首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
N S Sampson  J R Knowles 《Biochemistry》1992,31(36):8488-8494
A residue essential for proper closure of the active-site loop in the reaction catalyzed by triosephosphate isomerase is tyrosine-208, the hydroxyl group of which forms a hydrogen bond with the amide nitrogen of alanine-176, a component of the loop. Both residues are conserved, and mutagenesis of the tyrosine to phenylalanine results in a 2000-fold drop in the catalytic activity (kcat/Km) of the enzyme compared to the wild-type isomerase. The nature of the closure process has been elucidated from both viscosity dependence and primary isotope effects. The reaction catalyzed by the mutant enzyme shows a viscosity dependence using glycerol as the viscosogen. This dependence can be attributed to the rate-limiting motion of the active-site loop between the "open" and the "closed" conformations. Furthermore, a large primary isotope effect is observed with [1-2H]dihydroxyacetone phosphate as substrate [(kcat/Km)H/(kcat/Km)D = 6 +/- 1]. The range of isotopic experiments that were earlier used to delineate the energetics of the wild-type isomerase has provided the free energy profile of the mutant enzyme. Comparison of the energetics of the wild-type and mutant enzymes shows that only the transition states flanking the enediol intermediate have been substantially affected. The results suggest either that loop closure and deprotonation are coupled and occur in the same rate-limiting step or that these two processes happen sequentially but interdependently. This finding is consistent with structural information that indicates that the catalytic base glutamate-165 moves 2 A toward the substrate upon loop closure.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
St Maurice M  Bearne SL 《Biochemistry》2002,41(12):4048-4058
Mandelate racemase (EC 5.1.2.2) from Pseudomonas putida catalyzes the interconversion of the two enantiomers of mandelic acid with remarkable proficiency, producing a rate enhancement exceeding 15 orders of magnitude. The rates of the forward and reverse reactions catalyzed by the wild-type enzyme and by a sluggish mutant (N197A) have been studied in the absence and presence of several viscosogenic agents. A partial dependence on relative solvent viscosity was observed for values of kcat and kcat/Km for the wild-type enzyme in sucrose-containing solutions. The value of kcat for the sluggish mutant was unaffected by varying solvent viscosity. However, sucrose did have a slight activating effect on mutant enzyme efficiency. In the presence of the polymeric viscosogens poly(ethylene glycol) and Ficoll, no effect on kcat or kcat/Km for the wild-type enzyme was observed. These results are consistent with both substrate binding and product dissociation being partially rate-determining in both directions. The viscosity variation method was used to estimate the rate constants comprising the steady-state expressions for kcat and kcat/Km. The rate constant for the conversion of bound (R)-mandelate to bound (S)-mandelate (k2) was found to be 889 +/- 40 s(-1) compared with a value of 654 +/- 58 s(-1) for kcat in the same direction. From the temperature dependence of Km (shown to equal K(S)), k2, and the rate constant for the uncatalyzed reaction [Bearne, S. L., and Wolfenden, R. (1997) Biochemistry 36, 1646-1656], we estimated the enthalpic and entropic changes associated with substrate binding (DeltaH = -8.9 +/- 0.8 kcal/mol, TDeltaS = -4.8 +/- 0.8 kcal/mol), the activation barrier for conversion of bound substrate to bound product (DeltaH# = +15.4 +/- 0.4 kcal/mol, TDeltaS# = +2.0 +/- 0.1 kcal/mol), and transition state stabilization (DeltaH(tx) = -22.9 +/- 0.8 kcal/mol, TDeltaS(tx) = +1.8 +/- 0.8 kcal/mol) during mandelate racemase-catalyzed racemization of (R)-mandelate at 25 degrees C. Although the high proficiency of mandelate racemase is achieved principally by enthalpic reduction, there is also a favorable and significant entropic contribution.  相似文献   

3.
J P Richard 《Biochemistry》1991,30(18):4581-4585
Kinetic parameters for triosephosphate isomerase catalysis of the elimination reaction of an equilibrium mixture of dihydroxyacetone phosphate (DHAP) and D-glyceraldehyde-3-phosphate (DGAP) to form methylglyoxal and phosphate ion are reported for the enzyme from rabbit muscle. Pseudo-first-order rate constants for the disappearance of substrate (kelim) were determined for reactions at [Enzyme] much greater than [Substrate]. The second-order rate constant kEnz = 10.1 M-1 s-1 was determined from a plot of kelim against enzyme concentration. The kinetic parameters, determined from a steady-state kinetic analysis at [Substrate] much greater than [Enzyme], are kcat = 0.011 s-1, Km = 0.76 mM, and kcat/Km = 14 M-1 s-1. The estimated rate-constant ratio for partitioning of the enzyme-bound intermediate between protonation at carbon 2 and elimination, 1,000,000, is much larger than the ratio of 6.5 determined for the reaction of the enediolate phosphate in a loose complex with quinuclidinonium cation, a small buffer catalyst. There is a 10(5)-10(8)-fold decrease in the rate constant for the elimination reaction of the enediolate phosphate when this species binds to triosephosphate isomerase. The kinetic parameters for the elimination reaction catalyzed by the native triosephosphate isomerase and for the reaction catalyzed by a mutant form of the enzyme, which is missing a segment that forms hydrogen bonds with the phosphate group of substrate [Pompliano, D. L., Peyman, A., & Knowles, J. R. (1990) Biochemistry 29, 3186-3194] are similar.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
An important active-site residue in the glycolytic enzyme triosephosphate isomerase is His-95, which appears to act as an electrophilic component in catalyzing the enolization of the substrates. With the techniques of site-directed mutagenesis, His-95 has been replaced by Gln in the isomerase from Saccharomyces cerevisiae. The mutant isomerase has been expressed in Escherichia coli strain DF502 and purified to homogeneity. The specific catalytic activity of the mutant enzyme is less than that of wild type by a factor of nearly 400. The mutant enzyme can be resolved from the wild-type isomerase on nondenaturing isoelectric focusing gels, and an isomerase activity stain shows that the observed catalytic activity indeed derives from the mutant protein. The inhibition constants for arsenate and for glycerol phosphate with the mutant enzyme are similar to those with the wild-type isomerase, but the substrate analogues 2-phosphoglycolate and phosphoglycolohydroxamate bind 8- and 35-fold, respectively, more weakly to the mutant isomerase. The mutant enzyme shows the same stereospecificity of proton transfer as the wild type. Tritium exchange experiments similar to those used to define the free energy profile for the wild-type yeast isomerase, together with a new method of analysis involving 14C and 3H doubly labeled substrates, have been used to investigate the energetics of the mutant enzyme catalyzed reaction. When the enzymatic reaction is conducted in tritiated solvent, the mutant isomerase does not catalyze any appreciable exchange between protons of the remaining substrate and those of the solvent either in the forward reaction direction (using dihydroxyacetone phosphate as substrate) or in the reverse direction (using glyceraldehyde phosphate as substrate). However, the specific radioactivity of the product glyceraldehyde phosphate formed in the forward reaction is 31% that of the solvent, while that of the product dihydroxyacetone phosphate formed in the reverse reaction is 24% that of the solvent. The deuterium kinetic isotope effects observed with the mutant isomerase using [1(R)-2H]dihydroxyacetone phosphate and [2-2H]glyceraldehyde 3-phosphate are 2.15 +/- 0.04 and 2.4 +/- 0.1, respectively. These results lead to the conclusion that substitution of Gln for His-95 so impairs the ability of the enzyme to stabilize the reaction intermediate that there is a change in the pathways of proton transfer mediated by the mutant enzyme. The data allow us more closely to define the role of His-95 in the reaction catalyzed by the wild-type enzyme, while forcing us to be alert to subtle changes in mechanistic pathways when mutant enzymes are generated.  相似文献   

5.
The kinetic properties of Trypanosoma brucei brucei triose-phosphate isomerase are compared with those of the commercially available rabbit muscle and yeast enzymes and with published data on the chicken muscle enzyme. With glyceraldehyde 3-phosphate as substrate Km = 0.25 +/- 0.05 mM and kcat = 3.7 X 10(5) min-1. With dihydroxyacetone phosphate as substrate Km = 1.2 +/- 0.1 mM and kcat = 6.5 X 10(4) min-1. The pH dependence of Km and Vmax at 0.1 M ionic strength is in agreement with the results published for the yeast and chicken muscle enzymes. At ionic strength below 0.05 M the effect of a charged group specific for the trypanosomal enzyme and absent from the yeast and rabbit muscle enzymes becomes detectable. This effect significantly increases Km whereas Vmax becomes slightly higher. Trypanosomal triose-phosphate isomerase is inhibited by sulphate, phosphate and arsenate ions, by 2-phosphoglycolate and a number of documented inhibitors in the same concentration range as are the other triose-phosphate isomerases. The trypanocidal drug, Suramin inhibits T. brucei and rabbit muscle triose-phosphate isomerase to the same extent while leaving the yeast enzyme relatively unaffected.  相似文献   

6.
The function of aspartic acid residue 101 in the active site of Escherichia coli alkaline phosphatase was investigated by site-specific mutagenesis. A mutant version of alkaline phosphatase was constructed with alanine in place of aspartic acid at position 101. When kinetic measurements are carried out in the presence of a phosphate acceptor, 1.0 M Tris, pH 8.0, both the kcat and the Km for the mutant enzyme increase by approximately 2-fold, resulting in almost no change in the kcat/Km ratio. Under conditions of no external phosphate acceptor and pH 8.0, both the kcat and the Km for the mutant enzyme decrease by approximately 2-fold, again resulting in almost no change in the kcat/Km ratio. The kcat for the hydrolysis of 4-methyl-umbelliferyl phosphate and p-nitrophenyl phosphate are nearly identical for both the wild-type and mutant enzymes, as is the Ki for inorganic phosphate. The replacement of aspartic acid 101 by alanine does have a significant effect on the activity of the enzyme as a function of pH, especially in the presence of a phosphate acceptor. At pH 9.4 the mutant enzyme exhibits 3-fold higher activity than the wild-type. The mutant enzyme also exhibits a substantial decrease in thermal stability: it is half inactivated by treatment at 49 degrees C for 15 min compared to 71 degrees C for the wild-type enzyme. The data reported here suggest that this amino acid substitution alters the rates of steps after the formation of the phospho-enzyme intermediate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
To allow the detailed interpretation of the pH dependences of the steady-state parameters for the reaction catalyzed by triosephosphate isomerase, three kinds of experiments have been performed. First, the value of kcat/Km for enzyme-catalyzed isomerization of the phosphonate analogue of D-glyceraldehyde 3-phosphate (2-hydroxy-4-phosphonobutyraldehyde) has been shown to titrate with an apparent pKa of 7.5, which is close to the phosphonate's second ionization constant. Secondly, the sulfate ester analogue of dihydroxyacetone phosphate (dihydroxyacetone sulfate), which exists only as a monoanion over the pH range of interest, has been shown not to bind detectably to the enzyme. Thirdly, an isotopic discrimination experiment at pH 5.2 has been compared with a similar investigation at pH 7.6. The results together demonstrate that both enzyme and substrate ionizations control the reaction rate in the pH range 5 to 8.  相似文献   

8.
J Steyaert  L Wyns  P Stanssens 《Biochemistry》1991,30(35):8661-8665
We report on the effect of the viscogenic agents glycerol and ficoll on the RNase T1 catalyzed turnover of GpA, GpC, GpU, and Torula yeast RNA. For wild-type enzyme, we find that the kcat/Km values for the transesterification of GpC and GpA as well as for the cleavage of RNA are inversely proportional to the relative viscosity of glycerol-containing buffers; no such effect is observed for the conversion of GpU to cGMP and U. The second-order rate constants for His40Ala and Glu46Ala RNase T1, two mutants with a drastically reduced kcat/km ratio, are independent of the microviscosity, indicating that glycerol does not affect the intrinsic kinetic parameters. Consistent with the notion that molecular diffusion rates are unaffected by polymeric viscogens, addition of ficoll has no effect on the kcat/Km for GpC transesterification by wild-type enzyme. The data indicate that the second-order rate constants for GpC, GpA, and Torula yeast RNA are at least partly limited by the diffusion-controlled association rate of substrate and active site; RNase T1 obeys Briggs-Haldane kinetics for these substrates (Km greater than Ks). Calculations suggest that the equilibrium dissociation constants (Ks) for the various GpN-wild-type enzyme complexes are virtually independent of N whereas the measured kcat values follow the order GpC greater than GpA greater than GpU. This is also revealed by the steady-state kinetic parameters of Tyr38Phe and His40Ala RNase T1, two mutants that follow simple Michaelis-Menten kinetics because of a dramatically reduced kcat value (i.e., Km = Ks).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
N S Sampson  J R Knowles 《Biochemistry》1992,31(36):8482-8487
To determine what drives the closure of the active-site loop in the reaction catalyzed by triosephosphate isomerase, several residues involved in hydrogen bonding between the loop and the bulk of the protein have been altered. It was known from earlier work that the loop serves two functions: to stabilize the reaction intermediate (and the two transition states that flank it) and to prevent the loss of this unstable species into free solution. To discover what elements of the protein are necessary for proper closure of the loop, selective destabilization of the "open" and the "closed" forms of the enzyme with respect to one another has been attempted. The mutant Y164F isomerase has been prepared to evaluate the importance of the structure of the "open" form, and the mutant E129Q, Y208F, and S211A enzymes have allowed investigation of the "closed" form. The integrity of the loop itself has been destabilized by making the T172A isomerase. We have found that only those mutations that destabilize the "closed" form of the enzyme significantly perturb the catalytic properties of the isomerase. The second-order rate constants (kcat/Km) of the S211A and E129Q enzymes are reduced 30-fold, and that of the mutant Y208F enzyme is reduced 2000-fold, from the level of the wild-type enzyme. The dramatic drop in activity of the Y208F enzyme is accompanied by a 200-fold increase in the dissociation constant of the intermediate analogue phosphoglycolohydroxamate. The most important property of the mobile loop of triosephosphate isomerase lies, therefore, in the stability of the system when the active site contains ligand and the loop is closed.  相似文献   

10.
We have replaced asparagine residues at the subunit interface of yeast triosephosphate isomerase (TIM) using site-directed mutagenesis in order to elucidate the effects of substitutions on the catalytic activity and conformational stability of the enzyme. The mutant proteins were expressed in a strain of Escherichia coli lacking the bacterial isomerase and purified by ion-exchange and immunoadsorption chromatography. Single replacements of Asn-78 by either Thr or Ile residues had little effect on the enzyme's catalytic efficiency, while the single replacement Asn-78----Asp-78 and the double replacement Asn-14/Asn-78----Thr-14/Ile-78 appreciably lowered kcat for the substrate D-glyceraldehyde 3-phosphate. The isoelectric point of the mutant Asn-78----Asp-78 was equivalent to that of wild-type yeast TIM that had undergone a single, heat-induced deamidation, and this mutant enzyme was less resistant than wild-type TIM to denaturation and inactivation caused by elevated temperature, denaturants, tetrabutylammonium bromide, alkaline pH, and proteases.  相似文献   

11.
The function of the mobile loop of triosephosphate isomerase has been investigated by deleting four contiguous residues from the part of this loop that interacts directly with the bound substrate. From the crystal structure of the wild-type enzyme, it appears that this excision will not significantly alter the conformation of the rest of the main chain of the protein. The specific catalytic activity of the purified mutant enzyme is nearly 10(5)-fold lower than that of the wild type. Kinetic measurements and isotopic partitioning studies show that the decrease in activity is due to much higher activation barriers for the enolization of enzyme-bound substrate. Although the substrates bind somewhat more weakly to the mutant enzyme than to the wild type, the intermediate analogue phosphoglycolohydroxamate binds much less well (by 200-fold) to the mutant. It seems that the deleted residues of the loop contribute critically to the stabilization of the enediol phosphate intermediate. Consistent with this view, the mutant enzyme can no longer prevent the loss of the enediol phosphate from the active site and its rapid decomposition to methylglyoxal and inorganic phosphate. Indeed, when glyceraldehyde 3-phosphate is the substrate, the enediol phosphate intermediate is lost (and decomposes) 5.5 times faster than it reprotonates to form the product dihydroxyacetone phosphate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Electrophilic catalysis by histidine-95 in triosephosphate isomerase has been probed by using Fourier transform infrared spectroscopy and X-ray crystallography. The carbonyl stretching frequency of dihydroxyacetone phosphate bound to the wild-type enzyme is known to be 19 cm-1 lower (at 1713 cm-1) than that of dihydroxyacetone phosphate free in solution (at 1732 cm-1), and this decrease in stretching frequency has been ascribed to an enzymic electrophile that polarizes the substrate carbonyl group toward the transition state for the enolization. Infrared spectra of substrate bound to two site-directed mutants of yeast triosephosphate isomerase in which histidine-95 has been changed to glutamine or to asparagine show unperturbed carbonyl stretching frequencies between 1732 and 1742 cm-1. The lack of carbonyl polarization when histidine-95 is removed suggests that histidine-95 is indeed the catalytic electrophile, at least for dihydroxyacetone phosphate. Kinetic studies of the glutamine mutant (H95Q) have shown that the enzyme follows a subtly different mechanism of proton transfers involving only a single acid-base catalytic group. These findings suggest an additional role for histidine-95 as a general acid-base catalyst in the wild-type enzyme. The X-ray crystal structure of the H95Q mutant with an intermediate analogue, phosphoglycolohydroxamate, bound at the active site has been solved to 2.8-A resolution, and this structure clearly implicates glutamate-165, the catalytic base in the wild-type isomerase, as the sole acid-base catalyst for the mutant enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
S C Blacklow  K D Liu  J R Knowles 《Biochemistry》1991,30(34):8470-8476
Second-site suppressor changes that improve the catalytic potency of a sluggish mutant of the enzyme triosephosphate isomerase have been examined both individually and in combination. Each of the second-site mutations increases the specific catalytic activity of a triosephosphate isomerase in which the catalytic base, glutamate-165, has been changed to aspartate. These second-site suppressors are G10S, S96P, S96T, E97D, V167D, and G233R. Not one of these changes enhances the value of kcat/Km for the wild-type enzyme, which is consistent with the knowledge that the reaction catalyzed by the wild-type enzyme is already diffusion-controlled. Indeed, two of the changes, S96P and V167D, are catalytically deleterious to the wild-type isomerase. When pairs of second-site suppressors are combined with the primary lesion E165D, six pairs show additive independence while the effects of eight other pairs are less than additive. The sites fall into two clusters: pairs within a cluster always interfere with one another and do not produce additive improvements in catalytic activity, whereas combinations of changes from different clusters tend to be additive in their effects. No combination of second-site suppressor mutations behaves synergistically, though there seems to be no a priori reason to exclude this possibility. Since the catalytic potency of each of the six second-site suppressor mutants can be further improved by the introduction of (at least) one of the other five changes, it is evident that none of the double mutants lies at a local catalytic maximum. In these cases, therefore, the opportunity exists for at least two "steps" of monotonic catalytic improvement along each of six different "paths" in protein space.  相似文献   

14.
The function of arginine residue 166 in the active site of Escherichia coli alkaline phosphatase was investigated by site-directed mutagenesis. Two mutant versions of alkaline phosphatase, with either serine or alanine in the place of arginine at position 166, were generated by using a specially constructed M13 phage carrying the wild-type phoA gene. The mutant enzymes with serine and alanine at position 166 have very similar kinetic properties. Under conditions of no external phosphate acceptor, the kcat for the mutant enzymes decreases by approximately 30-fold while the Km increases by less than 2-fold. When kinetic measurements are carried out in the presence of a phosphate acceptor, 1.0 M Tris, the kcat for the mutant enzymes is reduced by less than 3-fold, while the Km increases by more than 50-fold. For both mutant enzymes, in either the absence or the presence of a phosphate acceptor, the catalytic efficiency as measured by the kcat/Km ratio decreases by approximately 50-fold as compared to the wild type. Measurements of the Ki for inorganic phosphate show an increase of approximately 50-fold for both mutants. Phenylglyoxal, which inactivates the wild-type enzyme, does not inactivate the Arg-166----Ala enzyme. This result indicates that Arg-166 is the same arginine residue that when chemically modified causes loss of activity [Daemen, F.J.M., & Riordan, J.F. (1974) Biochemistry 13, 2865-2871]. The data reported here suggest that although Arg-166 is important for activity is not essential. The analysis of the kinetic data also suggests that the loss of arginine-166 at the active site of alkaline phosphatase has two different effects on the enzyme. First, the binding of the substrate, and phosphate as a competitive inhibitor, is reduced; second, the rate of hydrolysis of the covalent phosphoenzyme may be diminished.  相似文献   

15.
The essential catalytic base at the active site of the glycolytic enzyme triosephosphate isomerase is the carboxylate group of Glu-165, which directly abstracts either the 1-pro-R proton of dihydroxyacetone phosphate or the 2-proton of (R)-glyceraldehyde 3-phosphate to yield the cis-enediol intermediate. Using the methods of site-directed mutagenesis, we have replaced Glu-165 by Asp. The three enzymes chicken isomerase from chicken muscle, wild-type chicken isomerase expressed in Escherichia coli, and mutant (Glu-165 to Asp) chicken isomerase expressed in E. coli have each been purified to homogeneity. The specific catalytic activities of the two wild-type isomerases are identical, while the specific activity of the mutant enzyme is reduced by a factor of about 1000. The observed kinetic differences do not derive from a change in mechanism in which the aspartate of the mutant enzyme acts as a general base through an intervening water molecule, because the D2O solvent isotope effects and the stoichiometries of inactivation with bromohydroxyacetone phosphate are identical for the wild-type and mutant enzymes. Using the range of isotopic experiments that were used to delineate the free-energy profile of the wild-type chicken enzyme, we here derive the complete energetics of the reaction catalyzed by the mutant protein. Comparison of the reaction energetics for the wild-type and mutant isomerases shows that only the free energies of the transition states for the two enolization steps have been seriously affected. Each of the proton abstraction steps is about 1000-fold slower in the mutant enzyme. Evidently, the excision of a methylene group from the side chain of the essential glutamate has little effect on the free energies of the intermediate states but dramatically reduces the stabilities of the transition states for the chemical steps in the catalyzed reaction.  相似文献   

16.
Mature boar spermatozoa oxidized glycerol to carbon dioxide in the absence of any detectable activity of glycerol kinase. With triosephosphate isomerase and glyceraldehyde-3-phosphate dehydrogenase inhibited by the presence of 3-chloro-1-hydroxypropanone (CHOP), dihydroxyacetone phosphate accumulated in incubates when glycerol-3-phosphate was the substrate, but not when it was glycerol. Both dihydroxyacetone and glyceraldehyde could be used as substrates; in the presence of CHOP, dihydroxyacetone phosphate and fructose-1,6-bisphosphate accumulated when dihydroxyacetone was the substrate, but not when it was glyceraldehyde. The metabolic pathways glycerol----glyceraldehyde----glyceraldehyde 3-phosphate and dihydroxyacetone----dihydroxyacetone phosphate have been shown to operate in these cells.  相似文献   

17.
Tanaka K  Suzuki T 《FEBS letters》2004,573(1-3):78-82
The purpose of this study is to elucidate the mechanisms of guanidine substrate specificity in phosphagen kinases, including creatine kinase (CK), glycocyamine kinase (GK), lombricine kinase (LK), taurocyamine kinase (TK) and arginine kinase (AK). Among these enzymes, LK is unique in that it shows considerable enzyme activity for taurocyamine in addition to its original target substrate, lombricine. We earlier proposed several candidate amino acids associated with guanidine substrate recognition. Here, we focus on amino-acid residue 95, which is strictly conserved in phosphagen kinases: Arg in CK, Ile in GK, Lys in LK and Tyr in AK. This residue is not directly associated with substrate binding in CK and AK crystal structures, but it is located close to the binding site of the guanidine substrate. We replaced amino acid 95 Lys in LK isolated from earthworm Eisenia foetida with two amino acids, Arg or Tyr, expressed the modified enzymes in Escherichia coli as a fusion protein with maltose-binding protein, and determined the kinetic parameters. The K95R mutant enzyme showed a stronger affinity for both lombricine (Km=0.74 mM and kcat/Km=19.34 s(-1) mM(-1)) and taurocyamine (Km=2.67 and kcat/Km=2.81), compared with those of the wild-type enzyme (Km=5.33 and kcat/Km=3.37 for lombricine, and Km=15.31 and kcat/ Km=0.48for taurocyamine). Enzyme activity of the other mutant, K95Y, was dramatically altered. The affinity for taurocyamine (Km=1.93 and kcat/Km=6.41) was enhanced remarkably and that for lombricine (Km=14.2 and kcat/Km=0.72) was largely decreased, indicating that this mutant functions as a taurocyamine kinase. This mutant also had a lower but significant enzyme activity for the substrate arginine (Km=33.28 and kcat/Km=0.01). These results suggest that Eisenia LK is an inherently flexible enzyme and that substrate specificity is strongly controlled by the amino-acid residue at position 95.  相似文献   

18.
Three mutant Escherichia coli aspartate aminotransferases in which Val39 was changed to Ala, Leu, and Phe by site-directed mutagenesis were prepared and characterized. Among the three mutant and the wild-type enzymes, the Leu39 enzyme had the lowest Km values for dicarboxylic substrates. The Km values of the Ala39 enzyme for dicarboxylates were essentially the same as those of the wild-type (Val39) enzyme. These two mutant enzymes showed essentially the same kcat values for dicarboxylic substrates as did the wild-type enzyme. On the other hand, incorporation of a bulky side-chain at position 39 (Phe39 enzyme) decreased both the affinity (1/Km) and catalytic ability (kcat) toward dicarboxylic substrates. These results show that the position 39 residue is involved in the modulation of both the binding of dicarboxylic substrates to enzyme and the catalytic ability of the enzyme. Although the replacement of Val39 with other residues altered both the kcat and Km values toward various substrates including dicarboxylic and aromatic amino acids and the corresponding oxo acids, it did not alter the ratio of the kcat/Km value of the enzyme toward a dicarboxylic substrate to that for an aromatic substrate. The affinity for aromatic substrates was not affected by changing the residue at position 39. These data indicate that, although the side chain bulkiness of the residue at position 39 correlates well with the activity toward aromatic substrates in the sequence alignment of several aminotransferases [Seville, M., Vincent M.G., & Hahn, K. (1988) Biochemistry 27, 8344-8349], the residue does not seem to be involved in the recognition of aromatic substrates.  相似文献   

19.
Mature epididymal boar spermatozoa converted fructose and glycerol to carbon dioxide but in the presence of 3-chloro-1-hydroxyacetone these oxidations were inhibited. When the substrate was fructose, there was an increase in the amounts of fructose 1,6-bisphosphate and dihydroxyacetone phosphate but these glycolytic intermediates did not accumulate when glycerol was the substrate. Examination of enzyme activities in mature boar spermatozoa incubated with 3-chloro-1-hydroxyacetone, which is metabolised in vitro to (S)-3-chlorolactaldehyde, confirmed that glyceraldehyde 3-phosphate dehydrogenase and triosephosphate isomerase were both inhibited to equivalent degrees by this metabolite.  相似文献   

20.
Using site-directed mutagenesis, Arginine-171 at the substrate-binding site of Bacillus stearothermophilus, lactate dehydrogenase has been replaced by lysine. In the closely homologous eukaryotic lactate dehydrogenase, this residue binds the carboxylate group of the substrate by forming a planar bifurcated bond. The mutation diminishes the binding energy of pyruvate, alpha-ketobutyrate and alpha-ketovalerate (measured by kcat/Km) by the same amount (about 6 kcal/mol). For each additional methylene group on the substrate, there is a loss of about 1.5 kcal/mol of binding energy in both mutant and wild-type enzymes. From these parallel trends in the two forms of enzyme, we infer that the mode of productive substrate binding is identical in each, the only difference being the loss of a strong carboxylate-guanidinium interaction in the mutant. In contrast to this simple pattern in kcat/Km, the Km alone increases with substrate-size in the wild-type enzyme, but decreases in the mutant. These results can be most simply explained by the occurrence of relatively tight unproductive enzyme-substrate complexes in the mutant enzyme as the substrate alkyl chain is extended. This does not occur in the wild-type enzyme, because the strong orienting effect of Arg-171 maximizes the frequency of substrates binding in the correct alignment.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号