of - pnasthe enzyme used as a model, yeast triosephosphate isomerase (tim) (mr 53,000, two identical...

5
Proc. Nati. Acad. Sci. USA Vol. 84, pp. 675-679, February 1987 Biochemistry Control of oligomeric enzyme thermostability by protein engineering (protein stability/irreversible enzyme thermoinactivation/thermostable enzymes/site-directed mutagenesis) TIM J. AHERN*t, JOSE I. CASAL*t, GREGORY A. PETSKO§, AND ALEXANDER M. KLIBANOV*¶ *Department of Applied Biological Sciences and §Department of Chemistry, Massachusetts Institute of Technology, Cambridge, MA 02139 Communicated by Robert A. Alberty, October 1, 1986 (received for review August 15, 1986) ABSTRACT The ability to control the resistance of an enzyme to inactivation due to exposure to elevated tempera- tures is essential for the understanding of thermophilic behav- ior and for developing rational approaches to enzyme stabili- zation. By means of site-directed mutagenesis, point mutations have been engineered in the dimeric enzyme yeast triose- phosphate isomerase that improve its thermostability. Cumu- lative replacement of asparagine residues at the subunit inter- face by residues resistant to heat-induced deterioration and approximating the geometry of asparagine (Asn-14 -* Thr-14 and Asn-78 -- Ile-78) nearly doubled the half-life of the enzyme at 100°C, pH 6. Moreover, in an attempt to model the deleterious effects of deamidation, we show that replacement of interfacial Asn-78 by an aspartic acid residue increases the rate constant of irreversible thermal inactivation, drastically de- creases the reversible transition temperature, and reduces the stability against dilution-induced dissociation. Enzyme inactivation at elevated temperatures can be divided into two distinct classes (1). Reversible loss of activity is due to disruption of the native conformation above the melting transition; as the name implies, the activity is regained when the enzyme preparation is subsequently cooled. Irreversible loss of activity is the result of deleterious changes in the enzyme structure that are not undone by reducing the temperature. We have recently elucidated the processes causing the irreversible thermoinactivation of enzymes: depending on the pH, inactivating reactions may include deamidation of asparagine residues, hydrolysis of the polypeptide chain at aspartic acid residues, p elimination of cystine residues, as well as certain conformational processes (2). For example, at 100°C and pH 6, hen egg white lysozyme irreversibly inac- tivates due to a single reaction-deamidation of asparagine residues. The general nature of the mechanisms uncovered in lysozyme has been corroborated in studies on thermostability of bovine pancreatic ribonuclease A (3) and bacterial a- amylases (4). Knowledge of the chemical processes respon- sible for irreversible thermoinactivation affords a straight- forward enzyme stabilization strategy: replacement of the "weak links" in the protein molecules with other, more thermoresistant amino acid residues. This strategy was realized in the present study by means of site-directed mutagenesis (5-9). The enzyme used as a model, yeast triosephosphate isomerase (TIM) (Mr 53,000, two identical subunits, no S-S bonds and nonprotein components), has been cloned and expressed in Escherichia coli (10); its complete three-dimensional structure has been determined by x-ray diffraction (11), and site-directed mutagenesis has been successfully implemented for the study of its catalytic properties (12). MATERIALS AND METHODS Materials. All reagents were of the highest grade available and were obtained from commercial sources. Phage M13mpl8 and plasmid pUC18 were generously supplied by J. Messing (Waksman Institute of Microbiology, Rutgers Uni- versity) (13). The E. coli strain JM101 (Alac, pro, SupE, thil, F', proAB+, laci+, lacz miS tra A36) was a gift of R. Reilly, and the strain DF 502 [A(rha, pfkA, tpi) pfkBl, his-, pyrD-, edd-, F-, strr] was a construct of D. Fraenkel (11). Minimal medium is M9 medium (14) supplemented with 0.001% vitamin B1 and 1% glycerol as sole carbon source. Rich medium is LB medium (15) supplemented with 50 ,ug of ampicillin per ml, 1% glycerol, and 1% lactose as inducer. Construction of Mutant Genes. Details of the construction of the system used for the expression of yeast TIM in E. coli and for the mutagenesis of the yTPI gene, which encodes TIM of the yeast Saccharomyces cerevisiae, are given in ref. 16. In summary, the vector used to express the wild-type and mutant proteins was pUC18 containing yTPI [from a frag- ment of pTPIclO (10)] under the control of the lac operon. After transformation by the engineered vector, an E. coli strain (DF502) lacking the bacterial TPI gene expressed levels of yeast TIM as high as 1-1.5 mg/liter of culture. For purposes of mutagenesis, a fragment bearing the TPI gene was excised from pUC18 by endonucleases and subsequently inserted into the double-stranded replicative form of M13mpl8. The replicative form containing the yTPI gene was used to transfect E. coli JM101 cells that were grown in minimal medium to yield the single-stranded M13mpl8 for use in mutagenesis. The indicated replacements of asparagine at the interface of yeast TIM were produced by the double primer method of mutagenesis (17). The mismatched oligonucleotide primers used were synthesized by means of solid-phase phosphotri- ester chemistry by a DNA synthesizer (Biosearch model SAM 1) for annealing to a single-stranded template of DNA from M13mpl8 phage containing the yTPI gene. Following purification and sequencing (18), the mutated gene was excised and inserted into the modified pUC18 expression vector used to transform competent DF502 E. coli cells. The identity of the inserted mutant gene was verified by sequenc- ing, and the mutant TIM was obtained from cultures of the transformed cells and purified (16). Wild-type yeast TIM produced by this expression system was indistinguishable from commercially available yeast TIM (type 1, Sigma) with respect to molecular weight, isoelectric point, amino acid composition, specific activity, and rate constant of irrevers- ible thermoinactivation. Abbreviations: TIM, triosephosphate isomerase; yTPI, gene coding for TIM of the yeast S. cerevisiae. tPresent address: Genetics Institute, 87 Cambridge/Park Drive, Cambridge, MA 02140. tPresent address: Ingenasa, Hermanos Garcia Noblejas 42, Madrid, Spain. STo whom reprint requests should be addressed. 675 The publication costs of this article were defrayed in part by page charge payment. This article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. §1734 solely to indicate this fact. Downloaded by guest on August 8, 2021

Upload: others

Post on 10-Mar-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: of - PNASThe enzyme used as a model, yeast triosephosphate isomerase (TIM) (Mr 53,000, two identical subunits, no S-S bonds and nonprotein components), has been cloned and expressed

Proc. Nati. Acad. Sci. USAVol. 84, pp. 675-679, February 1987Biochemistry

Control of oligomeric enzyme thermostabilityby protein engineering

(protein stability/irreversible enzyme thermoinactivation/thermostable enzymes/site-directed mutagenesis)

TIM J. AHERN*t, JOSE I. CASAL*t, GREGORY A. PETSKO§, AND ALEXANDER M. KLIBANOV*¶*Department of Applied Biological Sciences and §Department of Chemistry, Massachusetts Institute of Technology, Cambridge, MA 02139

Communicated by Robert A. Alberty, October 1, 1986 (received for review August 15, 1986)

ABSTRACT The ability to control the resistance of anenzyme to inactivation due to exposure to elevated tempera-tures is essential for the understanding of thermophilic behav-ior and for developing rational approaches to enzyme stabili-zation. By means of site-directed mutagenesis, point mutationshave been engineered in the dimeric enzyme yeast triose-phosphate isomerase that improve its thermostability. Cumu-lative replacement of asparagine residues at the subunit inter-face by residues resistant to heat-induced deterioration andapproximating the geometry of asparagine (Asn-14 -* Thr-14and Asn-78 -- Ile-78) nearly doubled the half-life of the enzymeat 100°C, pH 6. Moreover, in an attempt to model thedeleterious effects of deamidation, we show that replacement ofinterfacial Asn-78 by an aspartic acid residue increases the rateconstant of irreversible thermal inactivation, drastically de-creases the reversible transition temperature, and reduces thestability against dilution-induced dissociation.

Enzyme inactivation at elevated temperatures can be dividedinto two distinct classes (1). Reversible loss of activity is dueto disruption of the native conformation above the meltingtransition; as the name implies, the activity is regained whenthe enzyme preparation is subsequently cooled. Irreversibleloss of activity is the result of deleterious changes in theenzyme structure that are not undone by reducing thetemperature.We have recently elucidated the processes causing the

irreversible thermoinactivation of enzymes: depending onthe pH, inactivating reactions may include deamidation ofasparagine residues, hydrolysis of the polypeptide chain ataspartic acid residues, p elimination of cystine residues, aswell as certain conformational processes (2). For example, at100°C and pH 6, hen egg white lysozyme irreversibly inac-tivates due to a single reaction-deamidation of asparagineresidues. The general nature of the mechanisms uncovered inlysozyme has been corroborated in studies on thermostabilityof bovine pancreatic ribonuclease A (3) and bacterial a-amylases (4). Knowledge of the chemical processes respon-sible for irreversible thermoinactivation affords a straight-forward enzyme stabilization strategy: replacement of the"weak links" in the protein molecules with other, morethermoresistant amino acid residues. This strategy wasrealized in the present study by means of site-directedmutagenesis (5-9). The enzyme used as a model, yeasttriosephosphate isomerase (TIM) (Mr 53,000, two identicalsubunits, no S-S bonds and nonprotein components), hasbeen cloned and expressed in Escherichia coli (10); itscomplete three-dimensional structure has been determinedby x-ray diffraction (11), and site-directed mutagenesis hasbeen successfully implemented for the study of its catalyticproperties (12).

MATERIALS AND METHODSMaterials. All reagents were of the highest grade available

and were obtained from commercial sources. PhageM13mpl8 and plasmid pUC18 were generously supplied by J.Messing (Waksman Institute of Microbiology, Rutgers Uni-versity) (13). The E. coli strain JM101 (Alac, pro, SupE, thil,F', proAB+, laci+, lacz miS tra A36) was a gift of R. Reilly,and the strain DF 502 [A(rha, pfkA, tpi) pfkBl, his-, pyrD-,edd-, F-, strr] was a construct of D. Fraenkel (11).Minimal medium is M9 medium (14) supplemented with

0.001% vitamin B1 and 1% glycerol as sole carbon source.Rich medium is LB medium (15) supplemented with 50 ,ug ofampicillin per ml, 1% glycerol, and 1% lactose as inducer.

Construction of Mutant Genes. Details of the constructionof the system used for the expression of yeast TIM in E. coliand for the mutagenesis of the yTPI gene, which encodesTIM of the yeast Saccharomyces cerevisiae, are given in ref.16. In summary, the vector used to express the wild-type andmutant proteins was pUC18 containing yTPI [from a frag-ment of pTPIclO (10)] under the control of the lac operon.After transformation by the engineered vector, an E. colistrain (DF502) lacking the bacterial TPI gene expressed levelsof yeast TIM as high as 1-1.5 mg/liter of culture. Forpurposes of mutagenesis, a fragment bearing the TPI genewas excised from pUC18 by endonucleases and subsequentlyinserted into the double-stranded replicative form ofM13mpl8. The replicative form containing the yTPI gene wasused to transfect E. coli JM101 cells that were grown inminimal medium to yield the single-stranded M13mpl8 foruse in mutagenesis.The indicated replacements of asparagine at the interface

of yeast TIM were produced by the double primer method ofmutagenesis (17). The mismatched oligonucleotide primersused were synthesized by means of solid-phase phosphotri-ester chemistry by a DNA synthesizer (Biosearch modelSAM 1) for annealing to a single-stranded template of DNAfrom M13mpl8 phage containing the yTPI gene. Followingpurification and sequencing (18), the mutated gene wasexcised and inserted into the modified pUC18 expressionvector used to transform competent DF502 E. coli cells. Theidentity of the inserted mutant gene was verified by sequenc-ing, and the mutant TIM was obtained from cultures of thetransformed cells and purified (16). Wild-type yeast TIMproduced by this expression system was indistinguishablefrom commercially available yeast TIM (type 1, Sigma) withrespect to molecular weight, isoelectric point, amino acidcomposition, specific activity, and rate constant of irrevers-ible thermoinactivation.

Abbreviations: TIM, triosephosphate isomerase; yTPI, gene codingfor TIM of the yeast S. cerevisiae.tPresent address: Genetics Institute, 87 Cambridge/Park Drive,Cambridge, MA 02140.tPresent address: Ingenasa, Hermanos Garcia Noblejas 42, Madrid,Spain.STo whom reprint requests should be addressed.

675

The publication costs of this article were defrayed in part by page chargepayment. This article must therefore be hereby marked "advertisement"in accordance with 18 U.S.C. §1734 solely to indicate this fact.

Dow

nloa

ded

by g

uest

on

Aug

ust 8

, 202

1

Page 2: of - PNASThe enzyme used as a model, yeast triosephosphate isomerase (TIM) (Mr 53,000, two identical subunits, no S-S bonds and nonprotein components), has been cloned and expressed

Proc. Natl. Acad. Sci. USA 84 (1987)

Production and Purification of TIM. All steps of purifica-tion were monitored by NaDodSO4/PAGE (15% acrylamide)using the buffer system of Laemmli (19). Approximately25-30 g (wet weight) of DF502 E. coli cells was harvestedfrom 9.9 liters of culture grown to 6 OD6w units in shakingflasks in rich medium. Cells were lysed by treatment withTriton X-100 (0.1%) and lysozyme (0.5 mg/ml). The lysatewas centrifuged at 30,000 rpm for 120 min in a Ti 45 rotor(Beckman) to remove cell debris. The 35-95% saturatedammonium sulfate fraction of the clarified lysate was dia-lyzed against 0.1 M phosphate buffer (pH 7.4) containing 0.15M NaCl [phosphate-buffered saline (PBS)] (3 x 2 liters). Theinsoluble material was subsequently removed by centrifuga-tion, and the supernatant was applied to a DEAE-cellulose(Sigma) column and eluted with PBS. The pooled fractionscontaining TIM activity were concentrated before subjectionto immunoadsorption chromatography on immobilized anti-bodies specific for yeast TIM (16). The protein was elutedwith 3 M KSCN and dialyzed immediately against 0.1 Mphosphate buffer (pH 6) (3 x 2 liters). The final enzyme yieldswere typically 10-12 mg of TIM of >95% purity, as deter-mined by NaDodSO4/PAGE, isoelectrofocusing, andnondenaturing molecular sieve chromatography (16).

Activity Assay for TIM. The enzyme-linked spectrophoto-metric assay based on the decrease of absorbance at 340 nmdue to the oxidation ofNADH (e = 6220 M-1 cm-1) was usedto determine the activity ofTIM (20). In the presence ofTIM,D-glyceraldehyde 3-phosphate is converted to dihydroxyac-etone phosphate, whose reduction to a-glycerol 3-phosphateis catalyzed by a-glycerophosphate dehydrogenase withconcomitant oxidation of NADH. Samples of TIM (10 juI)were mixed in a solution containing D-glyceraldehyde 3-phosphate (1 mM), 10 ug of a-glycerophosphate dehydro-genase per ml (type 1, Sigma), EDTA (5 mM), and NADH(0.1 mg/ml) in 20 mM triethanolamine hydrochloride buffer(pH 7.9). One unit of enzymatic activity is the conversion of1 umol of D-glyceraldehyde 3-phosphate to dihydroxyace-tone phosphate per minute.

Kinetic Studies. Kinetic measurements were carried outaccording to the method of Plaut and Knowles (21) in a model552 spectrophotometer coupled to a model 561 recorder(Perkin-Elmer) at 30'C maintained by a circulating constanttemperature bath. The cuvette contained 0.1 M buffer solu-tion (sodium cacodylate and triethanolamine hydrochloridefor pH 6 and 7.9, respectively), 5 mM EDTA, 0.2 mMNADH, 0.017 mg of a-glycerophosphate dehydrogenase perml (Sigma, dialyzed to remove ammonium sulfate), 0.3-2.0mM D-glyceraldehyde 3-phosphate, and 1-25 ng of TIM perml to initiate the reaction. The total volume was 1.5 ml. Thekinetic parameters kcat (calculated per dimer) and Km wereobtained from unweighted least-squares analysis of plots ofvo versus vo/[SO] (the slope ofwhich gives Km) and [SO] versus[So]/vo (the slope of which gives kcat), where vo is the initialvelocity and [SO] is the initial substrate concentration. Thetabulated values for Km were calculated on the basis that only4% of the substrate is reactive, since =96% is hydrated underthe conditions of the experiment (22).

Irreversible Thermoinactivation of TIM. The purified en-zymes (1.5 mg/ml) were incubated at 100°C in 0.1 Mphosphate buffer (pH 6) containing 6 M guanidine hydro-chloride and 8.4 mM EDTA; periodically, samples werewithdrawn, cooled, and diluted (1:400) prior to assay forenzymatic activity. In all cases, the correlation coefficientwas >0.98.

Deamidation of Asparagine/Glutamine Residues. The de-gree of deamidation ofTIM during heating was determined bythe colorimetric assay of Forman (23) for the release ofammonia and by summation of the deamidated forms of TIMseparated according to charge by isoelectrofocusing and

quantified by gel-scanning densitometry as described byAhern and Klibanov (2).

Destruction of Other Amino Acid Residues. The cysteineresidues of TIM were titrated before and after incubationaccording to the modified Ellman's procedure (24). Theamino acid composition was determined by subjecting heatedsamples of TIM to acid hydrolysis, derivatization witho-phthalaldehyde, and separation by means of high-perform-ance liquid chromatography (2).

Peptide Chain Integrity. Heated samples of TIM weresubjected to NaDodSO4/PAGE (19). After staining withCoomassie blue, the protein in the gel was quantified bygel-scanning densitometry.

RESULTS AND DISCUSSION

When an aqueous solution of TIM (1.5 mg/ml) is heated at100'C, pH 6, the enzyme rapidly aggregates and forms avisible precipitate. Aggregation is an inherently ill-definedprocess that obscures the underlying mechanism -of inacti-vation and can be prevented by the addition of a strongdenaturant (2). When TIM is heated briefly in the presence of6M guanidine hydrochloride and then cooled and diluted, theenzymatic activity fully returns; longer exposures to hightemperature result in a gradually decreasing recovery of theactivity of TIM. We have found that under these conditionsonly one process, deamidation of asparagine/glutamine res-idues (resulting in the evolution of ammonia and a decreasein the isoelectric point, Fig. 1), takes place in the enzyme

X5 0

\'5

E

75 20

'0~~

c50 '

0 30 60Time, min

FIG. 1. Time course of irreversible thermal inactivation andcovalent processes in yeast TIM at 1001C, pH 6. o, Residual activityof TIM (1.5 mg/ml) after heating in 0.1 M phosphate buffercontaining 6 M guanidine hydrochloride and 8.4mM EDTA. Sampleswere cooled to 230C at the times indicated and diluted (1:400) priorto assay. e, Fraction of nondeamidated yeast TIM during heating, asdetermined by isoelectrofocusing and gel-scanning densitometry.(Inset) Evolution of ammonia during incubation, either measureddirectly (23) (o) or calculated on the basis of the isoelectrofocusingdata (*). Other covalent processes-namely, hydrolysis of thepolypeptide chain and destruction of disulfide bonds-have beenreported to cause irreversible thermoinactivation of enzymes (2);however, under our conditions, no more than 6% of TIM hadundergone polypeptide chain hydrolysis after 30 mmn, as determinedby NaDodSO4/PAGE. All free cysteines in TIM (which has nodisulfide bonds) were titratable before and after incubation. Nodestruction of amino acid residues was detectable by amino acidanalysis of acid hydrolyzates of thermoinactivated TIM.

676 Biochemistry: Ahern et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 8

, 202

1

Page 3: of - PNASThe enzyme used as a model, yeast triosephosphate isomerase (TIM) (Mr 53,000, two identical subunits, no S-S bonds and nonprotein components), has been cloned and expressed

Proc. Natl. Acad. Sci. USA 84 (1987) 677

FIG. 2. Location of asparagine residuesin yeast TIM. (A) Stereo drawing of thepolypeptide chains (blue and purple) and theasparagine residues (green and red) of thetwo identical subunits of TIM. (B) Stereodrawing that shows the position of aspara-gine residues at the subunit interface. Theportions of the interface due to the aspara-gine (green) and non-asparagine (blue) resi-dues of one subunit and the Asn-14, -65, and-78 residues (red) of the complementarysubunit are depicted as dots of the samecolors. The interfacial contacts were calcu-lated by means of a molecular surface pro-gram utilizing a spherical probe of 1.6-Aradius (25). The images were generated on anEvans & Sutherland PS 300 graphics systemby means of the HYDRA program of R.Hubbard (University of York).

molecule. [It has been shown (2, 3) that the presence of thedenaturant does not affect appreciably the rate of thiscovalent process.] Such deamidation is known to inactivateenzymes-e.g., lysozyme, ribonuclease, and bacterial a-

amylases (4)-and had the same effect on TIM: the catalyticactivity of deamidated forms of TIM separated by chro-matofocusing decreased with the degree of deamidation.Hence, replacement of those asparagine residues whosedeamidation is detrimental to enzyme activity with groupsthat do not undergo thermal destruction should decrease therate of irreversible thermoinactivation. Likewise, replace-ment of asparagine by aspartic acid residues should model thedamaging effects of deamidation of specific asparagine resi-dues on enzyme activity and serves as a useful control.The refined crystal structure ofTIM reveals that 3 of each

subunit's 12 asparagine residues (Asn-14, -65, and -78) arepresent at the interfacial Van der Waals surface (Fig. 2).Deamidation of such residues results in the formation ofcharged aspartic acid residues, thereby promoting dissocia-tion to monomers (26), which have been shown to becatalytically inactive (27). Examination of the primary struc-tures of TIM from various organisms (28, 29) revealed thatsmall uncharged amino acid residues at these sites conserveisomerase activity and that the enzyme from the thermophileBacillus stearothermophilus has asparagine residues at noneof these positions (28).We therefore produced mutant forms ofTIM having single

replacements at Asn-78 (isoleucine, threonine, and, as acontrol, aspartic acid) and a double replacement (Thr-14 plusIle-78) by means of the double-primer mutagenesis method(17). The wild-type and mutant forms of the enzyme werepurified from cell extracts by immunoadsorption chromatog-raphy to >95% purity, as determined by NaDodSO4/PAGEand isoelectrofocusing (16). The purified enzymes wereeluted in the form of dimers by molecular sieve chromatog-raphy in 0.1 M cacodylate buffer (pH 6); the Km value forD-glyceraldehyde 3-phosphate (at 30°C) was little affected byasparagine replacement at the interface, whereas kcat (19,500sec-1, calculated per dimer) decreased somewhat (17,800,

17,200, 12,900, and 11,500 sec-1 for Thr-78, Ile-78, Asp-78,and Thr-14/Ile-78 forms, respectively). The decrease of kcatto -66% of that of wild type when Asn-78 was replaced byaspartic acid is comparable to the decrease in specific activityof monodeamidated enzymes produced by heat treatment(Table 1): a single, random deamidation on average appearsto decrease the specific activity to =50-70% that ofthe nativeenzyme.As shown in Table 2, replacement of Asn-78 by isoleucine

or threonine increased the half-life of the enzyme at 100'C(pH 6) with respect to irreversible inactivation by -25%.Cumulative replacement of interfacial Asn-14 and -78 resi-dues by threonine and isoleucine, respectively, resulted in a4-fold greater effect. In contrast, the stability of the controlmutant Asn-78 -- Asp-78 decreased relative to wild type.For comparison, the effect of single amino acid substitu-

tions at the interface on the other type of enzyme stability-namely, conformational stability with respect to reversible

Table 1. Effect of deamidation on the relative specific activityof enzymes

Relativespecific

Enzyme activity Ref.

Hen egg white lysozyme 2Native 1.00Monodeamidated 0.53Di- and trideamidated 0.21

Bovine pancreatic ribonuclease A 3Native 1.00Monodeamidated 0.65Dideamidated 0.38Trideamidated 0.19

Yeast TIM This workNative 1.00(Asn-78 -* Asp-78)* 0.66

*The altered enzyme was produced by means of site-directedmutagenesis and expressed in E. coli.

Biochemistry: Ahern et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 8

, 202

1

Page 4: of - PNASThe enzyme used as a model, yeast triosephosphate isomerase (TIM) (Mr 53,000, two identical subunits, no S-S bonds and nonprotein components), has been cloned and expressed

Proc. Natl. Acad. Sci. USA 84 (1987)

Table 2. Effect of replacement of interfacial amino acid residueson the half-life of yeast TIM at 1000C

Enzyme type t1/2, minWild typeAsn-14/Asn-78 13

StabilizedThr-78 17Ile-78 16Thr-14/Ile-78 25

Destabilized*Asp-78 11

The purified enzymes (1.5 mg/ml) were incubated at 100TC in 0.1M phosphate buffer (pH 6) containing 6 M guanidine hydrochlorideand 8.4 mM EDTA; periodically, samples were withdrawn, cooled,and diluted (1:400) prior to assay for enzymatic activity. In all cases,the correlation coefficient was >0.98.*An attempt to make a double mutant, Asp-14 plus Asp-78, yieldedvery low enzyme activity (<0.4% relative to other mutant forms)and no soluble enzyme upon purification, despite the presence ofmutant plasmid in the cells, as determined by plasmid sequencing(16).

denaturation-is depicted in Fig. 3A. As expected, the"neutral" single mutations Asn-78 -* Thr-78 and Asn-78 --

Ile-78, which stabilized TIM against irreversible thermoin-activation, had little effect on reversible denaturation; how-ever, a single substitution of Asn-78 by aspartic acid loweredthe transition temperature by as much as 15TC. (Assay at 23TCafter such incubations resulted in complete recovery of theoriginal activity, thus confirming the reversible nature ofdenaturation.) In addition, the mutant Asn-78 -- Asp-78 was

also unstable at 230C at mildly alkaline pH and low enzymeconcentrations (Fig. 3B). Since the mutant Asn-78 -- Asp-78

was more susceptible than wild-type TIM to proteolysis aswell (16), we speculate that an attempted double mutant,Asn-14/Asn-78 Asp-14/Asp-78, which could not be ex-

100

'5 50

A B

42 53 64 0 30 60T, 0C Time, min

FIG. 3. Conformational stability of wild-type and monosub-stituted mutant TIM: the effect of temperature (A) and dilution andpH (B) on the catalytic activity. (A) Enzyme (0.1 mg/ml) wasincubated in 24 mM triethanolamine buffer (pH 7.9) containing 6mMEDTA at the temperatures indicated for 10 min prior to assay at thesame temperature. *, Wild type; A, Asn-78 -. Thr-78; o, Asn-78 --

Ile-78; e, Asn-78 -- Asp-78. (B) Wild-type enzyme [pH 6 (o) and 7.9(.)] and mutant Asn-78 -- Asp-78 [pH 6 (o) and 7.9 (e)], 50 ng/ml,were incubated at 30°C in either cacodylate or triethanolamine buffer(0.1 M, 6 mM EDTA, pH 6 and 7.9, respectively) for the timesindicated prior to addition of substrate for assay. It is worth notingthat deamidation of asparagine residues in proteins may occur byway of a succinimide intermediate that can hydrolyze to yield anaspartic acid residue linked by way of either its a- or ,B-carboxylgroup to the polypeptide chain (30) (the products of which ourexperimental methodology would not distinguish). Our modeling ofthe deamidation process by site-directed mutagenesis demonstratesthat lowered stability with respect to reversible and irreversiblethermoinactivation results when asparagine is converted to asparticacid even without a -* A peptide bond rearrangement.

pressed in the bacterium, was so unstable that it wasintracellularly degraded during fermentation at 37TC (Table 2,footnote). All of these observations are consistent withreversible thermal inactivation being due to conformationalinstability, in contrast to the chemical nature of irreversiblethermoinactivation of this enzyme.Our findings demonstrate how certain amino acid substi-

tutions affect protein stability, specifically (i) the suscepti-bility to heat-induced "melting" of native conformation, thefirst step of essentially all thermoinactivation processes (1),and (ii) the resistance to irreversible thermoinactivation dueto covalent mechanisms. In addition to demonstrating thestabilizing effect of substitution of specific asparagine resi-dues, our findings with asparagine -* aspartic acid controlmutations illustrate the potential damaging effects of heat-induced deamidation and the conformational instability in-troduced by the presence of charged residues at normallyburied interfaces (31, 32). Rational strategies such as thoseemployed here are essential for redesigning enzymes for useas industrial catalysts and for establishing limiting rules ofenzyme thermostability.

We are grateful to R. Campbell, R. C. Davenport, and E. Lolis forhelp and advice. This work was financially supported by grants fromthe National Science Foundation (PCM-8316020 and DMB-8520721)and Sun Oil Company.

1. Klibanov, A. M. (1983) Adv. Appl. Microbiol. 29, 1-28.2. Ahern, T. J. & Klibanov, A. M. (1985) Science 228,

1280-1284.3. Zale, S. E. & Klibanov, A. M. (1986) Biochemistry 25,

5432-5444.4. Ahern, T. J. & Klibanov, A. M. (1986) UCLA Symp. Mol.

Cell. Biol. 39, 283-289.5. Razin, A., Hirose, T., Itakura, K. & Riggs, A. D. (1978) Proc.

Nati. Acad. Sci. USA 75, 4268-4270.6. Villafranca, J. E., Howell, H. E., Voet, D. H., Strobel, M. S.,

Ogden, R. C., Abelson, J. N. & Kraut, J. (1983) Science 222,782-788.

7. Wilkinson, A. J., Fersht, A. R., Blow, D. M., Carter, P. &Winter, G. (1984) Nature (London) 307, 187-188.

8. Fersht, A. R., Shi, J., Knill-Jones, J., Lowe, D. M., Wilkin-son, A. J., Blow, D. M., Brick, P., Carter, P., Waye,M. M. Y. & Winter, G. (1985) Nature (London) 314, 235-238.

9. Straus, D., Raines, R., Kawashima, E., Knowles, J. R. &Gilbert, W. (1985) Proc. Natl. Acad. Sci. USA 82, 2272-2276.

10. Alber, T. & Kawasaki, G. (1982) J. Mol. Appl. Genet. 1,419-434.

11. Alber, T., Banner, D. W., Bloomer, A. C., Petsko, G. A.,Phillip, R. S. D., Rivers, P. S. & Wilson, I. A. (1981) Philos.Trans. R. Soc. London B 293, 159-171.

12. Petsko, G. A., Davenport, R. C., Frankel, D. & RajBhandary,U. L. (1984) Biochem. Soc. Trans. 12, 229-232.

13. Norrander, J., Kempe, T. & Messing, J. (1983) Gene 26,101-106.

14. Miller, J. H. (1972) Experiments in Molecular Genetics (ColdSpring Harbor Laboratory, Cold Spring Harbor, NY).

15. Maniatis, T., Fritsch, E. F. & Sambrook, J. (1982) MolecularCloning: A Laboratory Manual (Cold Spring Harbor Labora-tory, Cold Spring Harbor, NY).

16. Casal, I. J., Ahern, T. J., Davenport, R. C., Petsko, G. A. &Klibanov, A. M. (1987) Biochemistry 26, in press.

17. Zoller, M. J. & Smith, M. (1984) DNA 3, 479-488.18. Sanger, F., Nicklen, S. & Coulson, A. R. (1977) Proc. Natl.

Acad. Sci. USA 74, 5463-5467.19. Laemmli, U. K. (1970) Nature (London) 227, 680-685.20. Hartman, F. C. & Norton, I. L. (1975) Methods Enzymol. 41,

447-453.21. Plaut, B. & Knowles, J. R. (1972) Biochem. J. 129, 311-320.22. Reynolds, S. J., Yates, D. W. & Pogson, C. J. (1971) Bio-

chem. J. 122, 285-297.23. Forman, D. T. (1964) Clin. Chem. (Winston-Salem, N.C.) 10,

497-508.24. Riddles, P. W., Blakely, R. L. & Zerner, B. (1983) Methods

Enzymol. 91, 49-61.

678 Biochemistry: Ahern et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 8

, 202

1

Page 5: of - PNASThe enzyme used as a model, yeast triosephosphate isomerase (TIM) (Mr 53,000, two identical subunits, no S-S bonds and nonprotein components), has been cloned and expressed

Biochemistry: Ahern et al.

25. Connoly, M. L. (1983) J. Appl. Crystallogr. 16, 548-558.26. Yuan, P. M., Talent, J. M. & Gracy, R. W. (1981) Mech.

Aging Dev. 17, 151-162.27. Waley, S. G. (1973) Biochem. J. 135, 165-172.28. Lu, H. S., Yuan, M. P. & Gracy, R. W. (1984) J. Biol. Chem.

259, 11958-11968.

Proc. Natl. Acad. Sci. USA 84 (1987) 679

29. Pichersky, E., Gottlieb, L. & Hess, J. F. (1984) Mol. Gen.Genet. 195, 314-319.

30. Clarke, S. (1985) Annu. Rev. Biochem. 54, 479-506.31. Perutz, M. F. (1978) Science 201, 1187-1191.32. Jones, D. H., McMillan, A. J., Fersht, A. R. & Winter, G.

(1985) Biochemistry 24, 5852-5857.

Dow

nloa

ded

by g

uest

on

Aug

ust 8

, 202

1