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 15 毫秒
1.
S C Blacklow  J R Knowles 《Biochemistry》1990,29(17):4099-4108
The reaction energetics of four triosephosphate isomerase mutants are compared with those of the wild-type enzyme. The two primary mutants, E165D and H95N, contain site-specific alterations of active site residues. In one case the active site base has been altered (E165D), and in the other, an active site electrophile has been removed (H95N), yet the major effect in each case is the relative destabilization of the transition states for the two chemical (enolization) steps that constitute the catalytic reaction. When the genes encoding each of these sluggish mutant isomerases were subjected to random mutagenesis using chemical reagents and a selection for isomerases of increased catalytic potency was performed, pseudorevertant enzymes with dramatic increases in activity were found. Remarkably, the same second-site suppressor locus partially corrects each lesion. The E165D,S96P pseudorevertant is a 20-fold better catalyst than the E165D mutant from which it is derived, and the H95N,S96P pseudorevertant is about 60 times more active than its H95N parent. The S96P substitution thus increases the catalytic activity in each of two different contexts, H95N and E165D. The energetic consequences of the S96P change are suprisingly similar in each pseudorevertant. The H95N,S96P enzyme is more effective than H95N at stabilizing the intermediate enediol(ate) phosphate and its flanking transition states. The E165D,S96P enzyme likewise stabilizes the transition states for enolization better than E165D, and this pseudorevertant also forms a tighter enzyme-dihydroxyacetone phosphate complex than its parent. These data show how, in these two cases, the catalytic potency of sluggish mutant enzymes can be improved by second-site changes. The results thus provide the beginnings of a detailed understanding of the kinetic refinement of enzyme catalysts.  相似文献   

2.
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 With the techniques of site-directed mutagenesis, His-95 has been replaced by Gln in the isomerase from Saccharomyces cerevisiae.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   

3.
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   

4.
The kinetic characteristics of E. coli aspartate transcarbamylase, altered by site-specific mutagenesis of Glu-239----Gln, have been determined by equilibrium isotope-exchange kinetics and compared to the wild-type system. In wild-type enzyme, residue Glu-239 helps to stabilize the T-state structure by multiple bonding interactions with Tyr-165 and Lys-164 across the c1-c4 subunit interface; upon conversion to the R-state, these bonds are re-formed within c-chains. Catalysis of both the [14C]Asp in equilibrium C-Asp and [32P]ATP in equilibrium Pi exchanges by mutant enzyme occurs at rates comparable to those for wild-type enzyme. Saturation with different reactant/product pairs produced kinetic patterns consistent with strongly preferred order binding of carbamyl-P prior to Asp and carbamyl-Asp release before Pi. The kinetics for the Gln-239 mutant enzyme resemble those observed for catalytic subunits (c3), namely a R-state enzyme (Hill coefficient nH = 1.0) and Km (Asp) approximately equal to 6 mM. The Glu-239----Gln mutation appears to destablize both the T- and R-states, whereas the Tyr-240----Phe mutation destablizes only the T-state.  相似文献   

5.
We present an ab-initio self-consistent field calculation with a 4-31G basis set on a simple model for proton abstraction from hydroxyacetone (a model for dihydroxyacetone phosphate; DHAP) by formate, which is a model for Glu165 in triose phosphate isomerase. Earlier, we showed that the electrophilic groups on the enzyme (the NH3+ of Lys13 and the NH of His95) were essential to efficient catalysis by triose phosphate isomerase. These groups stabilized the enediolate formed by proton abstraction from the DHAP model so that proton transfer from this molecule to Glu165 became likely. In this study, we carry this analysis one step further. First, we re-examine the energy profile for proton transfer, using the fact that our earlier calculations showed that the combined effect of His95 and Lys13 on the reactant DHAP and intermediate enediolate was to make them equal in energy. Then, we analyze the likely effect of changing Glu165 to Asp165 and relate this to experiments on the kinetics of enzyme catalysis by the Glu165----Asp165 mutant.  相似文献   

6.
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   

7.
Two site-specifically mutated human dihydrolipoamide dehydrogenases (His-452----Gln and Glu-457----Gln) were expressed in pyruvate dehydrogenase complex-deletion mutant Escherichia coli JRG1342. The expressed mutant E3s were purified to near homogeneity using DEAE-Sephacel and hydroxyapatite columns. The initial velocity measurements in the absence of products for the Gln-452 mutant E3 in the direction of NAD+ reduction showed parallel lines in double-reciprocal plots, indicating that the mutant E3, like wild-type enzyme, catalyzed E3 reaction via a ping-pong mechanism. The specific activity of the Gln-452 mutant E3 was about 0.2% of that of wild-type enzyme. Its Km for dihydrolipoamide was dramatically increased by 63-fold. The substitution of His-452 to Gln resulted in a destabilization of the transition state of human E3 catalysis by about 6.4 kcal mol-1. The Gln-457 mutant E3, unlike wild-type enzyme, catalyzed E3 reaction via a sequential mechanism in the direction of NAD+ reduction based on the intersecting lines shown on double-reciprocal plots. Its specific activity decreased to 28% of that of wild-type enzyme. Its Km for dihydrolipoamide increased about 4.3-fold. The substitution of Glu-457 to Gln resulted in a destabilization of the transition state by about 1.7 kcal mol-1. These results indicate that His-452, which is a possible proton acceptor/donor in human E3 reaction, is critical to human E3 catalysis and that the local environment around His-452 and Glu-457, which are suggested to be hydrogen-bonded, is important in the binding of dihydrolipoamide to the enzyme.  相似文献   

8.
The gene coding for thermophilic xylose (glucose) isomerase of Clostridium thermosulfurogenes was isolated and its complete nucleotide sequence was determined. The structural gene (xylA) for xylose isomerase encodes a polypeptide of 439 amino acids with an estimated molecular weight of 50,474. The deduced amino acid sequence of thermophilic C. thermosulfurogenes xylose isomerase displayed higher homology with those of thermolabile xylose isomerases from Bacillus subtilis (70%) and Escherichia coli (50%) than with those of thermostable xylose isomerases from Ampullariella (22%), Arthrobacter (23%), and Streptomyces violaceoniger (24%). Several discrete regions were highly conserved throughout the amino acid sequences of all these enzymes. To identify the histidine residue of the active site and to elucidate its function during enzymatic xylose or glucose isomerization, histidine residues at four different positions in the C. thermosulfurogenes enzyme were individually modified by site-directed mutagenesis. Substitution of His101 by phenylalanine completely abolished enzyme activity whereas substitution of other histidine residues by phenylalanine had no effect on enzyme activity. When His101 was changed to glutamine, glutamic acid, asparagine, or aspartic acid, approximately 10-16% of wild-type enzyme activity was retained by the mutant enzymes. The Gln101 mutant enzyme was resistant to diethylpyrocarbonate inhibition which completely inactivated the wild-type enzyme, indicating that His101 is the only essential histidine residue involved directly in enzyme catalysis. The constant Vmax values of the Gln101, Glu101, Asn101, and Asp101 mutant enzymes over the pH range of 5.0-8.5 indicate that protonation of His101 is responsible for the reduced Vmax values of the wild-type enzyme at pH below 6.5. Deuterium isotope effects by D-[2-2H]glucose on the rate of glucose isomerization indicated that hydrogen transfer and not substrate ring opening is the rate-determining step for both the wild-type and Gln101 mutant enzymes. These results suggest that the enzymatic sugar isomerization does not involve a histidine-catalyzed proton transfer mechanism. Rather, essential histidine functions to stabilize the transition state by hydrogen bonding to the C5 hydroxyl group of the substrate and this enables a metal-catalyzed hydride shift from C2 to C1.  相似文献   

9.
A mutant (D165N) of clostridial glutamate dehydrogenase (GDH) in which the catalytic Asp is replaced by Asn surprisingly showed a residual 2% of wild-type activity when purified after expression in Escherichia coli at 37 degrees C. This low-level activity also displayed Michaelis constants for substrates that were remarkably similar to those of the wild-type enzyme. Expression at 8 degrees C gave a mutant enzyme preparation 1000 times less active than the first preparation, but progressively, over 2 weeks' incubation at 37 degrees C in sealed vials, this enzyme regained 90% of the specific activity of wild type. This suggested that the mutant might undergo spontaneous deamidation. Mass spectrometric analysis of tryptic peptides derived from D165N samples treated in various ways showed (i) that the Asn is in place in D165N GDH freshly prepared at 8 degrees C; (ii) that there is a time-dependent reversion of this Asn to Asp over the 2-week incubation period; (iii) that detectable deamidation of other Asn residues, in Asn-Gly sequences, mainly occurred in sample workup rather than during the 2-week incubation; (iv) that there is no significant deamidation of other randomly chosen Asn residues in this mutant over the same period; and (v) that when the protein is denatured before incubation, no deamidation at Asn-165 is detectable. It appears that this deamidation depends on the residual catalytic machinery of the mutated GDH active site. A literature search indicates that this finding is not unique and that Asn may not be a suitable mutational replacement in the assessment of putative catalytic Asp residues by site-directed mutagenesis.  相似文献   

10.
Chloroacetol phosphate covalently reacts with Glu-165 in the catalytic center of triosephosphate isomerase. Reaction of the enzyme with the substrate analogue results in two 31P resonances at 6.8 and 5.5 ppm. Dissociation with guanidinium chloride results in a single resonance at 4.5 ppm. Reassociation and redimerization of the triosephosphate isomerase-chloroacetol phosphate complex restores only the resonance at 5.5 ppm. The two 31P resonances appear to represent different conformations of the enzyme which are trapped upon reaction with the affinity label.  相似文献   

11.
The reaction mechanism of the catalytic phosphoryl transfer of cAMP-dependent protein kinase (cAPK) was investigated by semi-empirical AM1 molecular orbital computations of an active site model system derived from the crystal structure of the catalytic subunit of the enzyme. The activation barrier is calculated as 20.7 kcal mol(-1) and the reaction itself to be exothermic by 12.2 kcal mol(-1). The active site residue Asp166, which was often proposed to act as a catalytic base, does not accept a proton in any of the reaction steps. Instead, the hydroxyl hydrogen of serine is shifted to the simultaneously transferred phosphate group of ATP. Although the calculated transition state geometry indicates an associative phosphoryl transfer, no concentration of negative charge is found. To study the influence of protein mutations on the reaction mechanism, we compared two-dimensional energy hypersurfaces of the protein kinase wild-type model and a corresponding mutant in which Asp166 was replaced by alanine. Surprisingly, they show similar energy profiles despite the experimentally known decrease of catalytic activity for corresponding mutants. Furthermore, a model structure was examined, where the charged NH3 group of Lys168 was replaced by a neutral methyl group. The energetic hypersurface of this hypothetical mutant shows two possible pathways for phosphoryl transfer, which both require significantly higher activation energies than the other systems investigated, while the energetic stabilization of the reaction product is similar in all systems. As the position of the amino acid side chains and the substrate peptide is virtually unchanged in all model systems, our results suggest that the exchange of Asp166 by other amino acid is less important to the phosphoryl transfer itself, but crucial to maintain the configuration of the active site in vivo. The positively charged side chain of Lys168, however, is necessary to stabilize the intermediate reaction states, particularly the side chain of the substrate peptide.  相似文献   

12.
Delta(3,5),Delta(2,4)-Dienoyl-CoA isomerase (DI), an auxiliary enzyme of unsaturated fatty acid beta-oxidation, was purified from rat mitochondria and peroxisomes and subjected to N-terminal sequencing to facilitate a mechanistic study of this enzyme. The mature mitochondrial DI from rat heart was lacking its 34 N-terminal amino acid residues that have the properties of a mitochondrial targeting sequence. The peroxisomal isomerase was identified as a product of the same gene with a truncated and ragged N terminus. Expression of the cDNA coding for the mature mitochondrial DI in Escherichia coli yielded an enzyme preparation that was as active as the native DI. Because the recombinant DI also exhibited Delta(3,5,7),Delta(2,4,6)-trienoyl-CoA isomerase (TI) activity, both isomerases reside on the same protein. Mutations of any of the 3 acidic amino acid residues located at the active site (Modis, Y., Filppula, S. A., Novikov, D. K., Norledge, B., Hiltunen, J. K., and Wierenga, R. K. (1998) Structure 6, 957-970) caused activity losses. In contrast to only a 10-fold decrease in activity upon replacement of Asp(176) by Ala, substitutions of Asp(204) by Asn and of Glu(196) by Gln resulted in 10(5)-fold lower activities. Such activity losses are consistent with the direct involvement of these latter two residues in the proposed proton transfers at carbons 2 and 6 or 8 of the substrates. Probing of the wild-type and mutants forms of the enzyme with 2,5-octadienoyl-CoA as substrate revealed low Delta(2),Delta(3)-enoyl-CoA isomerase and Delta(5),Delta(4)-enoyl-CoA isomerase activities catalyzed by Glu(196) and Asp(204), respectively. Altogether, these data reveal that positional isomerizations of the diene and triene are facilitated by simultaneous proton transfers involving Glu(196) and Asp(204), whereas each residue alone can catalyze, albeit less efficiently, a monoene isomerization.  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.
Isopentenyl diphosphate (IPP) isomerase catalyzes an essential activation step in the isoprenoid biosynthetic pathway. A database search based on probes from the highly conserved regions in three eukaryotic IPP isomerases revealed substantial similarity with ORF176 in the photosynthesis gene cluster in Rhodobacter capsulatus. The open reading frame was cloned into an Escherichia coli expression vector. The encoded 20-kDa protein, which was purified in two steps by ion exchange and hydrophobic interaction chromatography, catalyzed the interconversion of IPP and dimethylallyl diphosphate. Thus, the photosynthesis gene cluster encodes all of the enzymes required to incorporate IPP into the ultimate carotenoid and bacteriochlorophyll metabolites in R. capsulatus. More recent searches uncovered additional putative open reading frames for IPP isomerase in seed-bearing plants (Oryza sativa, Arabadopsis thaliana, and Clarkia breweri), a worm (Caenorhabiditis elegans), and another eubacterium (Escherichia coli). The R. capsulatus enzyme is the smallest of the IPP isomerases to be identified thus far and may consist mostly of a fundamental catalytic core for the enzyme.  相似文献   

15.
S G Miran  S H Chang  F M Raushel 《Biochemistry》1991,30(32):7901-7907
Carbamoyl phosphate synthetase from Escherichia coli catalyzes the formation of carbamoyl phosphate from ATP, bicarbonate, and glutamine. The amidotransferase activity of this enzyme is catalyzed by the smaller of the two subunits of the heterodimeric protein. The roles of four conserved histidine residues within this subunit were probed by site-directed mutagenesis to asparagine. The catalytic activities of the H272N and H341N mutants are not significantly different than that of the wild-type enzyme. The H353N mutant is unable to utilize glutamine as a nitrogen source in the synthetase reaction or the partial glutaminase reaction. However, binding to the glutamine active site is not impaired in the H353N enzyme since glutamine is found to activate the partial ATPase reaction by 40% with a Kd of 54 microM. The H312N mutant has a Michaelis constant for glutamine that is 2 orders of magnitude larger than the wild-type value, but the maximal rate of glutamine hydrolysis is unchanged. These results are consistent with His-353 functioning as a general acid/base catalyst for proton transfers while His-312 serves a critical role for the binding of glutamine to the active site.  相似文献   

16.
P J Lodi  J R Knowles 《Biochemistry》1991,30(28):6948-6956
To illuminate the role of histidine-95 in the catalytic reaction mediated by triosephosphate isomerase, 13C and 15N NMR titration studies have been carried out both on the wild-type enzyme and on a mutant isomerase in which the single remaining histidine (that at the active site) has been isotopically enriched in the imidazole ring. 15N NMR has proved especially useful in the unambiguous demonstration that the imidazole ring of histidine-95 is uncharged over the entire pH range of isomerase activity, between pH 5 and pH 9.9. The results require that the first pKa of histidine-95 is below 4.5. This abnormally low pKa rules out the traditional view that the positively charged imidazolium cation of histidine-95 donates a proton to the developing charge on the substrate's carbonyl oxygen. 15N NMR experiments on the enzyme in the presence of the reaction intermediate analogue phosphoglycolohydroxamate show the presence of a strong hydrogen bond between N epsilon 2 of histidine-95 and the bound inhibitor. These findings indicate that, in the catalyzed reaction, proton abstraction from C-1 of dihydroxyacetone phosphate first yields an enediolate intermediate that is strongly hydrogen bonded to the neutral imidazole side chain of histidine-95. The imidazole proton involved in this hydrogen bond then protonates the enediolate, with the transient formation of the enediol-imidazolate ion pair. Abstraction of the hydroxyl proton on O-1 now produces the other enediolate intermediate, which collapses to give the product glyceraldehyde 3-phosphate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Glu-50 of aspartate transcarbamoylase from Escherichia coli forms a set of interdomain bridging interactions between the 2 domains of the catalytic chain; these interactions are critical for stabilization of the high-activity high-affinity form of the enzyme. The mutant enzyme with an alanine substituted for Glu-50 (Glu-50-->Ala) exhibits significantly reduced activity, little cooperativity, and altered regulatory behavior (Newton CJ, Kantrowitz ER, 1990, Biochemistry 29:1444-1451). A study of the structural consequences of replacing Glu-50 by alanine using solution X-ray scattering is reported here. Correspondingly, in the absence of substrates, the mutant enzyme is in the same, so-called T quaternary conformation as is the wild-type enzyme. In the presence of a saturating concentration of the bisubstrate analog N-phosphonacetyl-L-aspartate (PALA), the mutant enzyme is in the same, so-called R quaternary conformation as the wild-type enzyme. However, the Glu-50-->Ala enzyme differs from the wild-type enzyme, in that its scattering pattern is hardly altered by a combination of carbamoyl phosphate and succinate. Addition of ATP under these conditions does result in a slight shift toward the R structure. Steady-state kinetic studies indicate that, in contrast to the wild-type enzyme, the Glu-50-->Ala enzyme is activated by PALA at saturating concentrations of carbamoyl phosphate and aspartate, and that PALA increases the affinity of the mutant enzyme for aspartate. These data suggest that the enzyme does not undergo the normal T to R transition upon binding of the physiological substrates and verifies the previous suggestion that the interdomain bridging interactions involving Glu-50 are critical for the creation of the high-activity, high-affinity R state of the enzyme.  相似文献   

18.
Aspartate transcarbamylase (EC 2.1.3.2) from E. coli is a multimeric enzyme consisting of two catalytic subunits and three regulatory subunits whose activity is regulated by subunit interactions. Differential scanning calorimetric (DSC) scans of the wild-type enzyme consist of two peaks, each comprised of at least two components, corresponding to denaturation of the catalytic and regulatory subunits within the intact holoenzyme (Vickers et al., J. Biol. Chem. 253 (1978) 8493; Edge et al., Biochemistry 27 (1988) 8081). We have examined the effects of nine single-site mutations in the catalytic chains. Three of the mutations (Asp-100-Gly, Glu-86-Gln, and Arg-269-Gly) are at sites at the C1: C2 interface between c chains within the catalytic subunit. These mutations disrupt salt linkages present in both the T and R states of the molecule (Honzatko et al., J. Mol. Biol. 160 (1982) 219; Krause et al., J. Mol. Biol. 193 (1987) 527). The remainder (Lys-164-Ile, Tyr-165-Phe, Glu-239-Gln, Glu-239-Ala, Tyr-240-Phe and Asp-271-Ser) are at the C1: C4 interface between catalytic subunits and are involved in interactions which stabilize either the T or R state. DSC scans of all of the mutants except Asp-100-Gly and Arg-269-Gly consisted of two peaks. At intermediate concentrations, Asp-100-Gly and Arg-269-Gly had only a single peak near the Tm of the regulatory subunit transition in the holoenzyme, although their denaturational profiles were more complex at high and low protein concentrations. The catalytic subunits of Glu-86-Gln, Lys-164-Ile and Asp-271-Ser appear to be significantly destabilized relative to wild-type protein while Tyr-165-Phe and Tyr-240-Phe appear to be stabilized. Values of delta delta G degree cr, the difference between the subunit interaction energy of wild-type and mutant proteins, evaluated as suggested by Brandts et al. (Biochemistry 28 (1989) 8588) range from -3.7 kcal mol-1 for Glu-86-Gln to 2.4 kcal mol-1 for Tyr-165-Phe.  相似文献   

19.
D J Weber  A K Meeker  A S Mildvan 《Biochemistry》1991,30(25):6103-6114
The mechanism of the phosphodiesterase reaction catalyzed by staphylococcal nuclease is believed to involve concerted general acid-base catalysis by Arg-87 and Glu-43. The mutual interactions of Arg-87 and Glu-43 were investigated by comparing kinetic and thermodynamic properties of the single mutant enzymes E43S (Glu-43 to Ser) and R87G (Arg-87 to Gly) with those of the double mutant, E43S + R87G, in which both the basic and acidic functions have been inactivated. Denaturation studies with guanidinium chloride, CD, and 600-MHz 1D and 2D proton NMR spectra, indicate all enzyme forms to be predominantly folded in absence of the denaturant and reveal small antagonistic effects of the E43S and R87G mutations on the stability and structure of the wild-type enzyme. The free energies of binding of the divalent cation activator Ca2+, the inhibitor Mn2+, and the substrate analogue 3',5'-pdTp show simple additive effects of the two mutations in the double mutant, indicating that Arg-87 and Glu-43 act independently to facilitate the binding of divalent cations and of 3',5'-pdTP by the wild-type enzyme. The free energies of binding of the substrate, 5'-pdTdA, both in binary E-S and in active ternary E-Ca(2+)-S complexes, show synergistic effects of the two mutations, suggesting that Arg-87 and Glu-43 interact anticooperatively in binding the substrate, possibly straining the substrate by 1.6 kcal/mol in the wild-type enzyme. The large free energy barriers to Vmax introduced by the R87G mutation (delta G1 = 6.5 kcal/mol) and by the E43S mutation (delta G2 = 5.0 kcal/mol) are partially additive in the double mutant (delta G1+2 = 8.1 kcal/mol). These partially additive effects on Vmax are most simply explained by a cooperative component to transition state binding by Arg-87 and Glu-43 of -3.4 kcal/mol. The combination of anticooperative, cooperative, and noncooperative effects of Arg-87 and Glu-43 together lower the kinetic barrier to catalysis by 8.1 kcal/mol.  相似文献   

20.
Aspartate transcarbamoylase from Escherichia coli shows homotropic cooperativity for aspartate as well as heterotropic regulation by nucleotides. Structurally, it consists of two trimeric catalytic subunits and three dimeric regulatory subunits, each chain being comprised of two domains. Glu-50 and Ser-171 are involved in stabilizing the closed conformation of the catalytic chain. Replacement of Glu-50 or Ser-171 by Ala in the holoenzyme has been shown previously to result in marked decreases in the maximal observed specific activity, homotropic cooperativity, and affinity for aspartate (Dembowski NJ, Newton CJ, Kantrowitz ER, 1990, Biochemistry 29:3716-3723; Newton CJ, Kantrowitz ER, 1990, Biochemistry 29:1444-1451). We have constructed a double mutant enzyme combining both mutations. The resulting Glu-50/ser-171-->Ala enzyme is 9-fold less active than the Ser-171-->Ala enzyme, 69-fold less active than the Glu-50-->Ala enzyme, and shows 1.3-fold and 1.6-fold increases in the [S]0.5Asp as compared to the Ser-171-->Ala and Glu-50-->Ala enzymes, respectively. However, the double mutant enzyme exhibits some enhancement of homotropic cooperativity with respect to aspartate, relative to the single mutant enzymes. At subsaturating concentrations of aspartate, the Glu-50/Ser-171 -->Ala enzyme is activated less by ATP than either the Glu-50-->Ala or Ser-171-->Ala enzyme, whereas CTP inhibition is intermediate between that of the two single mutants. As opposed to the wild-type enzyme, the Glu-50/Ser-171 -->Ala enzyme is activated by ATP and inhibited by CTP at saturating concentrations of aspartate. Structural analysis of the Ser-171-->Ala and Glu-50/Ser-171-->Ala enzymes by solution X-ray scattering indicates that both mutants exist in the same T quaternary structure as the wild-type enzyme in the absence of ligands, and in the same R quaternary structure in the presence of saturating N-(phosphonoacetyl)-L-aspartate. However, saturating concentrations of carbamoyl phosphate and succinate are unable to convert a significant fraction of either mutant enzyme population to the R quaternary structure, as has been observed previously for the Glu-50-->Ala enzyme. The curves for both the Ser-171-->Ala and Glu-50/Ser-171-->Ala enzymes obtained in the presence of substoichiometric amounts of PALA are linear combinations of the two extreme T and R states. The structural consequences of nucleotide binding to these two enzymes were also investigated. Most surprisingly, the direction and amplitude of the effect of ATP upon the double mutant enzyme were shown to vary depending upon the substrate analogue used.  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