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.
We generated replacement sets for three highly conserved residues, Pro196, Pro197 and His199, that flank the catalytic nucleophile, Cys198. Pro196 and Pro197 have restricted mobility that could be important for the structural transitions known to be essential for activity. To test this hypothesis we obtained and characterized 13 amino acid substitutions for Pro196, 14 for Pro197 and 14 for His199. All of the Pro196 and Pro197 variants, except P197R, and four of the His199 variants complemented TS-deficient Escherichia coli cells, indicating they had at least 1% of wild-type activity. For all His199 mutations, k(cat)/K(m) for substrate and cofactor decreased more than 40-fold, suggesting that the conserved hydrogen bond network co-ordinated by His199 is important for catalysis. Pro196 can be substituted with small hydrophilic residues with little loss in k(cat), but 15- to 23-fold increases in K(m)(dUMP). Small hydrophobic substitutions for Pro197 were most active, and the most conservative mutant, P197A, had only a 5-fold lower k(cat)/K(m)(dUMP) than wild-type TS. Several Pro196 and Pro197 variants were temperature sensitive. The small effects of Pro196 or Pro197 mutations on enzyme kinetics suggest that the conformational restrictions encoded by the Pro-Pro sequence are largely maintained when either member of the pair is mutated.  相似文献   

2.
The Escherichia coli aspartate (AATase) and tyrosine (TATase) aminotransferases share 43% sequence identity and 72% similarity, but AATase has only 0.08% and 0.01% of the TATase activities (k(cat)/K(m)) for tyrosine and phenylalanine, respectively. Approximately 5% of TATase activity was introduced into the AATase framework earlier both by rational design (six mutations, termed HEX) and by directed evolution (9-17 mutations). The enzymes realized from the latter procedure complement tyrosine auxotrophy in TATase deficient E. coli. HEX complements even more poorly than does wild-type AATase, even though the (k(cat)/K(m)) value for tyrosine exhibited by HEX is similar to those of the enzymes found from directed evolution. HEX, however, is characterized by very low values of K(m) and K(D) for dicarboxylic ligands, and by a particularly slow release for oxaloacetate, the product of the reaction with aspartate and a TCA cycle intermediate. These observations suggest that HEX exists largely as an enzyme-product complex in vivo. HEX was therefore subjected to a single round of directed evolution with selection for complementation of tyrosine auxotrophy. A variant with a single amino acid substitution, A293D, exhibited substantially improved TATase function in vivo. The A293D mutation alleviates the tight binding to dicarboxylic ligands as K(m)s for aspartate and alpha-ketoglutarate are >20-fold higher in the HEX + A293D construct compared to HEX. This mutation also increased k(cat)/K(m)(Tyr) threefold. A second mutation, I73V, elicited smaller but similar effects. Both residues are in close proximity to Arg292 and the mutations may function to modulate the arginine switch mechanism responsible for dual substrate recognition in TATases and HEX.  相似文献   

3.
In the present study, glutaryl-7-amino cephalosporanic acid acylase from Pseudomonas sp. strain 130 (CA130) was mutated to improve its enzymatic activity and stability. Based on the crystal structure of CA130, two series of amino acid residues, one from those directly involved in catalytic function and another from those putatively involved in surface charge, were selected as targets for site-directed mutagenesis. In the first series of experiments, several key residues in the substrate-binding pocket were substituted, and the genes were expressed in Escherichia coli for activity screening. Two of the mutants constructed, Y151alphaF and Q50betaN, showed two- to threefold-increased catalytic efficiency (k(cat)/K(m)) compared to wild-type CA130. Their K(m) values were decreased by ca. 50%, and the k(cat) values increased to 14.4 and 16.9 s(-1), respectively. The ability of these mutants to hydrolyze adipoyl 6-amino penicillinic acid was also improved. In the second series of mutagenesis, several mutants with enhanced stabilities were identified. Among them, R121betaA and K198betaA had a 30 to 58% longer half-life than wild-type CA130, and K198betaA and D286betaA showed an alkaline shift of optimal pH by about 1.0 to 2.0 pH units. To construct an engineered enzyme with the properties of both increased activity and stability, the double mutant Q50betaN/K198betaA was expressed. This enzyme was purified and immobilized for catalytic analysis. The immobilized mutant enzyme showed a 34.2% increase in specific activity compared to the immobilized wild-type CA130.  相似文献   

4.
We have designed a heterodimerizing leucine zipper system to target a radionuclide to prelocalized noninternalizing tumor-specific antibodies. The modular nature of the leucine zipper allows us to iteratively use design rules to achieve specific homodimer and heterodimer affinities. We present circular-dichroism thermal denaturation measurements on four pairs of heterodimerizing leucine zippers. These peptides are 47 amino acids long and contain four or five pairs of electrostatically attractive g <--> e' (i, i' +5) interhelical heterodimeric interactions. The most stable heterodimer consists of an acidic leucine zipper and a basic leucine zipper that melt as homodimers in the micro (T(m) = 28 degrees C) or nanomolar (T(m) = 40 degrees C) range, respectively, but heterodimerize with a T(m) >90 degrees C, calculated to represent femtamolar affinities. Modifications to this pair of acidic and basic zippers, designed to destabilize homodimerization, resulted in peptides that are unstructured monomers at 4 microM and 6 degrees C but that heterodimerize with a T(m) = 74 degrees C or K(d(37)) = 1.1 x 10(-11) M. A third heterodimerizing pair was designed to have a more neutral isoelectric focusing point (pI) and formed a heterodimer with T(m) = 73 degrees C. We can tailor this heterodimerizing system to achieve pharmacokinetics aimed at optimizing targeted killing of cancer cells.  相似文献   

5.
Singh N  Frushicheva MP  Warshel A 《Proteins》2012,80(4):1110-1122
The current challenge in designing effective drugs against HIV-1 is to find novel candidates with high potency, but with a lower susceptibility to mutations associated with drug resistance. Trying to address this challenge, we developed in our previous study (Ishikita and Warshel, Angew Chem Int Ed Engl 2008; 47:697-700) a novel computational strategy for fighting drug resistance by predicting the likely moves of the virus through constraints on binding and catalysis. This has been based on calculating the ratio between the vitality values ((K(i) k(cat)/K(M))(mutant)/(K(i) k(cat)/K(M))(wild-type)) and using it as a guide for predicting the moves of the virus. The corresponding calculations of the binding affinity, K(i), were carried out using the semi-macroscopic version of the protein dipole Langevin dipole (PDLD/S) in its linear response approximation (LRA) in its β version (PDLD/S-LRA/β). We also calculate the proteolytic efficiency, k(cat)/K(M), by evaluating the transition state (TS) binding free energies using the PDLD/S-LRA/β method. Here we provide an extensive validation of our strategy by calculating the vitality of six existing clinical and experimental drug candidates. It is found that the computationally determined vitalities correlate reasonably well with those derived from the corresponding experimental data. This indicates that the calculated vitality may be used to identify mutations that would be most effective for the survival of the virus. Thus, it should be possible to use our approach in screening for mutations that would provide the most effective resistance to any proposed antiviral drug. This ability should be very useful in guiding the design of drug molecules that will lead to the slowest resistance.  相似文献   

6.
To elucidate the functions of the surface loops of VIIa, we prepared two mutants, VII-30 and VII-39. The VII-30 mutant had all of the residues in the 99 loop replaced with those of trypsin. In the VII-39 mutant, both the 99 and 170 loops were replaced with those of trypsin. The k(cat)/K(m) value for hydrolysis of the chromogenic peptidyl substrate S-2288 by VIIa-30 (103 mm(-)1s(-)1) was 3-fold higher than that of wild-type VIIa (30.3 mm(-)1 s(-)1) in the presence of soluble tissue factor (sTF). This enhancement was due to a decrease in the K(m) value but not to an increase in the k(cat) value. On the other hand, the k(cat)/K(m) value for S-2288 hydrolysis by VIIa-39 (17.9 mm(-)1 s(-)1) was 18-fold higher than that of wild-type (1.0 mm(-)1 s(-)1) in the absence of sTF, and the value was almost the same as that of wild-type measured in the presence of sTF. This enhancement was due to not only a decrease in the K(m) value but also to an increase in the k(cat) value. These results were in good agreement with their susceptibilities to a subsite 1-directed serine protease inhibitor. In our previous paper (Soejima, K., Mizuguchi, J., Yuguchi, M., Nakagaki, T., Higashi, S., and Iwanaga, S. (2001) J. Biol. Chem. 276, 17229-17235), the replacement of the 170 loop of VIIa with that of trypsin induced a 10-fold enhancement of the k(cat) value for S-2288 hydrolysis as compared with that of wild-type VIIa in the absence of sTF. These results suggested that the 99 and the 170 loop structures of VIIa independently affect the K(m) and k(cat) values, respectively. Furthermore, we studied the effect of mutations on proteolytic activity toward S-alkylated lysozyme as a macromolecular substrate and the activation of natural macromolecular substrate factor X.  相似文献   

7.
Hirano N  Haruki M  Morikawa M  Kanaya S 《Biochemistry》2000,39(43):13285-13294
A genetic method for isolating a mutant enzyme of ribonuclease HI (RNase HI) from Thermus thermophilus HB8 with enhanced activity at moderate temperatures was developed. T. thermophilus RNase HI has an ability to complement the RNase H-dependent temperature-sensitive (ts) growth phenotype of Escherichia coli MIC3001. However, this complementation ability was greatly reduced by replacing Asp(134), which is one of the active site residues, with His, probably due to a reduction in the catalytic activity. Random mutagenesis of the gene encoding the resultant D134H enzyme, followed by screening for second-site revertants, allowed us to isolate three single mutations (Ala(12) --> Ser, Lys(75) --> Met, and Ala(77) --> Pro) that restore the normal complementation ability to the D134H enzyme. These mutations were individually or simultaneously introduced into the wild-type enzyme, and the kinetic parameters of the resultant mutant enzymes for the hydrolysis of a DNA-RNA-DNA/DNA substrate were determined at 30 degrees C. Each mutation increased the k(cat)/K(m) value of the wild-type enzyme by 2.1-4.8-fold. The effects of the mutations on the enzymatic activity were roughly cumulative, and the combination of these three mutations increased the k(cat)/K(m) value of the wild-type enzyme by 40-fold (5.5-fold in k(cat)). Measurement of thermal stability of the mutant enzymes with circular dichroism spectroscopy in the presence of 1 M guanidine hydrochloride and 1 mM dithiothreitol showed that the T(m) value of the triple mutant enzyme, in which all three mutations were combined, was comparable to that of the wild-type enzyme (75.0 vs 77.4 degrees C). These results demonstrate that the activity of a thermophilic enzyme can be improved without a cost of protein stability.  相似文献   

8.
We have performed random mutagenesis coupled with selection to isolate mutant enzymes with high catalytic activities at low temperature from thermophilic 3-isopropylmalate dehydrogenase (IPMDH) originally isolated from Thermus thermophilus. Five cold-adapted mutant IPMDHs with single-amino-acid substitutions were obtained and analyzed. Kinetic analysis revealed that there are two types of cold-adapted mutant IPMDH: k(cat)-improved (improved in k(cat)) and K(m)-improved (improved in k(cat)/K(m)) types. To determine the mechanisms of cold adaptation of these mutants, thermodynamic parameters were estimated and compared with those of the Escherichia coli wild-type IPMDH. The Delta G(m) values for Michaelis intermediate formation of the k(cat)-improved-type enzymes were larger than that of the T. thermophilus wild-type IPMDH and similar to that of the E. coli wild-type IPMDH. The Delta G(m) values of K(m)-improved-type enzymes were smaller than that of the T. thermophilus wild-type IPMDH. Fitting of NAD(+) binding was improved in the K(m)-improved-type enzymes. The two types of cold-adapted mutants employed one of the two strategies of E. coli wild-type IPMDH: relative destabilization of the Michaelis complex in k(cat)-improved-type, and destabilization of the rate-limiting step in K(m)-improved type mutants. Some cold-adapted mutant IPMDHs retained thermostability similar to that of the T. thermophilus wild-type IPMDH.  相似文献   

9.
A rapid purification procedure produces milligram amounts of the T7 gene 4A' primase/helicase, 4B helicase, and the wild-type 4AB proteins expressed from the clones described in the accompanying paper (Rosenberg, A. H., Patel, S. S., Johnson, K. A., and Studier, F. W. (1992) J. Biol. Chem. 267, 15005-15012). Purified 4A' protein (in which the wild-type methionine at amino acid 64 has been replaced by leucine to eliminate the 4B initiation codon) appears to be equivalent to the wild-type 4A protein in primase, helicase, and NTPase activities. Gel filtration chromatography and polyacrylamide gel electrophoresis of native proteins indicate that the 4A' and 4B proteins form homodimers and heterodimers in solution. Heterodimer formation presumably accounts for an observed 3-fold increase in the primase activity of 4A' upon addition of 4B that lacks primase activity of its own. Steady-state k(cat) and Km values for hydrolysis of the nucleoside triphosphates ATP, dATP, dTTP, and dGTP were measured for 4A', 4B, 4A'B (1:1), and wild-type 4AB (1:2) proteins. The dependence of the dNTPase activities on the concentration was hyperbolic, suggesting single or noncooperative binding sites, whereas ATPase activity was sigmoidal, suggesting more than one ATP binding site. The k(cat)/Km ratios for hydrolysis of the dNTPs by the four protein preparations were within a factor of 6 of each other. The 1:1 mixture of 4A'B had the highest k(cat)/Km ratios, with a preference for dATP and dTTP.  相似文献   

10.
Glutathione transferase (GST) A3-3 is the most efficient human steroid double-bond isomerase known. The activity with Delta(5)-androstene-3,17-dione is highly dependent on the phenolic hydroxyl group of Tyr-9 and the thiolate of glutathione. Removal of these groups caused an 1.1 x 10(5)-fold decrease in k(cat); the Y9F mutant displayed a 150-fold lower isomerase activity in the presence of glutathione and a further 740-fold lower activity in the absence of glutathione. The Y9F mutation in GST A3-3 did not markedly decrease the activity with the alternative substrate 1-chloro-2,4-dinitrobenzene. Residues Phe-10, Leu-111, and Ala-216 selectively govern the activity with the steroid substrate. Mutating residue 111 into phenylalanine caused a 25-fold decrease in k(cat)/K(m) for the steroid isomerization. The mutations A216S and F10S, separate or combined, affected the isomerase activity only marginally, but with the additional L111F mutation k(cat)/K(m) was reduced to 0.8% of that of the wild-type value. In contrast, the activities with 1-chloro-2,4-dinitrobenzene and phenethylisothiocyanate were not largely affected by the combined mutations F10S/L111F/A216S. K(i) values for Delta(5)-androstene-3,17-dione and Delta(4)-androstene-3,17-dione were increased by the triple mutation F10S/L111F/A216S. The pK(a) of the thiol group of active-site-bound glutathione, 6.1, increased to 6.5 in GST A3-3/Y9F. The pK(a) of the active-site Tyr-9 was 7.9 for the wild-type enzyme. The pH dependence of k(cat)/K(m) of wild-type GST A3-3 for the isomerase reaction displays two kinetic pK(a) values, 6.2 and 8.1. The basic limb of the pH dependence of k(cat) and k(cat)/K(m) disappears in the Y9F mutant. Therefore, the higher kinetic pK(a) reflects ionization of Tyr-9, and the lower one reflects ionization of glutathione. We propose a reaction mechanism for the double-bond isomerization involving abstraction of a proton from C4 in the steroid accompanied by protonation of C6, the thiolate of glutathione serving as a base and Tyr-9 assisting by polarizing the 3-oxo group of the substrate.  相似文献   

11.
Apparent kinetic constants k(cat) and K(m) were determined for tyrocidine thioesterase (TycC TE) using randomized peptide N-acetylcysteamine thioesters as substrate analogues. The enzyme has been found to be adequately active for the synthesis of positional-scanning libraries for novel antibiotic screening with reduced k(cat)/K(m) in the range of 2 to 82 folds lower than that of the wild-type sequence  相似文献   

12.
Xylose reductase from the yeast Candida tenuis (CtXR) is a family 2 member of the aldo-keto reductase (AKR) superfamily of proteins and enzymes. Active site His-113 is conserved among AKRs, but a unified mechanism of how it affects catalytic activity is outstanding. We have replaced His-113 by alanine using site-directed mutagenesis, determined a 2.2 A structure of H113A mutant bound to NADP(+), and compared catalytic reaction profiles of NADH-dependent reduction of different aldehydes catalyzed by the wild type and the mutant. Deuterium kinetic isotope effects (KIEs) on k(cat) and k(cat)/K(m xylose) show that, relative to the wild type, the hydride transfer rate constant (k(7) approximately 0.16 s(-1)) has decreased about 1000-fold in H113A whereas xylose binding was not strongly affected. No solvent isotope effect was seen on k(cat) and k(cat)/K(m xylose) for H113A, suggesting that proton transfer has not become rate-limiting as a result of the mutation. The pH profiles of log(k(cat)/K(m xylose)) for the wild type and H113A decreased above apparent pK(a) values of 8.85 and 7.63, respectively. The DeltapK(a) of -1.2 pH units likely reflects a proximally disruptive character of the mutation, affecting the position of Asp-50. A steady-state kinetic analysis for H113A-catalyzed reduction of a homologous series of meta-substituted benzaldehyde derivatives was carried out, and quantitative structure-reactivity correlations were used to factor the observed kinetic substituent effect on k(cat) and k(cat)/K(m aldehyde) into an electronic effect and bonding effects (which are lacking in the wild type). Using the Hammett sigma scale, electronic parameter coefficients (rho) of +0.64 (k(cat)) and +0.78 (k(cat)/K(m aldehyde)) were calculated and clearly differ from rho(k(cat)/K(aldehyde)) and rho(k(cat)) values of +1.67 and approximately 0.0, respectively, for the wild-type enzyme. Hydride transfer rate constants of H113A, calculated from kinetic parameters and KIE data, display a substituent dependence not seen in the corresponding wild-type enzyme rate constants. An enzymic mechanism is proposed in which His-113, through a hydrogen bond from Nepsilon2 to aldehyde O1, assists in catalysis by optimizing the C=O bond charge separation and orbital alignment in the ternary complex.  相似文献   

13.
To identify protein split sites quickly, a selection procedure by using chloramphenicol acetyl transferase (CAT) as reporter was introduced to search for folded protein fragments from libraries generated by random digestion and reassembly of the target gene, which yielded an abundant amount of DNA fragments with controllable lengths. Experimental results of tryptophan synthase alpha subunit (TSα) and TEM‐1 β‐lactamase agreed well with what the literature has reported. The solubility of these fragments correlated roughly with the minimum inhibitory concentrations of the CAT fusions. The application of this dissection protocol to protein fragment complementation assay (PCA) was evaluated using aminoglycoside‐3′‐phosphotransferase I (APH(3′)‐I) as a model protein. Three nearly bisectional sites and a number of possible split points were identified, and guided by this result, four novel pairs of fragments were tested for complementation. Three out of four pairs partially restored the APH activity with the help of leucine zippers, and a truncated but active APH(3′)‐I (Δ1–25) was also found. Finally, the weakly active APH(3′)‐I‐(1–253)NZ/CZ (254–271) containing a short 18 residue tag was further improved by error‐prone PCR, and a best mutant was obtained showing a fourfold improvement after just one round of evolution. These results demonstrate that protein random dissection based on the CAT selection can provide an efficient search for protein breakage points and guide the design of fragments for protein complementation assay. Furthermore, more active fragment pairs can be achieved with the classical directed evolution approach.  相似文献   

14.
Random PCR mutagenesis was applied to the Thermus thermophilus xylA gene encoding xylose isomerase. Three cold-adapted mutants were isolated with the following amino-acid substitutions: E372G, V379A (M-1021), E372G, F163L (M-1024) and E372G (M-1026). The wild-type and mutated xylA genes were cloned and expressed in Escherichia coli HB101 using the vector pGEM-T Easy, and their physicochemical and catalytic properties were determined. The optimum pH for xylose isomerization activity for the mutants was approximately 7.0, which is similar to the wild-type enzyme. Compared with the wild-type, the mutants were active over a broader pH range. The mutants exhibited up to nine times higher catalytic rate constants (k(cat)) for d-xylose compared with the wild-type enzyme at 60 degrees C, but they did not show any increase in catalytic efficiency (k(cat)/K(m)). For d-glucose, both the k(cat) and the k(cat)/K(m) values for the mutants were increased compared with the wild-type enzyme. Furthermore, the mutant enzymes exhibited up to 255 times higher inhibition constants (K(i)) for xylitol than the wild-type, indicating that they are less inhibited by xylitol. The thermal stability of the mutated enzymes was poorer than that of the wild-type enzyme. The results are discussed in terms of increased molecular flexibility of the mutant enzymes at low temperatures.  相似文献   

15.
Human thrombin utilizes Na+ as a driving force for the cleavage of substrates mediating its procoagulant, prothrombotic, and signaling functions. Murine thrombin has Asp-222 in the Na+ binding site of the human enzyme replaced by Lys. The charge reversal substitution abrogates Na+ activation, which is partially restored with the K222D mutation, and ensures high activity even in the absence of Na+. This property makes the murine enzyme more resistant to the effect of mutations that destabilize Na+ binding and shift thrombin to its anticoagulant slow form. Compared with the human enzyme, murine thrombin cleaves fibrinogen and protein C with similar k(cat)/K(m) values but activates PAR1 and PAR4 with k(cat)/K(m) values 4- and 26-fold higher, respectively. The significantly higher specificity constant toward PAR4 accounts for the dominant role of this receptor in platelet activation in the mouse. Murine thrombin can also cleave substrates carrying Phe at P1, which potentially broadens the repertoire of molecular targets available to the enzyme in vivo.  相似文献   

16.
Pan JJ  Yang LW  Liang PH 《Biochemistry》2000,39(45):13856-13861
Undecaprenyl pyrophosphate synthase (UPPs) catalyzes condensation of eight molecules of isopentenyl pyrophosphate with farnesyl pyrophosphate to yield C(55)-undecaprenyl pyrophosphate. We have mutated the aspartates and glutamates in the five conserved regions (I to V) of UPPs protein sequence to evaluate their effects on substrate binding and catalysis. The mutant enzymes including D26A, E73A, D150A, D190A, E198A, E213A, D218A, and D223A were expressed and purified to great homogeneity. Kinetic analyses of these mutant enzymes indicated that the substitution of D26 in region I with alanine resulted in a 10(3)-fold decrease of k(cat) value compared to wild-type UPPs. Its IPP K(m) value has only minor change. The mutagenesis of D150A has caused a much lower IPP affinity with IPP K(m) value 50-fold larger than that of wild-type UPPs but did not affect the FPP K(m) and the k(cat). The E213A mutant UPPs has a 70-fold increased IPP K(m) value and has a 100-fold decreased k(cat) value compared to wild-type. These results suggest that D26 of region I is critical for catalysis and D150 in region IV plays a significant role of IPP binding. The E213 residue in region V is also important in IPP binding as well as catalysis. Other mutant UPPs enzymes in this study have shown no significant change (<5-fold) of k(cat) with exception of E73A and D218A. Both enzymes have 10-fold lower k(cat) value relative to wild-type UPPs.  相似文献   

17.
Tripp BC  Ferry JG 《Biochemistry》2000,39(31):9232-9240
Four glutamate residues in the prototypic gamma-class carbonic anhydrase from Methanosarcina thermophila (Cam) were characterized by site-directed mutagenesis and chemical rescue studies. Alanine substitution indicated that an external loop residue, Glu 84, and an internal active site residue, Glu 62, are both important for CO(2) hydration activity. Two other external loop residues, Glu 88 and Glu 89, are less important for enzyme function. The two E84D and -H variants exhibited significant activity relative to wild-type activity in pH 7.5 MOPS buffer, suggesting that the original glutamate residue could be substituted with other ionizable residues with similar pK(a) values. The E84A, -C, -K, -Q, -S, and -Y variants exhibited large decreases in k(cat) values in pH 7.5 MOPS buffer, but only exhibited small changes in k(cat)/K(m). These same six variants were all chemically rescued by pH 7.5 imidazole buffer, with 23-46-fold increases in the apparent k(cat). These results are consistent with Glu 84 functioning as a proton shuttle residue. The E62D variant exhibited a 3-fold decrease in k(cat) and a 2-fold decrease in k(cat)/K(m) relative to those of the wild type in pH 7.5 MOPS buffer, while other substitutions (E62A, -C, -H, -Q, -T, and -Y) resulted in much larger decreases in both k(cat) and k(cat)/K(m). Imidazole did not significantly increase the k(cat) values and slightly decreased the k(cat)/K(m) values of most of the Glu 62 variants. These results indicate a primary preference for a carboxylate group at position 62, and support a proposed catalytic role for residue Glu 62 in the CO(2) hydration step, but do not definitively establish its role in the proton transport step.  相似文献   

18.
Lietz EJ  Truher H  Kahn D  Hokenson MJ  Fink AL 《Biochemistry》2000,39(17):4971-4981
Lysine 73 is a conserved active-site residue in the class A beta-lactamases, as well as other members of the serine penicillin-sensitive enzyme family; its role in catalysis remains controversial and uncertain. Mutation of Lys73 to alanine in the beta-lactamase from Bacillus licheniformis resulted in a substantial reduction in both turnover rate (k(cat)) and catalytic efficiency (k(cat)/K(m)), and a very significant shift in pK(1) to higher pH in the bell-shaped pH-rate profiles (k(cat)/K(m)) for several penicillin and cephalosporin substrates. The increase in pK(1) is consistent with the removal of the positive ammonium group of the lysine from the proximity of Glu166, to which the acid limb has been ascribed. The alkaline limb of the k(cat)/K(m) vs profiles is not shifted appreciably, as might have been expected if this limb reflected the ionization of Lys73 in the wild-type enzyme. The k(cat)/K(m) at the pH optimum for the mutant was down about 200-fold for penicillins and around 10(4) for cephalosporins, compared to the wild-type, suggesting significant differences in the mechanisms for catalysis of penicillins compared to cephalosporins. Burst kinetics were observed with several substrates assayed with K73A beta-lactamase, indicating an underlying branched-pathway kinetic scheme, and rate-limiting deacylation. FTIR analysis was used to determine whether acylation or deacylation was rate-limiting. In general, acylation was the rate-limiting step for cephalosporin substrates, whereas deacylation was rate-limiting for penicillin substrates. The results indicate that Lys73 plays an important role in both the acylation and deacylation steps of the catalytic mechanism. The effects of this mutation (K73A) indicate that Lys73 does not function as a general base in the catalytic mechanism of beta-lactamase. The existence of bell-shaped pH-rate profiles for the K73A variant suggests that Lys73 is not directly responsible for either limb in such plots. It is likely that both Glu166 and Lys73 are important to each other in terms of maintaining the optimum electrostatic environment for fully efficient catalytic activity to occur.  相似文献   

19.
The affinity between the 1–157 and 158–238 fragments of green fluorescent protein (GFP) is too low for spontaneous in vivo reassembly of the protein upon co-expression of the two fragments. This prevents chromophore maturation and the cells lack GFP fluorescence. We have utilized the very high affinity between the two EF-hands of calbindin D9k to facilitate GFP assembly from its fragments and to introduce a calcium dependent molecular switch. In GFPN-EF1, residues 1–157 of GFP are fused to residues 1–43 of calbindin, and in EF2-GFPC, residues 44–75 of calbindin are fused to residues 158–238 of GFP. When co-expressed, GFPN-EF1 and EF2-GFPC associate spontaneously and rapidly resulting in a folded reconstituted protein with bright GFP fluorescence. The high affinity of GFPN-EF1 for EF2-GFPC leads to brighter fluorescence of the cells compared to cells with a control constructs carrying leucine zippers (Wilson et al., Nature Methods 2004;3:255). The complex of GFPN-EF1 and EF2-GFPC was purified from cells using metal-ion chelate chromatography and the temperature dependence of GFP fluorescence was found to be calcium dependent. The GFPN-EF1 and EF2-GFPC fragments were separated by ion exchange chromatography. The assembly of the fragments was found to be reversible and the complex was regained upon mixing, as evidenced by surface plasmon resonance (SPR) data. The affinity between GFPN-EF1 and EF2-GFPC as well as rates of association and dissociation were found to be Ca2+-dependent.  相似文献   

20.
In the blood coagulation cascade, thrombin cleaves fibrinopeptides A and B from fibrinogen revealing sites for fibrin polymerization that lead to insoluble clot formation. Factor XIII stabilizes this clot by catalyzing the formation of intermolecular cross-links in the fibrin network. Thrombin activates the Factor XIII a(2) dimer by cleaving the Factor XIII activation peptide segment at the Arg(37)-Gly(38) peptide bond. Using a high performance liquid chromatography assay, the kinetic constants K(m), k(cat), and k(cat)/K(m) were determined for thrombin hydrolysis of fibrinogen Aalpha-(7-20), Factor XIII activation peptide-(28-41), and Factor XIII activation peptide-(28-41) with a Val(34) to Leu substitution. This Val to Leu mutation has been correlated with protection from myocardial infarction. In the absence of fibrin, the Factor XIII activation peptide-(28-41) exhibits a 10-fold lower k(cat)/K(m) value than fibrinogen Aalpha-(7-20). With the Factor XIII V34L mutation, decreases in K(m) and increases in k(cat) produce a 6-fold increase in k(cat)/K(m) relative to the wild-type Factor XIII sequence. A review of the x-ray crystal structures of known substrates and inhibitors of thrombin leads to a hypothesis that the new Leu generates a peptide with more extensive interactions with the surface of thrombin. As a result, the Factor XIII V34L is proposed to be susceptible to wasteful conversion of zymogen to activated enzyme. Premature depletion may provide cardioprotective effects.  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