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.
The kinetic characterization of horseradish peroxidase (HRPC) substrates is difficult because the reaction products are free radicals. The application of a spectrophotometrical method, which is based on determining the time necessary for a given quantity of L-ascorbic acid to be consumed (lag period) during its reaction with the free radicals generated by the enzyme acting on the reducing substrate, makes it possible to obtain the initial steady-state rates (v0). From the kinetic study of a series of derivates of phenol and aniline, the following parameters were determined for the first time: the global catalytic constant (kcat), the Michaelis constant of HRPC for H2O2 in the presence of each reducing substrate (K(M)H2O2), the Michaelis constant of HRPC for the reducing substrate (KMS), the binding constant of the reducing substrate with HRPC compound II (k5) and the rate constant of substrate oxidation by HRPC compound II (k6). The values obtained are disccussed.  相似文献   

2.
Transient kinetic analysis of biphasic, single turnover data for the reaction of 2,2'-azino-bis[3-ethylbenzthiazoline-6-sulfonic acid] (ABTS) with horseradish peroxidase (HRPC) compound II demonstrated preequilibrium binding of ABTS (k(+5) = 7.82 x 10(4) M(-)(1) s(-)(1)) prior to rate-limiting electron transfer (k(+6) = 42.1 s(-)(1)). These data were obtained using a stopped-flow method, which included ascorbate in the reaction medium to maintain a low steady-state concentration of ABTS (pseudo-first-order conditions) and to minimize absorbance changes in the Soret region due to the accumulation of ABTS cation radicals. A steady-state kinetic analysis of the reaction confirmed that the reduction of HRPC compound II by this substrate is rate-limiting in the complete peroxidase cycle. The reaction of HRPC with o-diphenols has been investigated using a chronometric method that also included ascorbate in the assay medium to minimize the effects of nonenzymic reactions involving phenol-derived radical products. This enabled the initial rates of o-diphenol oxidation at different hydrogen peroxide and o-diphenol concentrations to be determined from the lag period induced by the presence of ascorbate. The kinetic analysis resolved the reaction of HRPC compound II with o-diphenols into two steps, initial formation of an enzyme-substrate complex followed by electron transfer from the substrate to the heme. With o-diphenols that are rapidly oxidized, the heterolytic cleavage of the O-O bond of the heme-bound hydrogen peroxide (k(+2) = 2.17 x 10(3) s(-)(1)) is rate-limiting. The size and hydrophobicity of the o-diphenol substrates are correlated with their rate of binding to HRPC, while the electron density at the C-4 hydroxyl group predominantly influences the rate of electron transfer to the heme.  相似文献   

3.
A direct mass spectrometric approach was used for the determination of steady-state kinetic parameters, the turnover number (k(cat)), the Michaelis constant (K(M)), and the specificity constant (k(cat)/K(M)) for an enzyme-catalyzed hydrolysis of xylooligosaccharides. Electrospray ionization mass spectrometry was performed to observe product distributions and to determine k(cat), K(M), and k(cat)/K(M) values for Trichoderma reesei endo-1,4-beta-xylanase II (TRX II) with xylohexaose (Xyl(6)), xylopentaose (Xyl(5)), xylotetraose (Xyl(4)), and xylotriose (Xyl(3)) as substrates. The determined k(cat)/K(M) values (0.93, 0.37, 0.027, and 0.00015 microM(-1) s(-1), respectively) indicated that Xyl(6) was the most preferred substrate of TRX II. In addition, the obtained K(M) value for Xyl(5) (136 microM) was roughly twice as high as that for Xyl(6) (73 microM), suggesting that at least six putative subsites contribute to the substrate binding in the active site of TRX II. Previous mass spectrometric assays for enzyme kinetics have been used mostly in the case of reactions that result in a transfer of acidic groups (e.g., phosphate) into neutral oligosaccharides giving rise to negatively charged products. Here we demonstrate that such analysis is also feasible in the case of neutral underivatized oligosaccharides. Implications of the results for the catalytic mechanism of TRX II in particular are discussed.  相似文献   

4.
Wu J  Kinoshita K  Khosla C  Cane DE 《Biochemistry》2004,43(51):16301-16310
The beta-ketoacyl-acyl carrier protein synthase (KS) domain of the modular 6-deoxyerythronolide B synthase (DEBS) catalyzes the fundamental chain building reaction of polyketide biosynthesis. The KS-catalyzed reaction involves two discrete steps consisting of formation of an acyl-enzyme intermediate generated from the incoming acylthioester substrate and an active site cysteine residue, and the conversion of this intermediate to the beta-ketoacyl-acyl carrier protein product by a decarboxylative condensation with a paired methylmalonyl-SACP. We have determined the rate constants for the individual biochemical steps by a combination of protein acylation and transthioesterification experiments. The first-order rate constant (k(2)) for formation of the acyl-enzyme intermediate from [1-(14)C]-(2S,3R)-2-methyl-3-hydroxypentanoyl-SNAC (2) and recombinant DEBS module 2 is 5.8 +/- 2.6 min(-)(1), with a dissociation constant (K(S)) of 3.5 +/- 2.8 mM. The acyl-enzyme adduct was formed at a near-stoichiometric ratio of approximately 0.8:1. Transthioesterification between unlabeled diketide-SNAC 2 and N-[1-(14)C-acetyl]cysteamine gave a k(exch) of 0.15 +/- 0.06 min(-)(1), with a K(m) for HSNAC of 5.7 +/- 4.9 mM and a K(m) for 2 of 5.3 +/- 0.9 mM. Under the conditions that were used, k(exch) was equal to k(-)(2), the first-order rate constant for reversal of the acyl-enzyme-forming reaction. Since the rate of the decarboxylative condensation is much greater that the rate of reversion to the starting material (k(3) > k(-)(2)), formation of the acyl-enzyme adduct is effectively irreversible, thereby establishing that the observed value of the specificity constant (k(cat)/K(m)) is solely a reflection of the intrinsic substrate specificity of the KS-catalyzed acyl-enzyme-forming reaction. These findings were also extended to a panel of diketide- and triketide-SNAC analogues, revealing that some substrate analogues that are not converted to product by DEBS module 2 form dead-end acyl-enzyme intermediates.  相似文献   

5.
A solvent deuterium isotope effect on the catalytic affinity (K(m)) and rate constant (k(cat)) of tyrosinase in its action on 4-tert-butylcatechol (TBC) was observed. Both parameters decreased as the molar fraction of deuterated water in the medium increased, while the k(cat)/K(m) ratio remained constant. In a proton inventory study, the representation of k(cat)(f(n))/k(cat)(f(0)) and K(m)(f(n))/K(m)(f(0)) vs. n (atom fractions of deuterium) was linear, indicating that, of the four protons transferred from the two molecules of substrate and which are oxidized in one turnover, only one is responsible for the isotope effects. The fractionation factor of 0.64+/-0.02 contributed to identifying the possible proton acceptor. Possible mechanistic implications are discussed.  相似文献   

6.
The determination of kinetic parameters of chitinases using natural substrates is difficult due to low K(m) values, which require the use of low substrate concentrations that are hard to measure. Using the natural substrate (GlcNAc)(4), we have developed an assay for the determination of k(cat) and K(m)values of chitinases. Product concentrations as low as 0.5 microM were detected using normal-phase high-performance liquid chromatography (HPLC) with an amide 80 column (0.20 x 25 cm) using spectrophotometric detection at 210 nm. By means of this assay, k(cat) and K(m)values for chitinases A (ChiA) and B (ChiB) of Serratia marcescens were found to be 33+/-1s(-1) and 9+/-1 microM and 28+/-2s(-1) and 4+/-2 microM, respectively. For ChiB, these values were compared to those found with commonly used substrates where the leaving group is a (nonnatural) chromophore, revealing considerable differences. For example, assays with 4-methylumbelliferyl-(GlcNAc)(2) yielded a k(cat) value of 18+/-2s(-1) and a K(m) value of 30+/-6 microM. For two ChiB mutants containing a Trp --> Ala mutation in the +1 or +2 subsites, the natural substrate and the 4-methylumbelliferyl-(GlcNAc)(2) assays yielded rather similar K(m) values (5-fold difference at most) but showed dramatic differences in k(cat) values (up to 90-fold). These results illustrate the risk of using artificial substrates for characterization of chitinases and, thus, show that the new HPLC-based assay is a valuable tool for future chitinase research.  相似文献   

7.
Vocadlo DJ  Wicki J  Rupitz K  Withers SG 《Biochemistry》2002,41(31):9727-9735
The catalytic mechanism of Thermoanaerobacterium saccharolyticum beta-xylosidase (XynB) from family 39 of glycoside hydrolases has been subjected to a detailed kinetic investigation using a range of substrates. The enzyme exhibits a bell-shaped pH dependence of k(cat)/K(m), reflecting apparent pK(a) values of 4.1 and 6.8. The k(cat) and k(cat)/K(m) values for a series of aryl xylosides have been measured and used to construct two Br?nsted plots. The plot of log(k(cat)/K(m)) against the pK(a) of the leaving group reveals a significant correlation (beta(lg) = -0.97, r(2) = 0.94, n = 8), indicating that fission of the glycosidic bond is significantly advanced in the transition state leading to the formation of the xylosyl-enzyme intermediate. The large negative value of the slope indicates that there is relatively little proton donation to the glycosidic oxygen in the transition state. A biphasic, concave-downward plot of log(k(cat)) against pK(a) provides good evidence for a two-step double-displacement mechanism involving a glycosyl-enzyme intermediate. For activated leaving groups (pK(a) < 9), the breakdown of the xylosyl-enzyme intermediate is the rate-determining step, as indicated by the absence of any effect of the pK(a) of the leaving group on log(k(cat)) (beta(lg) approximately 0). However, a strong dependence of the first-order rate constant on the pK(a) value of relatively poor leaving groups (pK(a) > 9) suggests that the xylosylation step is rate-determining for these substrates. Support for the dexylosylation chemical step being rate-determining for activated substrates comes from nucleophilic competition experiments in which addition of dithiothreitol results in an increase in turnover rates. Normal secondary alpha-deuterium kinetic isotope effects ((alpha-D)(V) or (alpha-D)(V/K) = 1.08-1.10) for three different substrates of widely varying pK(a) value (5.15-9.95) have been measured and these reveal that the transition states leading to the formation and breakdown of the intermediate are similar and both steps involve rehybridization of C1 from sp(3) to sp(2). These results are consistent only with "exploded" transition states, in which the saccharide moiety bears considerable positive charge, and the intermediate is a covalent acylal-ester where C1 is sp(3) hybridized.  相似文献   

8.
The complex reaction mechanism of tyrosinase involves three enzymatic forms, two overlapping catalytic cycles and a dead-end complex. Analytical expressions for the catalytic and Michaelis constants of tyrosinase towards phenols and oxygen were derived for both, monophenolase and diphenolase activities of the enzyme. Thus, the Michaelis constants of tyrosinase towards the oxygen (K(mO(2))) are related with the respective catalytic constants for monphenols (k(M)(cat)) and o-diphenols (k(D)(cat)), as well as with the rate constant, k(+8). We recently determined the experimental value of the rate constant for the binding of oxygen to deoxytyrosinase (k(+8)) by stopped-flow assays. In this paper, we calculate theoretical values of K(mO(2)) from the experimental values of catalytic constants and k(+8) towards several monophenols and o-diphenols. The reliability and the significance of the values of K(mO(2)) are discussed.  相似文献   

9.
The reaction of mushroom (Agaricus bisporus) tyrosinase with dioxygen in the presence of several o-diphenolic substrates has been studied by steady-state and transient-phase kinetics in order to elucidate the rate-limiting step and to provide new insights into the mechanism of oxidation of these substrates. A kinetic analysis has allowed for the first time the determination of individual rate constants for several of the partial reactions that comprise the catalytic cycle. Mushroom tyrosinase rapidly reacts with dioxygen with a second-order rate constant k(+8) = 2.3 x 10(7) M(-)(1) s(-)(1), which is similar to that reported for hemocyanins [(1.3 x 10(6))-(5.7 x 10(7)) M(-)(1) s(-)(1)]. Deoxytyrosinase binds dioxygen reversibly at the binuclear Cu(I) site with a dissociation constant K(D)(O)()2 = 46.6 microM, which is similar to the value (K(D)(O)()2 = 90 microM) reported for the binding of dioxygen to Octopus vulgaris deoxyhemocyanin [Salvato et al. (1998) Biochemistry 37, 14065-14077]. Transient and steady-state kinetics showed that o-diphenols such as 4-tert-butylcatechol react significantly faster with mettyrosinase (k(+2) = 9.02 x 10(6) M(-)(1) s(-)(1)) than with oxytyrosinase (k(+6) = 5.4 x 10(5) M(-)(1) s(-)(1)). This difference is interpreted in terms of differential steric and polar effects that modulate the access of o-diphenols to the active site for these two forms of the enzyme. The values of k(cat) for several o-diphenols are also consistent with steric and polar factors controlling the mobility, orientation, and thence the reactivity of substrates at the active site of tyrosinase.  相似文献   

10.
D-Lactate dehydrogenase (D-LDH) from Pediococcus pentosaceus ATCC 25745 was found to produce D-3-phenyllactic acid from phenylpyruvate. The optimum pH and temperature for enzyme activity were pH 5.5 and 45 °C. The Michaelis-Menten constant (K(m)), turnover number (k(cat)), and catalytic efficiency (k(cat)/K(m)) values for the substrate phenylpyruvate were estimated to be 1.73 mmol/L, 173 s(-1), and 100 (mmol/L)(-1) s(-1) respectively.  相似文献   

11.
Three genes encoding heme hydroperoxidases (katA, katB, and katC) have been identified in the soil bacterium Sinorhizobium meliloti. The recombinant proteins were overexpressed in Escherichia coli and purified in order to achieve a spectral and kinetic characterization. The three proteins contain heme b with high-spin Fe(III). KatB is an acidic bifunctional homodimeric catalase-peroxidase exhibiting both catalase (k(cat) = 2400 s(-1)) and peroxidase activity and having a high affinity for hydrogen peroxide (apparent K(M) = 1.6 mM). KatA and KatC are acidic monofunctional homotetrameric catalases. Although different in size (KatA is a small subunit catalase while KatC is a large subunit catalase) both enzymes exhibit the same heme type and a similar affinity for H(2)O(2) (apparent K(M) values of 160 and 150 mM). However, the turnover rate of KatA (k(cat) = 279000 s(-1)) exceeds that of KatC (k(cat) = 3100 s(-1)) significantly. The kinetic parameters are in good agreement with the physiological role of these heme proteins. KatB is the housekeeping hydroperoxidase exhibiting the highest affinity for hydrogen peroxide, while KatA has the lowest H(2)O(2) affinity but the highest k(cat)/K(M) value (1.75 x 10(6) M(-1) s(-1)), in agreement with the hydrogen peroxide inducibility of the encoding gene. Moreover, the lower catalytic efficiency of KatC (2.1 x 10(4) M(-1) s(-1)) appears to be enough for growing in the stationary phase and/or under heat or salt stress (conditions that are known to favor katC expression).  相似文献   

12.
The apparent catalytic constant (k(cat)) of artichoke (Cynara scolymus L.) peroxidase (AKPC) with 2,2'-azino-bis(3-ethylbenzthiazoline-6-sulphonic acid) (ABTS) increased 130-fold in the presence of calcium ions (Ca2+) but the affinity (K(m)) of the enzyme for ABTS was 500 times lower than for Ca2+-free AKPC. AKPC is known to exhibit an equilibrium between 6-aquo hexa-coordinate and penta-coordinate forms of the haem iron that is modulated by Ca2+ and affects compound I formation. Measurements of the Ca2+ dissociation constant (K(D)) were complicated by the water-association/dissociation equilibrium yielding a global value more than 1000 times too high. The value for the Ca2+ binding step alone has now been determined to be K(D) approximately 10 nM. AKPC-Ca2+ was more resistant to inactivation by hydrogen peroxide (H(2)O(2)) and exhibited increased catalase activity. An analysis of the complex H(2)O(2) concentration dependent kinetics of Ca2+-free AKPC is presented.  相似文献   

13.
We report here on the stereospecificity observed in the action of horseradish peroxidase (HRPC) on monophenol and diphenol substrates. Several enantiomers of monophenols and o-diphenols were assayed: L-tyrosinol, D-tyrosinol, L-tyrosine, DL-tyrosine, D-tyrosine, L-dopa, DL-dopa, D-dopa, L-alpha-methyldopa, DL-alpha-methyldopa, DL-adrenaline, D-adrenaline, L-isoproterenol, DL-isoproterenol and D-isoproterenol. The electronic density at the carbon atoms in the C-1 and C-2 positions of the benzene ring were determined by NMR assays (delta1 and delta2). This value is related to the nucleophilic power of the oxygen atom of the hydroxyl groups and to its oxidation-reduction capacity. The spatial orientation of the ring substituents resulted in lower Km values for L- than for D-isomers. The kcat values for substrates capable of saturating the enzyme were lower for D- than for L-isomers, although both have the same delta1 and delta2 NMR values for carbons C-1 and C-2, and therefore the same oxidation-reduction potential. In the case of substrates that cannot saturate the enzyme, the values of the binding constant for compound II (an intermediate in the catalytic cycle) followed the order: L-isomer>DL-isomer>D-isomer. Therefore, horseradish peroxidase showed stereospecificity in its affinity toward its substrates (K m) and in their transformation reaction rates (k cat).  相似文献   

14.
Bar-Shalom A  Moore MJ 《Biochemistry》2000,39(33):10207-10218
Group II introns self-splice via a two-step mechanism: cleavage at the 5' splice site followed by exon ligation at the 3' splice site. The second step has been difficult to study in vitro because it is generally faster than the first. Herein we describe development and partial kinetic characterization of a novel assay for studying the second step in isolation. In this system, a truncated linear intron (nucleotides 1-881) mediates exon ligation between two oligonucleotide substrates: a 19 nt 5' exon and a 3' substrate consisting of the last 6 nucleotides of the intron plus a 6 nucleotide 3' exon. We found that neither the exact structure of domain 6 nor the identity of nucleotides flanking the 3' splice site is critical for accurate 3' splice site choice by the ai5gamma group II intron. The multiple turnover k(cat) (0.14 min(-)(1)) is slower than the single turnover k(obs) (0.6-0.7 min(-)(1)), consistent with rate-limiting product release under steady-state conditions. Decreased single turnover rates at lower pHs were more consistent with loss of catalytic activity than with rate-limiting chemistry. Binding of the 3' substrate (K(m) = 2.6 microM) could be improved by changing a long-range A:U base pair involving the last intronic nucleotide (the gamma-gamma' interaction) to G:C (K(m(3)(')(substrate)) = 1 microM).  相似文献   

15.
Catechol 2,3-dioxygenase [EC 1.13.11.2] from Pseudomonas putida mt-2 (Mpc) catalyzes the extradiol cleavage of catechol to produce 2-hydroxymuconate semialdehyde. The K(m) values for the catecholic substrate (K(mA)) and O(2) (K(mO2)), and catalytic constants (k(cat)) were kinetically determined for eight C3/C4-substituted catechols at 25 degrees C and pH 6.5 or 7.5. The first pK(a) values (pK(1)) were determined for eleven catechols (pK(1) = 7.26-9.47), correlated with Hammett substituent constants, and electron-withdrawing substituents significantly stabilized the monoanionic species of free catechols. Mpc preferred catechols with non-ionic substituents at the C3 or C4 position. 3-Phenylcatechol, a biphenyl, was cleaved, while 4-tert-butylcatechol was not. The logarithm of k(cat)/K(mA) (substrate specificity constant) exhibited a good linear correlation with pK(1), with the exception of those for 4-halocatechols. The logarithm of k(cat)/K(mO2) showed a good linear correlation with pK(1), with the exception of that of 3-phenylcatechol. These results demonstrate that catechol binding to the Mpc active site, the following O(2) binding, and the activation of the bound O(2) are all sensitive to electronic effects of the substituents. However, k(cat) did not correlate significantly with pK(1). The present study distinguishes clearly between the electronic and the steric effects of catecholic substrates in the reactivity of Mpc, and provides important insight into the mechanistic basis for a vast range of substrate specificities of extradiol dioxygenases.  相似文献   

16.
Hardy LW  Kirsch JF 《Biochemistry》1984,23(6):1282-1287
The solvent kinetic isotope effects (SKIE's) on k(cat) (D(V)) and on k(cat/Km[D(V/K)] were determined for the Bacillus cereus beta-lactamase I catalyzed hydrolysis of five substrates that have values of k(cat)/K(m) varying over the range (0.014-46.3) X 10(6)M(-1) s(-1) and of k(cat) between 0.5 and 2019 s(-1). The variation of D(V/K) was only from 1.06 to 1.25 among these compounds and that in D(V) was from 1.50 to 2.16. These results require that Dk(1), the SKIE on the enzyme-substrate association rate constant, and D(k-1/k2), that on the partition ratio of the ES complex, both be near 1. The larger SKIE observed on D(V) requires that an exchangeable proton be in flight for either or both the acylation and the deacylation reaction. The pH dependence of the values k(cat)/K(m) for three substrates shows identical pK(a)s of 5.5. and 8.4. This identity combined with the fact that only one of these three substrates is kinetically "sticky" proves that the substrates can combine productively with only one protonic form of the enzyme. There is considerable substrate variation in the pK(a) values of k(cat) observed vs. pH profiles; the inflection points for all substrates studied are at pH values more extreme than are observed in the pH profiles for k(cat)/K(m).  相似文献   

17.
Studies of the proton-transfer reaction by Pseudomonas testosteroni 3-oxo steroid Delta(4)-Delta(5)-isomerase with Delta(5(6))- and Delta(5(10))-steroid substrates demonstrate the importance of the position of the double bond for the efficiency of the isomerization process. Thus 3-oxo-Delta(5(6))-substrates have markedly high k(cat.) values, whereas those of 3-oxo-Delta(5(10))-substrates are very low and their apparent K(m) values approach equilibrium dissociation constants. The first step in the isomerization process is: [Formula: see text] which is governed by the k(-1)/k(+1) ratio and is shown to be very similar for the two classes of substrates (3-oxo-Delta(5(6))- and -Delta(5(10))-steroids). They therefore differ in the steps distal to the initial formation of the Michaelis-Menten complex. The use of the deuterated androst-5(6)-ene-3,17-dione substrate enabled us to calculate individual rate constants k(+1) and k(-1) as well as to determine the apparent rate-limiting step in the isomerization process. With the deuterated oestr-5(10)-ene-3,17-dione substrate, no significant isotope effect was observed suggesting that a different rate-limiting step may be operative in this isomerization process. Data are presented that indicate that under optimal concentrations of the efficient androst-5(6)-ene-3,17-dione substrate, the forward reaction for ES complex formation (as defined by k(+1)) is limited only by diffusion and the apparent K(m) does not approach the equilibrium constant, suggesting that the evolution of this enzyme has proceeded close to ;catalytic perfection'.  相似文献   

18.
Tyrosine 343 in human sulfite oxidase (SO) is conserved in all SOs sequenced to date. Intramolecular electron transfer (IET) rates between reduced heme (Fe(II)) and oxidized molybdenum (Mo(VI)) in the recombinant wild-type and Y343F human SO were measured for the first time by flash photolysis. The IET rate in wild-type human SO at pH 7.4 is about 37% of that in chicken SO with a similar decrease in k(cat). Steady-state kinetic analysis of the Y343F mutant showed an increase in K(m)(sulfite) and a decrease in k(cat) resulting in a 23-fold attenuation in the specificity constant k(cat)/K(m)(sulfite) at the optimum pH value of 8.25. This indicates that Tyr-343 is involved in the binding of the substrate and catalysis within the molybdenum active site. Furthermore, the IET rate constant in the mutant at pH 6.0 is only about one-tenth that of the wild-type enzyme, suggesting that the OH group of Tyr-343 is vital for efficient IET in SO. The pH dependences of IET rate constants in the wild-type and mutant SO are consistent with the previously proposed coupled electron-proton transfer mechanism.  相似文献   

19.
The Staphylococcus aureus transpeptidase Sortase A (SrtA) anchors virulence and colonization-associated surface proteins to the cell wall. SrtA selectively recognizes a C-terminal LPXTG motif, whereas the related transpeptidase Sortase B (SrtB) recognizes a C-terminal NPQTN motif. In both enzymes, cleavage occurs after the conserved threonine, followed by amide bond formation between threonine and the pentaglycine cross-bridge of cell wall peptidoglycan. Genetic and biochemical studies strongly suggest that SrtA and SrtB exhibit exquisite specificity for their recognition motifs. To better understand the origins of substrate specificity within these two isoforms, we used sequence and structural analysis to predict residues and domains likely to be involved in conferring substrate specificity. Mutational analyses and domain swapping experiments were conducted to test their function in substrate recognition and specificity. Marked changes in the specificity profile of SrtA were obtained by replacing the beta6/beta7 loop in SrtA with the corresponding domain from SrtB. The chimeric beta6/beta7 loop swap enzyme (SrtLS) conferred the ability to acylate NPQTN-containing substrates, with a k(cat)/K(m)(app) of 0.0062 +/- 0.003 m(-1) s(-1). This enzyme was unable to perform the transpeptidation stage of the reaction, suggesting that additional domains are required for transpeptidation to occur. The overall catalytic specificity profile (k(cat)/K(m)(app)(NPQTN)/k(cat)/K(m)(app)(LPETG)) of SrtLS was altered 700,000-fold from SrtA. These results indicate that the beta6/beta7 loop is an important site for substrate recognition in sortases.  相似文献   

20.
To elucidate the interaction between substrate inhibition and substrate transglycosylation of retaining glycoside hydrolases (GHs), a steady-state kinetic study was performed for the GH family 3 glucan (1-->3)-beta-glucosidase from the white-rot fungus Phanerochaete chrysosporium, using laminarioligosaccharides as substrates. When laminaribiose was incubated with the enzyme, a transglycosylation product was detected by thin-layer chromatography. The product was purified by size-exclusion chromatography, and was identified as a 6-O-glucosyl-laminaribiose (beta-D-Glcp-(1-->6)-beta-D-Glcp-(1-->3)-D-Glc) by 1H NMR spectroscopy and electrospray ionization mass spectrometry analysis. In steady-state kinetic studies, an apparent decrease of laminaribiose hydrolysis was observed at high concentrations of the substrate, and the plots of glucose production versus substrate concentration were thus fitted to a modified Michaelis-Menten equation including hydrolytic and transglycosylation parameters (K(m), K(m2), k(cat), k(cat2)). The rate of 6-O-glucosyl-laminaribiose production estimated by high-performance anion-exchange chromatography coincided with the theoretical rate calculated using these parameters, clearly indicating that substrate inhibition of this enzyme is fully explained by substrate transglycosylation. Moreover, when K(m), k(cat), and affinity for glucosyl-enzyme intermediates (K(m2)) were estimated for laminarioligosaccharides (DP=3-5), the K(m) value of laminaribiose was approximately 5-9 times higher than those of the other oligosaccharides (DP=3-5), whereas the K(m2) values were independent of the DP of the substrates. The kinetics of transglycosylation by the enzyme could be well interpreted in terms of the subsite affinities estimated from the hydrolytic parameters (K(m) and k(cat)), and a possible mechanism of transglycosylation is proposed.  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