 222: 67. Rost B., Schneider R., and Sander C. (1993). Progress in protein structure prediction. TIBS 18(4): 120-123. Sambrook, J., Fritsch, E.F. and Maniatis, T. (1989). Molecular Cloning: A Laboratory Manual, 2nd ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Sandercock, L.E., MacLeod, A.M., Ong, E. and Warren RA.J. (1994). Non-S-layer glycoproteins in eubacteria. FEMS Microbiol. Lett. 118: 1-8. Sawadogo M. and Van Dyke, M.W. (1991). A rapid method for the purification of deprotected oligodeoxynucleotides. Nuc. Acids Res. 19(3): 674. Settineri, W. J. and Marchessault, R.H. (1965). Derivation of possible chain conformations for poly(f3-1,4-anhydroxylose). J. Polym. Sci. Part C 11: 253-264. 114 Schagger, H. and Von Jagow G. (1987). Tricine-sodium dodecyl sulfate-polyacrylamide gel electrophoresis for the separation of proteins in the range from 1 to 100 kDa. Anal. Biochem. 166: 368-379. Schimmel, P. (1990). Hazards and their exploitation in the applications of molecular biology to structure-function relationships. Biochemistry. 29: 9495-9502. Schowen, R.L. (1978). In Transition States of Biocehmical Processes (Candour R. D. and Schowen, R.L., eds.), Plenum Press, New York. pp. 77-114 Scopes, R.K. (1974). Measurement of protein by spectrophotometry at 205 nm. Anal. Biochem. 59: 277-282. Shareck, F., Roy, C., Yaguchi, M., Morosoli, R. and Kluepfel D. (1991). Sequences of three genes specifying xylanases in Streptomyces lividans. Gene 107: 75-82. Shen H., Tomme, P., Meinke, A., Gilkes, N.R., Kilburn, D., Warren, R.A.J. and Miller Jr., R.C. (1994) Biochem. Biophys. Res. Commun. 199(3): 1223-1228. Sinnott, M.L. (1990). Catalytic mechanisms of enzymic glycosyl transfer. Chem Rev. 90: 1171. Souchon H, Spinelli, S., Beguin P. and Aizari, P.M. (1994). Crystallization and preliminary diffraction analysis of the catalytic domain of xylanase Z from Clostridium thermocellum. J. Mol. Bio. 235(4): 1348-50. Stakebrandt E. and Woese, C.R. (1981). Towards a phylogeny of the actinomycetes and related organisms. Curr. Microbiol. 5: 197-202. Steiert J.G., Pogell, B.M., Speedie, M.K. and Laredo, J. (1989). A gene coding for a membrane-bound hydrolase is expressed as a secreted, soluble enzyme in Streptornyces lividans. Bio/Technology 7: 65-68. Tabor, S. and Richardson, C.C. (1987). DNA sequence analysis with a modified bacteriophage T7 DNA polymerase. Proc. Natl. Acad. Sci. USA. 84: 4767-4771. Taguchi, S., Kumagai, I., Nakayama, J., Suzuki, A. and Miura, K. (1989). Efficient extracellular expression of a foreign protein in Streptornyces using secretory protease inhibitor (SIT) gene fusions. Bio/Technology 7: 1063-1066 Towbin, H., Staehelin, T. and Gordon, J. (1979). Electrophoretic transfer of proteins from polyacrylamide gels to nitrocellulose sheets: procedure and some applications. Proc. Nati. Acad. Sci. USA 76:4350-4354. 115 Tull, D., Withers, S.G., Gilkes, N.R., Kilburn, D.C., Warren, R.A.J., and Aebersold, R. (1991). Glutamic acid 274 is the nucleophile in a retaining exoglucanase (Cex) from Cellulomonasfimi. J. Biol. Chem. 266: 15621 Tull, D. and Withers, S.G. (1994). Mechanisms of cellulases and xylanases: A detailed kinetic study of the exo-13-1,4-glycanase from Cellulornonas fimi. Biochemistry 33: 63636370. Vieira, J. and Messing, J. (1987). Production of single-stranded plasmid DNA. Methods Enzymol. 160: 3-11. Wakarchuk. W.W., Campbell, R.L., Sung. W.L., Davoodo, J. and Yaguchi. M. (1994). Mutational and crystallographic analyses of the active site residues of the Bacillus circulans xylanase. Protein Science 3: 467-475. Nhite, A., Withers, S.G., Gilkes, N.R and Rose, D.R. Crystal structure of the catalytic domain of the f3-1,4-glycanase Cex from Cellulomonas fimi. Biochemistry (in press). Withers, S. G., Dombroski, D., Berven, L.A., Kilburn, D.G., Miller, R.C., Jr., Warren, R.A.J. and Gilkes, N.R. (1986). Direct 1 H N.M.R. determination of the stereochemical course of hydrolysis catalysed by glucanase components of the cellulase complex. Biochem. Biophys. Res. Commun. 139: 487. Withers, S.G., and Street, I. (1988). Identification of a Covalent cL-D-glucopyranosyl enzyme intermediate formed on a 3-glucosidase. J. Am. Chem. Soc. 110 (25): 8551-8553. Withers, S.G., Warren, R.A.J., Street, I.P., Rupitz, K., Kempton, J.B. and Aebersold, R. (1990). Unequivocal demonstration of the involvement of a glutamate residue as a nucleophile in the mechanism of a ‘ retaining’ glycosidase. J. Am. Chem. Soc. 112: 15 t 5887-5889. Withers S.G., Rupitz, K., Trimbur, D. and Warren, R.A.J. (1992). Mechanistic consequences of mutation of the active site nucleophile Glu 358 in Agrobacterium f3glucosidase. Biochemistry 31: 9979-9985 Wong, W.K.R., Gerhard, B., Guo, Z.M., Kilburn, D.C., Warren, R.A.J. and Miller Jr., R.C. (1986). Characterization and structure of an endoglucanase gene cenA of Cellulomonas fimi. Gene 44: 3 15-324. Yanisch-Perron, C., Vieira, J. and Messing, J. (1985). Improved M13 phage cloning vectors and host strains: nucleotide sequences of the M13mp and pUC vectors. Gene 33: 103-119. 116 Yang, J.L., Luo G. and Eriksson, K.E.L. (1992). The impact of xylanase on bleaching of kraft puips. TAPPI 75: 95-101. Zacharius, R.M., Zell, T.E., Morrison, J.H. and Woodlock, 1.1. (1969). Glycoprotein staining following electrophoresis on acrylamide gels. Anal. Biochem. 30: 148-152. Zvelebil, M.J.J.M. and Sternberg, M.J.E. (1988). Analysis and prediction of the location of catalytic residues in enzymes. Protein Eng. 2: 127-138.