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.
Zhan X  Carpenter RA  Ellis HR 《Biochemistry》2008,47(7):2221-2230
The two-component alkanesulfonate monooxygenase system from Escherichia coli includes an FMN reductase (SsuE) and an FMNH2-dependent alkanesulfonate monooxygenase (SsuD) involved in the acquisition of sulfur from alkanesulfonates during sulfur starvation. The SsuD enzyme directly catalyzes the oxidation of alkanesulfonate to aldehyde and sulfite in the presence of O2 and FMNH2. The goal of these studies was to investigate the kinetic mechanism of SsuD through rapid reaction kinetics and substrate binding studies. The SsuD enzyme shows a clear preference for FMNH2 (Kd, 0.32 +/- 0.15 microM) compared to FMN (Kd, 10.2 +/- 0.4 microM) with a 1:1 binding stoichiometry for each form of the flavin. The kinetic trace of premixed SsuD and FMNH2 mixed with oxygenated buffer was best fit to a double exponential with no observed formation of the C4a-(hydro)peroxyflavin. However, when FMNH2 was mixed with SsuD and oxygenated buffer an initial fast phase (kobs, 12.9 s-1) was observed, suggesting that the mixing order is critical for the accumulation of the C4a-(hydro)peroxyflavin. Results from fluorimetric titrations with octanesulfonate imply that reduced flavin must bind first to promote octanesulfonate binding. When octanesulfonate was included in the kinetic studies the C4a-(hydro)peroxyflavin was observed at 370 nm when FMNH2 was not premixed with SsuD, which correlated with an increase in octanal product. There was a clear hyperbolic dependence on octanesulfonate binding, indicating that octanesulfonate binds in rapid equilibrium, and further results indicated there was a second isomerization step following binding. These results suggest that an ordered substrate binding mechanism is important in the desulfonation reaction by SsuD with reduced flavin binding first followed by either O2 or octanesulfonate.  相似文献   

2.
3.
JM Robbins  HR Ellis 《Biochemistry》2012,51(32):6378-6387
The alkanesulfonate monooxygenase enzyme (SsuD) catalyzes the oxygenolytic cleavage of a carbon-sulfur bond from sulfonated substrates. A mechanism involving acid-base catalysis has been proposed for the desulfonation mechanism by SsuD. In the proposed mechanism, base catalysis is involved in abstracting a proton from the alkane peroxyflavin intermediate, while acid catalysis is needed for the protonation of the FMNO(-) intermediate. The pH profiles of k(cat) indicate that catalysis by SsuD requires a group with a pK(a) of 6.6 ± 0.2 to be deprotonated and a second group with a pK(a) of 9.5 ± 0.1 to be protonated. The upper pK(a) value was not present in the pH profiles of k(cat)/K(m). Several conserved amino acid residues (His228, His11, His333, Cys54, and Arg226) have been identified as having potential catalytic importance due to the similar spatial arrangements with close structural and functional relatives of SsuD. Substitutions to these amino acid residues were generated, and the pH dependencies were evaluated and compared to wild-type SsuD. Although a histidine residue was previously proposed to be the active site base, the His variants possessed similar steady-state kinetic parameters as wild-type SsuD. Interestingly, R226A and R226K SsuD variants possessed undetectable activity, and there was no detectable formation of the C4a-(hydro)peroxyflavin intermediate for the Arg226 SsuD variants. Guanidinium rescue with the R226A SsuD variant resulted in the recovery of 1.5% of the wild-type SsuD k(cat) value. These results implicate Arg226 playing a critical role in catalysis and provide essential insights into the mechanistic steps that guide the SsuD desulfonation process.  相似文献   

4.
The structure of the flavin-dependent alkanesulfonate monooxygenase (SsuD) exists as a TIM-barrel structure with an insertion region located over the active site that contains a conserved arginine (Arg297) residue present in all SsuD homologues. Substitution of Arg297 with alanine (R297A SsuD) or lysine (R297K SsuD) was performed to determine the functional role of this conserved residue in SsuD catalysis. While the more conservative R297K SsuD possessed a lower k(cat)/K(m) value (0.04 ± 0.01 μM(-1) min(-1)) relative to wild-type (1.17 ± 0.22 μM(-1) min(-1)), there was no activity observed with the R297A SsuD variant. Each of the arginine variants had similar K(d) values for flavin binding as wild-type SsuD (0.32 ± 0.15 μM), but there was no measurable binding of octanesulfonate. The low levels of activity for the R297A and R297K SsuD variants correlated with the absence of any detectable C4a-(peroxy)flavin formation in stopped-flow kinetic studies. Single-turnover experiments were performed in the presence of SsuE to evaluate both the reductive and oxidative half-reaction. With wild-type SsuD a lag phase is observed following the reductive half-reaction by SsuE that represents flavin transfer or conformational changes associated with the binding of substrates. Evaluation of the Arg297 SsuD variants in the presence of SsuE showed no lag phase following reduction by SsuE, and the flavin was oxidized immediately following the reductive half-reaction. These results corresponded with a lack of detectable changes in the proteolytic susceptibility of R297A and R297K SsuD in the presence of reduced flavin and/or octanesulfonate, signifying the absence of a conformational change in these variants with the substitution of Arg297.  相似文献   

5.
Cytochrome P450 27C1 (P450 27C1) is a retinoid desaturase expressed in the skin that catalyzes the formation of 3,4-dehydroretinoids from all-trans retinoids. Within the skin, retinoids are important regulators of proliferation and differentiation. In vivo, retinoids are bound to cellular retinol-binding proteins (CRBPs) and cellular retinoic acid–binding proteins (CRABPs). Interaction with these binding proteins is a defining characteristic of physiologically relevant enzymes in retinoid metabolism. Previous studies that characterized the catalytic activity of human P450 27C1 utilized a reconstituted in vitro system with free retinoids. However, it was unknown whether P450 27C1 could directly interact with holo-retinoid-binding proteins to receive all-trans retinoid substrates. To assess this, steady-state kinetic assays were conducted with free all-trans retinoids and holo-CRBP-1, holo-CRABP-1, and holo-CRABP-2. For holo-CRBP-1 and holo-CRABP-2, the kcat/Km values either decreased 5-fold or were equal to the respective free retinoid values. The kcat/Km value for holo-CRABP-1, however, decreased ∼65-fold in comparison with reactions with free all-trans retinoic acid. These results suggest that P450 27C1 directly accepts all-trans retinol and retinaldehyde from CRBP-1 and all-trans retinoic acid from CRABP-2, but not from CRABP-1. A difference in substrate channeling between CRABP-1 and CRABP-2 was also supported by isotope dilution experiments. Analysis of retinoid transfer from holo-CRABPs to P450 27C1 suggests that the decrease in kcat observed in steady-state kinetic assays is due to retinoid transfer becoming rate-limiting in the P450 27C1 catalytic cycle. Overall, these results illustrate that, like the CYP26 enzymes involved in retinoic acid metabolism, P450 27C1 interacts with cellular retinoid-binding proteins.  相似文献   

6.
Onconase® (ONC) is a homolog of bovine pancreatic ribonuclease (RNase A) from the frog Rana pipiens. ONC displays antitumoral activity and is in advanced clinical trials for the treatment of cancer. Here, we report the first atomic structures of ONC-nucleic acid complexes: a T89N/E91A ONC-5′-AMP complex at 1.65 Å resolution and a wild-type ONC-d(AUGA) complex at 1.90 Å resolution. The latter structure and site-directed mutagenesis were used to reveal the atomic basis for substrate recognition and turnover by ONC. The residues in ONC that are proximal to the scissile phosphodiester bond (His10, Lys31, and His97) and uracil nucleobase (Thr35, Asp67, and Phe98) are conserved from RNase A and serve to generate a similar bell-shaped pH versus kcat/KM profile for RNA cleavage. Glu91 of ONC forms two hydrogen bonds with the guanine nucleobase in d(AUGA), and Thr89 is in close proximity to that nucleobase. Installing a neutral or cationic residue at position 91 or an asparagine residue at position 89 virtually eliminated the 102-fold guanine:adenine preference of ONC. A variant that combined such substitutions, T89N/E91A ONC, actually preferred adenine over guanine. In contrast, installing an arginine residue at position 91 increased the guanine preference and afforded an ONC variant with the highest known kcat/KM value. These data indicate that ONC discriminates between guanine and adenine by using Coulombic interactions and a network of hydrogen bonds. The structure of the ONC-d(AUGA) complex was also used to probe other aspects of catalysis. For example, the T5R substitution, designed to create a favorable Coulombic interaction between ONC and a phosphoryl group in RNA, increased ribonucleolytic activity by twofold. No variant, however, was more toxic to human cancer cells than wild-type ONC. Together, these findings provide a cynosure for understanding catalysis of RNA cleavage in a system of high medicinal relevance.  相似文献   

7.
A comparison of the primary structures among psychrophilic, mesophilic, and thermophilic subtilases revealed that the turn between the β8 and β9 strands (β8-β9 turn, BPN′ numbering) of psychrophilic subtilases are more flexible than those of their mesophilic and thermophilic counterparts. To investigate the relationship between structure of this turn and enzyme activity as well as thermostability of mesophilic subtilisin Carlsberg (sC), we analyzed 6 mutants of sC with a single, double, or triple Gly or Ala substitutions for Pro210Thr211Asn212 at the β8-β9 turn. Among the single Gly substitutions, the P210G substitution most significantly (1.5-fold) increased the specific activity on N-succinyl-Ala-Ala-Pro-Phe-p-nitroanilide (AAPF) substrate and 12-fold decreased the thermostability. All mutants tested showed the increased kcat for the AAPF substrate and reduced thermostability compared with the wild-type sC. The kcat values of the P210G, P210G/T211G, and P210G/T211G/N212G mutants were 1.5-, 1.7-, and 1.8-fold higher than that of the wild-type sC. There were significant positive correlations between kcat and thermal inactivation rates as well as kcat and Km of the wild-type and mutants. These results demonstrate that the structure of β8-β9 turn, despite its distance from the active site, has significant effects on the catalytic rate and thermostability of sC through a global network of intramolecular interactions and suggest that the lack of flexibility of this turn stabilizes the wild-type sC against thermal inactivation in compensation for some loss of catalytic activity.  相似文献   

8.
Lysine 315 of mouse polyamine amine oxidase corresponds to a lysine residue that is conserved in the flavoprotein amine oxidases of the monoamine oxidase structural family. In several structures, this lysine residue forms a hydrogen bond to a water molecule that is hydrogen-bonded to the flavin N(5). Mutation of Lys315 in polyamine oxidase to methionine was previously shown to have no effect on the kinetics of the reductive half-reaction of the enzyme (M. Henderson Pozzi, V. Gawandi, P.F. Fitzpatrick, Biochemistry 48 (2009) 1508-1516). In contrast, the mutation does affect steps in the oxidative half-reaction. The kcat value is unaffected by the mutation; this kinetic parameter likely reflects product release. At pH 10, the kcat/Km value for oxygen is 25-fold lower in the mutant enzyme. The kcat/KO2 value is pH-dependent for the wild-type enzyme, decreasing below a pKa of 7.0, while this kinetic parameter for the mutant enzyme is pH-independent. This is consistent with the neutral form of Lys315 being required for more rapid flavin oxidation. The solvent isotope effect on the kcat/KO2 value increases from 1.4 in the wild-type enzyme to 1.9 in the mutant protein, and the solvent inventory changes from linear to bowed. The effects of the mutation can be explained by the lysine orienting the bridging water so that it can accept the proton from the flavin N(5) during flavin oxidation. In the mutant enzyme the lysine amine would be replaced by a water chain.  相似文献   

9.
The loop following helix α2 in glutathione transferase P1-1 has two conserved residues, Cys48 and Tyr50, important for glutathione (GSH) binding and catalytic activity. Chemical modification of Cys48 thwarts the catalytic activity of the enzyme, and mutation of Tyr50 generally decreases the kcat value and the affinity for GSH in a differential manner. Cys48 and Tyr50 were targeted by site-specific mutations and chemical modifications in order to investigate how the α2 loop modulates GSH binding and catalysis. Mutation of Cys48 into Ala increased KMGSH 24-fold and decreased the binding energy of GSH by 1.5 kcal/mol. Furthermore, the protein stability against thermal inactivation and chemical denaturation decreased. The crystal structure of the Cys-free variant was determined, and its similarity to the wild-type structure suggests that the mutation of Cys48 increases the flexibility of the α2 loop rather than dislocating the GSH-interacting residues. On the other hand, replacement of Tyr50 with Cys, producing mutant Y50C, increased the Gibbs free energy of the catalyzed reaction by 4.8 kcal/mol, lowered the affinity for S-hexyl glutathione by 2.2 kcal/mol, and decreased the thermal stability. The targeted alkylation of Cys50 in Y50C increased the affinity for GSH and protein stability. Characterization of the most active alkylated variants, S-n-butyl-, S-n-pentyl-, and S-cyclobutylmethyl-Y50C, indicated that the affinity for GSH is restored by stabilizing the α2 loop through positioning of the key residue into the lock structure of the neighboring subunit. In addition, kcat can be further modulated by varying the structure of the key residue side chain, which impinges on the rate-limiting step of catalysis.  相似文献   

10.
LodA is a novel lysine-ε-oxidase which possesses a cysteine tryptophylquinone cofactor. It is the first tryptophylquinone enzyme known to function as an oxidase. A steady-state kinetic analysis shows that LodA obeys a ping-pong kinetic mechanism with values of kcat of 0.22 ± 0.04 s−1, Klysine of 3.2 ± 0.5 μM and KO2 of 37.2 ± 6.1 μM. The kcat exhibited a pH optimum at 7.5 while kcat/Klysine peaked at 7.0 and remained constant to pH 8.5. Alternative electron acceptors could not effectively substitute for O2 in the reaction. A mechanism for the reductive half reaction of LodA is proposed that is consistent with the ping-pong kinetics.  相似文献   

11.
The genome of the nematode Caenorhabditis elegans contains several genes that appear to encode proteins similar to CTP:phosphocholine cytidylyltransferase (CCT). We have isolated a 1044-nucleotide cDNA clone from a C. elegans cDNA library that encodes the 347-amino acid version of CCT that is most similar to previously-identified CCTs. Native and His-tagged forms were expressed and purified using a baculovirus expression system. The enzyme was maximally activated by 5 μM phosphatidylcholine:oleate (50:50) vesicles with a kcat value in the presence of lipid 37-fold greater than the kcat value in the absence of lipid. To localize the region of C. elegans CCT critical for lipid activation, a series of C-terminal truncation mutants was analyzed. CCT truncated after amino acids 225 or 245 was quite active in the absence of lipids and not further activated in the presence of lipids, supporting the concept that the lipid-activation segment is inhibitory to catalysis in the absence of lipids. CCT truncated after amino acids 266, 281, or 319 was activated by lipid similar to wild-type enzyme. Kinetic analysis in the absence of lipid revealed the lipid-independent CCT truncated after amino acid 245 to have a kcat value 15-fold greater than either full-length CCT or CCT truncated after amino acid 266. We conclude that elements critical for activation of C. elegans CCT by lipids are contained within amino acids 246–266, that this region is inhibitory in the absence of lipids, and that the inhibition is relieved by the association of the enzyme with lipid.  相似文献   

12.
Cystathionine γ-lyase (CGL) catalyzes the hydrolysis of l-cystathionine (l-Cth), producing l-cysteine (l-Cys), α-ketobutyrate and ammonia, in the second step of the reverse transsulfuration pathway, which converts l-homocysteine (l-Hcys) to l-Cys. Site-directed variants substituting residues E48 and E333 with alanine, aspartate and glutamine were characterized to probe the roles of these acidic residues, conserved in fungal and mammalian CGL sequences, in the active-site of CGL from Saccharomyces cerevisiae (yCGL). The pH optimum of variants containing the alanine or glutamine substitutions of E333 is increased by 0.4–1.2 pH units, likely due to repositioning of the cofactor and modification of the pKa of the pyridinium nitrogen. The pH profile of yCGL-E48A/E333A resembles that of Escherichia coli cystathionine β-lyase. The effect of substituting E48, E333 or both residues is the 1.3–3, 26–58 and 124–568-fold reduction, respectively, of the catalytic efficiency of l-Cth hydrolysis. The Kml-Cth of E333 substitution variants is increased ~ 17-fold, while Kml-OAS is within 2.5-fold of the wild-type enzyme, indicating that residue E333 interacts with the distal amine moiety of l-Cth, which is not present in the alternative substrate O-acetyl-l-serine. The catalytic efficiency of yCGL for α,γ-elimination of O-succinyl-l-homoserine (kcat/Kml-OSHS = 7 ± 2), which possesses a distal carboxylate, but lacks an amino group, is 300-fold lower than that of the physiological l-Cth substrate (kcat/Kml-Cth = 2100 ± 100) and 260-fold higher than that of l-Hcys (kcat/Kml-Hcys = 0.027 ± 0.005), which lacks both distal polar moieties. The results of this study suggest that the glutamate residue at position 333 is a determinant of specificity.  相似文献   

13.
The prevalence of paralogous enzymes implies that novel catalytic functions can evolve on preexisting protein scaffolds. The weak secondary activities of proteins, which reflect catalytic promiscuity and substrate ambiguity, are plausible starting points for this evolutionary process. In this study, we observed the emergence of a new enzyme from the ASKA (A Complete Set of E. coli K-12 ORF Archive) collection of Escherichia coli open reading frames. The overexpression of (His)6-tagged glutamine phosphoribosylpyrophosphate amidotransferase (PurF) unexpectedly rescued a ΔtrpF E. coli strain from starvation on minimal media. The wild-type PurF and TrpF enzymes are unrelated in sequence, tertiary structure and catalytic mechanism. The promiscuous phosphoribosylanthranilate isomerase activity of the ASKA PurF variant apparently stems from a preexisting affinity for phosphoribosylated substrates. The relative fitness of the (His)6-PurF/ΔtrpF strain was improved 4.8-fold to nearly wild-type levels by random mutagenesis of purF and genetic selection. The evolved and ancestral PurF proteins were purified and reacted with phosphoribosylanthranilate in vitro. The best evolvant (kcat/KM = 0.3 s− 1 M− 1) was ∼ 25-fold more efficient than its ancestor but > 107-fold less efficient than the wild-type phosphoribosylanthranilate isomerase. These observations demonstrate in quantitative terms that the weak secondary activities of promiscuous enzymes can dramatically improve the fitness of contemporary organisms.  相似文献   

14.
A bacterial arylmalonate decarboxylase (AMDase) catalyzes asymmetric decarboxylation of unnatural arylmalonates to produce optically pure (R)-arylcarboxylates without the addition of cofactors. Previously, we designed an AMDase variant G74C/C188S that displays totally inverted enantioselectivity. However, the variant showed a 20,000-fold reduction in activity compared with the wild-type AMDase. Further studies have demonstrated that iterative saturation mutagenesis targeting the active site residues in a hydrophobic pocket of G74C/C188S leads to considerable improvement in activity where all positive variants harbor only hydrophobic substitutions. In this study, simultaneous saturation mutagenesis with a restricted set of amino acids at each position was applied to further heighten the activity of the (S)-selective AMDase variant toward α-methyl-α-phenylmalonate. The best variant (V43I/G74C/A125P/V156L/M159L/C188G) showed 9,500-fold greater catalytic efficiency kcat/Km than that of G74C/C188S. Notably, a high level of decarboxylation of α-(4-isobutylphenyl)-α-methylmalonate by the sextuple variant produced optically pure (S)-ibuprofen, an analgesic compound which showed 2.5-fold greater activity than the (R)-selective wild-type AMDase.  相似文献   

15.

Background

The commercially important glycoside hydrolase family 3 (GH3) β-glucosidases from Aspergillus niger are anomeric-configuration-retaining enzymes that operate through the canonical double-displacement glycosidase mechanism. Whereas the catalytic nucleophile is readily identified across all GH3 members by sequence alignments, the acid/base catalyst in this family is phylogenetically variable and less readily divined.

Methods

In this report, we employed three-dimensional structure homology modeling and detailed kinetic analysis of site-directed mutants to identify the catalytic acid/base of a GH3 β-glucosidase from A. niger ASKU28.

Results

In comparison to the wild-type enzyme and other mutants, the E490A variant exhibited greatly reduced kcat and kcat/Km values toward the natural substrate cellobiose (67,000- and 61,000-fold, respectively). Correspondingly smaller kinetic effects were observed for artificial chromogenic substrates p-nitrophenyl β-d-glucoside and 2,4-dinitrophenyl β-d-glucoside, the aglycone leaving groups of which are less dependent on acid catalysis, although changes in the rate-determining catalytic step were revealed for both. pH-rate profile analyses also implicated E490 as the general acid/base catalyst. Addition of azide as an exogenous nucleophile partially rescued the activity of the E490A variant with the aryl β-glucosides and yielded β-glucosyl azide as a product.

Conclusions and general significance

These results strongly support the assignment of E490 as the acid/base catalyst in a β-glucosidase from A. niger ASKU28, and provide crucial experimental support for the bioinformatic identification of the homologous residue in a range of related GH3 subfamily members.  相似文献   

16.
Rhodanese is a component of the mitochondrial H2S oxidation pathway. Rhodanese catalyzes the transfer of sulfane sulfur from glutathione persulfide (GSSH) to sulfite generating thiosulfate and from thiosulfate to cyanide generating thiocyanate. Two polymorphic variations have been identified in the rhodanese coding sequence in the French Caucasian population. The first, 306A→C, has an allelic frequency of 1% and results in an E102D substitution in the encoded protein. The second polymorphism, 853C→G, has an allelic frequency of 5% and leads to a P285A substitution. In this study, we have examined differences in the stability between wild-type rhodanese and the E102D and P285A variants and in the kinetics of the sulfur transfer reactions. The Asp-102 and Ala-285 variants are more stable than wild-type rhodanese and exhibit kcat/Km,CN values that are 17- and 1.6-fold higher, respectively. All three rhodanese forms preferentially catalyze sulfur transfer from GSSH to sulfite, generating thiosulfate and glutathione. The kcat/Km,sulfite values for the variants in the sulfur transfer reaction from GSSH to sulfite were 1.6- (Asp-102) and 4-fold (Ala-285) lower than for wild-type rhodanese, whereas the kcat/Km,GSSH values were similar for all three enzymes. Thiosulfate-dependent H2S production in murine liver lysate is low, consistent with a role for rhodanese in sulfide oxidation. Our studies show that polymorphic variations that are distant from the active site differentially modulate the sulfurtransferase activity of human rhodanese to cyanide versus sulfite and might be important in differences in susceptibility to diseases where rhodanese dysfunction has been implicated, e.g. inflammatory bowel diseases.  相似文献   

17.
The roles of local interactions in the laboratory evolution of a highly active, computationally designed retroaldolase (RA) are examined. Partial Order Optimum Likelihood (POOL) is used to identify catalytically important amino acid interactions in several RA95 enzyme variants. The series RA95.5, RA95.5–5, RA95.5–8, and RA95.5–8F, representing progress along an evolutionary trajectory with increasing activity, is examined. Computed measures of coupling between charged states of residues show that, as evolution proceeds and higher activities are achieved, electrostatic coupling between the biochemically active amino acids and other residues is increased. In silico residue scanning suggests multiple coupling partners for the catalytic lysine K83. The effects of two predicted partners, Y51 and E85, are tested using site‐directed mutagenesis and kinetic analysis of the variants Y51F and E85Q. The Y51F variants show decreases in k cat relative to wild type, with the greatest losses observed for the more evolved constructs; they also exhibit significant decreases in k cat/K M across the series. Only modest decreases in k cat/K M are observed for the E85Q variants with little effect on k cat. Computed metrics of the degree of coupling between protonation states rise significantly as evolution proceeds and catalytic turnover rate increases. Specifically, the charge state of the catalytic lysine K83 becomes more strongly coupled to those of other amino acids as the enzyme evolves to a better catalyst.  相似文献   

18.
Fluorescently labeled derivatives of a xyloglucan (XG) nonasaccharide Glc4Xyl3Gal2 (XLLG) were used as glycosyl acceptors in assays of xyloglucan endotransglycosylase (XET) from germinated nasturtium (Tropaeolum majus) seeds. We have investigated how the type of the oligosaccharide label influences the kinetic parameters of the reaction. The fluorescent probes used to label XLLG were anthranilic acid (AA), 8-aminonaphtalene-1,3,6-trisulfonic acid (ANTS), fluorescein isothiocyanate (FITC), and sulforhodamine (SR), respectively. The obtained data were compared with those of the reactions where aldose and/or alditol forms of tritium-labeled xyloglucan-derived nonasaccharide served as the respective acceptors. Modification at C-1 of the reducing-end glucose in XLLG by substitution with the fluorophore markedly affected the kinetic parameters of the reaction. The Michaelis constants Km for individual acceptors increased in the order [1-3H]XLLG < XLLG-SR < [1-3H]XLLGol < XLLG-FITC < XLLG-ANTS < XLLG-AA, while the turnover numbers characterized by kcat decreased in the order XLLG-FITC > XLLG-SR > XLLG-ANTS > [1-3H]XLLGol > [1-3H]XLLG > XLLG-AA. Catalytic efficiency (expressed as kcat/Km) with XLLG labeled with SR or FITC was 15 and 28 times, respectively, higher than with the tritium-labeled natural substrate [1-3H]XLLG. Comparison of the kinetic parameters found with acceptors labeled with different types of labels enables to select the most effective substrates for the high-throughput assays of XET.  相似文献   

19.
The deacetoxycephalosporin C synthase (DAOCS) from Streptomyces clavuligerus was engineered with the aim of enhancing the conversion of penicillin G into phenylacetyl-7-aminodeacetoxycephalosporanic acid, a precursor of 7-aminodeacetoxycephalosporanic acid, for industrial application. A single round of random mutagenesis followed by the screening of 5,500 clones identified three mutants, G79E, V275I, and C281Y, that showed a two- to sixfold increase in the kcat/Km ratio compared to the wild-type enzyme. Site-directed mutagenesis to modify residues surrounding the substrate resulted in three mutants, N304K, I305L, and I305M, with 6- to 14-fold-increased kcat/Km values. When mutants containing all possible combinations of these six sites were generated to optimize the ring expansion activity for penicillin G, the double mutant, YS67 (V275I, I305M), showed a significant 32-fold increase in the kcat/Km ratio and a 5-fold increase in relative activity for penicillin G, while the triple mutant, YS81 (V275I, C281Y, I305M), showed an even greater 13-fold increase in relative activity toward penicillin G. Our results demonstrate that this is a robust approach to the modification of DAOCS for an optimized DAOCS-penicillin G reaction.  相似文献   

20.
Enzyme function depends on specific conformational motions. We show that the temperature dependence of enzyme kinetic parameters can provide insight into these functionally relevant motions. While investigating the catalytic properties of IPMDH from Escherichia coli, we found that its catalytic efficiency (kcat/KM,IPM) for the substrate IPM has an unusual temperature dependence, showing a local minimum at ∼35°C. In search of an explanation, we measured the individual constants kcat and KM,IPM as a function of temperature, and found that the van 't Hoff plot of KM,IPM shows sigmoid-like transition in the 20-40°C temperature range. By means of various measurements including hydrogen-deuterium exchange and fluorescence resonance energy transfer, we showed that the conformational fluctuations, including hinge-bending domain motions increase more steeply with temperatures >30°C. The thermodynamic parameters of ligand binding determined by isothermal titration calorimetry as a function of temperature were found to be strongly correlated to the conformational fluctuations of the enzyme. Because the binding of IPM is associated with a hinge-bending domain closure, the more intense hinge-bending fluctuations at higher temperatures increasingly interfere with IPM binding, thereby abruptly increasing its dissociation constant and leading to the observed unusual temperature dependence of the catalytic efficiency.  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