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 661 毫秒
1.
Circular dichroism (CD) spectra and UV absorption spectra of two obligatory intermediates in rhodanese catalysis were compared. A broad CD band between 250 and 287 nm increased in a manner stoichiometrically related to the content of enzyme-bound persulfide. Titration of a sample of sulfur-substituted rhodanese (ES) with either cyanide or sulfite gave a stoichiometry that is consistent with one persulfide/molecule of rhodanese (Mr = 33,000). This result agrees with that determined by x-ray crystallography and a method based on quenching of intrinsic fluorescence. Cyanolysis of the persulfide in ES is accompanied by a decrease of UV absorption in the region between 250 and 300 nm. Cyanide titrations followed by the change in absorbance at 263, 272, and 292 nm gave the expected stoichiometry. The magnitude of the difference between the far UV-CD spectra of E and ES found here is smaller than reported previously. This variability suggests that the differences in the secondary structure of these intermediates may not be obligatorily related to the cyanolysis of the persulfide. This view is compatible with recent evidence which suggested that E and ES may be made different by structural relaxation events that occur outside of the catalytic cycle. Furthermore, the methods developed here will be useful in studies on the stability of the catalytic persulfide that has been suggested to be central in the mechanism of several enzymes important in sulfur metabolism.  相似文献   

2.
The accessibilities of sites on the surface of the enzyme rhodanese for binding to macromolecular apolarity have been measured for the two forms of the enzyme related to obligatory catalytic intermediates: the free enzyme, E and the sulfur substituted enzyme, ES. This study was done using a micromethod developed for this purpose which allows facile assessment of the apolar binding of proteins to commercially available beads of cross-linked agarose on which hydrophobic groups have been immobilized. The results indicate that the enzyme rhodanese can bind to macromolecular apolarity and that there is considerably more binding of the E form than the ES form. The fact that the binding is relatively slow implicates a protein conformational change in the rate limiting binding step. In fact, there is a large increase in the binding when the temperature is raised from 23° to 40° which correlates with previous results showing a conformational change in rhodanese over the same temperature range. These results in comparison with other solution studies and with x-ray studies are consistent with a model for rhodanese which has an apolar active site and a mechanism for catalysis that includes a conformational change.  相似文献   

3.
The steady-state and pre-steady-state kinetic parameters for the interaction of E with the fluorogenic substrate 2-aminobenzoyl-Thr-Ile-Nle-Phe(p-NO(2))-Gln-Arg-NH(2) were determined in 1.25 M NaCl, 0.1 M MES-TRIS at pH 6.0 at 25 degrees C. At low concentrations of enzyme, the values of the K(m) and k(cat) calculated from steady-state data were 2.1 microM and 7.4 s(-1), respectively. At high concentrations of enzyme, the time-courses of fluorescence enhancement associated with catalysis were very dependent on the excitation wavelength used to monitor the reaction. Because the absorbance spectrum of the substrate overlapped the fluorescence emission spectrum of the enzyme, these abnormalities were attributed to fluorescence energy transfer between the enzyme and the substrate in an enzyme-substrate intermediate. The kinetic data collected with lambda(ex) = 280 nm and lambda(em) > 435 nm were analyzed according to the following mechanism in which EX was the species with enhanced fluorescence relative to substrate or products: [formula see text]. The values of the kinetic parameters with (1)H(2)O as the solvent were K = 13 microM, k(2) = 150 s(-1), k(-2) = 25 s(-1), and k(3) = 11 s(-1). The values of the kinetic parameters with (2)H(2)O as the solvent were K = 13 microM, k(2) = 210 s(-1), k(-2) = 12 s(-1), and k(3) = 4.4 s(-1). These values yielded solvent isotope effects of 2 on k(cat) and 0.9 on k(cat)/K(m). From analysis of the complete time-course of the fluorescence change (lambda(ex) = 280 nm and lambda(em) > 435 nm) during the course of substrate hydrolysis, the intermediate EX was determined to be 6.3-fold more fluorescent than the product, which, in turn, was 4.5-fold more fluorescent than ES or S. Rapid quench experiments with 2 N HCl as the quenching reagent confirmed that EX was a complex between enzyme and substrate. Consequently, the small burst in fluorescence observed when monitoring with lambda(ex) = 340 nm (0.3 product equiv per enzyme equivalent) was attributed to the fluorescence change upon transfer of substrate from an aqueous environment to a nonaqueous environment in the enzyme. These results were consistent with carbon-nitrogen bond cleavage being the major contributor to k(cat).  相似文献   

4.
Bovine liver rhodanese (thiosulfate:cyanide sulfurtransferase, EC 2.8.1.1) was prepared in dilute solutions and subjected to conditions that led to a time-dependent loss of enzyme activity. The rate of this activity loss was found to be dependent upon the sulfur substitution state of the enzyme, and the presence or absence of the substrates, thiosulfate and cyanide. In the absence of excess substrates, free enzyme (E), and the covalent intermediate form of the enzyme bearing a divalent sulfur atom in the active site (ES), are of approximately equal functional stability. In comparison, E, in the presence of excess cyanide, was markedly more labile, while ES, supported by 10-50 mM thiosulfate, showed no significant loss of activity under any of the conditions tested. All the enzyme solutions were shown to be losing assayable protein from solution. However, it was demonstrated that, for rhodanese in the E form, the amount of protein lost was insufficient to account for the activity lost, and a marked decline in specific activity was observed. Enzyme in the ES form, whether supported by additional thiosulfate or not, did not decline in the specific activity, though comparable protein loss did occur from these solutions. Intrinsic fluorescence measurements of rhodanese in the ES form, before and after removal of the persulfide sulfur through the addition of cyanide, indicated that loss of enzymic activity was not accompanied by loss of the bound sulfur atom. Therefore, the stabilizing effect observed with thiosulfate could not be explained simply by its ability to maintain enzyme in the sulfur-substituted state. Since the concentration of thiosulfate employed in these experiments was insufficient to maintain all the enzyme in ES.S2O3 form, thiosulfate was acting as a chemical reagent rather than a substrate in stabilizing enzyme activity.  相似文献   

5.
Rhodanese is oxidatively inactivated by several reagents, some of which are not normally considered oxidants. Rhodanese, in a form not containing persulfide sulfur (E), was inactivated by phenylglyoxal under conditions where disulfides are formed. There was the concomitant increase in the fluorescence of the apolar probe 1,1'-bi(4-anilino)naphthalene-5,5'-disulfonic acid (bisANS). At 0.2 mg/ml protein, there was no turbidity, while at 1 mg/ml, turbidity formed after an induction period of 23 min. Phenylglyoxal-inactivated E was extensively digested by endoproteinase glutamate C (V8 protease) to give two discrete high molecular weight fragments (Mr = 29,500 and 16,000). Enzymatically active E or ES, the form of rhodanese containing transferred sulfur (Mr = 33,000) was totally refractory to V8 protease and gave only small fluorescent enhancement of bisANS. Phenylglyoxal inactivated ES (reaction at arginine) gave very little fluorescence enhancement of bisANS and was not digested by V8. Hydrogen peroxide rapidly inactivated E (t1/2 less than 2 min) giving a slow increase in bisANS fluorescence (t1/2 greater than 10 min) identical to that observed with phenylglyoxal. The turbidity also increased after an induction period of approximately 30 min. Inactivation of E by hydrogen peroxide gave the same digestion pattern as that observed with phenylglyoxal inactivation. The turbidity was associated with the formation of disulfide-bonded structures that formed with the stoichiometry of E, 2E, 4E, 6E, 8E, etc. relative to the native enzyme, E. E was inactivated with several other reagents that lead to oxidatively inactivated rhodanese including NADH, dithiothreitol, mercaptoethanol, and m-dinitrobenzene. Enzyme inactivated with dithiothreitol or NADH gave an identical digestion pattern as above. In addition, with the exception of NADH which could not be used due to optical interference, each of the reagents gave rise to increased fluorescence of bisANS after inactivation. The results are consistent with a model in which the oxidized rhodanese resulting from diverse treatments is in a new conformation that has extensive exposed apolar surfaces and can form both noncovalent and disulfide-bonded aggregates.  相似文献   

6.
The enzyme rhodanese in the form without transferred sulfur, (E), was inactivated by carboxymethylation with iodoacetic acid (E.IAA), and its conformation was compared with that of E inactivated by oxidative processes (Eox). Formation of E.IAA led to the exposure of binding sites for the fluorescent apolar probe 1,1'-bi(4-anilino)naphthalene-5,5'-disulfonic acid (BisANS). The dissociation constant for BisANS decreased as the concentration of E.IAA decreased and ranged from approximately 200 microM at 1 mg/ml protein to approximately 2 microM at protein concentrations below 0.1 mg/ml. Centrifugation confirmed that E.IAA, but not the underivatized enzyme, could associate. E.IAA was proteolyzable by chymotrypsin or endoproteinase Glu C (V8), while rhodanese containing bound sulfur, ES, was totally refractory, and E was only clipped to a small extent. This constellation of consequences was only previously observed with oxidatively inactivated rhodanese. Fluorescence depolarization measurements of bound BisANS were consistent with exposure of apolar surfaces and association of the protein. The fluorescence spectra of BisANS bound to E.IAA or Eox were identical and distinct from the spectrum of BisANS bound to phenylglyoxal-inactivated ES. Digestion with chymotrypsin was followed using protein and BisANS fluorescence and showed a similar response for E.IAA and Eox. These results indicate that the consequences of forming Eox and E.IAA are very similar. Thus, reaction of the active site sulfhydryl group apparently triggers a conformational change leading to increased protein flexibility and increased exposure of hydrophobic surfaces. In the case of oxidation, the trigger might involve initial formation of an active site sulfenic acid which ultimately gives higher oxidation states that could include disulfides.  相似文献   

7.
The structure of the rhodanese-tetracyanonickelate (E X Ni(CN)2-4) complex has been characterized here in spectral and physical studies using urea as a structural perturbant. UV difference absorption, sedimentation velocity ultracentrifugation, fluorescence, and circular dichroism data show no significant conformational differences between sulfur-free rhodanese (E) and the E X Ni(CN)2-4 complex. The urea-induced enzyme structural transition curves were noncoincident when different structural parameters were monitored. For E, the urea concentrations giving half-maximal change (Cm) were: Cm = 3.0 M for activity measurement; Cm = 2.8 M for protein intrinsic fluorescence intensity; Cm = 4.3 M for ellipticity at 220 nm; and Cm = 3.3 M for wavelength of fluorescence emission maximum. For the E X Ni(CN)2-4 complex, Cm was shifted to a higher urea concentration relative to that found for E when activity (Cm = 3.6 M) and native protein fluorescence (Cm = 3.6 M) were the measured parameters but not when the wavelength of the emission maximum and ellipticity were monitored. Furthermore, urea-induced rhodanese structural changes were time-dependent and Ni(CN)2-4 binding on E slowed enzyme inactivation that is associated with structural relaxations. These findings, that Ni(CN)2-4 affects structural relaxations in rhodanese, are of particular interest in light of the recent suggestion that the E X Ni(CN)2-4 complex mimics a normally inaccessible intermediate in catalysis.  相似文献   

8.
Studies have been performed to quantitate the binding of the fluorescent probe 8-anilinonaphthalene-2-sulfonic acid (2,8-ANS) to catalytic intermediates of the enzyme rhodanese: the sulfur-substituted form (ES) and the sulfur-free form (E). The molecular 2,8-ANS has not been extensively used for protein studies, and some characterization is presented to demonstrate its usefulness as a probe for apolar binding sites. The molecule 2,8-ANS binds to at least two classes of sites on rhodanese. One class (class 1) is present in the ES form and has a Kd of 1.7 mM. The E form of rhodanese appears to have a second class of sites (class 2) in addition to the class 1 sites. Two independent fluorometric methods of analyzing the class 2 binding of 2,8-ANS to the E form gave an average value for Kd congruent to 179 microM. These fluorometric titrations, together with a Job plot, clearly indicate that 2,8-ANS binds to more than one site on the E form of rhodanese. The apparent apolarity is slightly higher for class 2 sites than for the class 1 sites, but both give Z factors of greater than 85. The substrate thiosulfate is able to displace the probe that is bound to the class 2 sites on the E form of the enzyme. Further, 2,8-ANS is found to be a competitive inhibitor of the catalyzed reaction with an apparent Kd of 170 microM. Circular dichroism measurements detect no significant changes in the average conformation of rhodanese that can be ascribed to the presence of 2,8-ANS.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Sulfhydryl groups of bovine liver rhodanese (thiosulfate: cyanide sulfurtransferase, EC 2.8.1.1) were modified by treatment with tetrathionate. There was a linear relationship between loss of enzyme activity and the amount of tetrathionate used. At a ratio of one tetrathionate per mole of rhodanese, 100% of enzyme activity was lost in the sulfur-free E-form as compared with a 70% loss for the sulfur-containing ES-form of the enzyme. Addition of up to a 100-fold molar excess of tetrathionate to ES gave no further inactivation. Addition of cyanide to the maximally inactivated ES-tetrathionate complex gave complete loss of activity. Kinetic studies of maximally inactivated ES and partially inactivated E gave Km (Ks) values that were essentially the same as native enzyme, indicating that the active enzyme, in all cases, bound thiosulfate similarly. Reactivation was faster with the ES-form than with the E-form. The substrate, thiosulfate, could reactivate the enzyme up to 70% in 1 h with ES as compared to 24 h with E. Tetrathionate modification of rhodanese could be correlated with the changes in intrinsic fluorescence and with the binding of the active site reporter 2-anilinonaphthalene-8-sulfonic acid (2,8-ANS). Circular dichroism spectra of the protein suggested increased ordered secondary structure in the protein after reaction with tetrathionate. Cadmium chloride and phenylarsine oxide totally inactivated the enzyme at levels usually associated with their effect on enzymes containing vicinal sulfhydryl groups. Further, cadmium inhibition could be reversed by EDTA. Tetrathionate modification of rhodanese may proceed through the formation of sulfenylthiosulfate intermediates at sulfhydryl groups, close to but not identical with the active-site sulfhydryl group, which then can react further with the active-site sulfhydryl group to form disulfide bridges.  相似文献   

10.
The enzyme rhodanese (thiosulfate sulfurtransferase, EC 2.8.1.1) is inactivated on incubation with reducing sugars such as glucose, mannose, or fructose, but is stable with non-reducing sugars or related polyhydroxy compounds. The enzyme is inactivated with (ES) or without (E) the transferable sulfur atom, although E is considerably more sensitive, and inactivation is accentuated by cyanide. Inactivation of E is accompanied by increased proteolytic susceptibility, a decreased sulfhydryl titer, a red-shift and quenching of the protein fluorescence, and the appearance of hydrophobic surfaces. Superoxide dismutase and/or catalase protect rhodanese. Inactive enzyme can be partially reactivated during assay and almost completely reactivated by incubation with thiosulfate, lauryl maltoside, and 2-mercaptoethanol. These results are similar to those observed when rhodanese is inactivated by hydrogen peroxide. These observations, as well as the cyanide-dependent, oxidative inactivation by phenylglyoxal, are explained by invoking the formation of reactive oxygen species such as superoxide or hydrogen peroxide from autooxidation of alpha-hydroxy carbonyl compounds, which can be facilitated by cyanide.  相似文献   

11.
The bacterial enzyme sulfane sulfurtransferase has been studied using spectroscopic techniques. The enzyme was characterized in terms of its near-UV absorption spectrum, molar ellipticity, intrinsic fluorescence spectra and the effects of general and ionic quenching reagents upon its fluorescence. Fluorescence model studies are consistent with sulfane sulfurtransferase having only a single tryptophan residue, which accounts for its low UV absorption coefficient and suggested that this residue is at least partially exposed to solvent. Second derivative absorption spectroscopy studies revealed that most of the bacterial enzyme's tyrosine residues are exposed to solvent. Unlike the better known sulfurtransferase, bovine liver rhodanese, sulfane sulfurtransferase does not undergo a detectable increase in quantum yield when shifting from the sulfur-containing covalent enzyme intermediate to the free enzyme form (which lacks sulfur) during catalysis. CD studies suggest that sulfane sulfurtransferase has a significantly higher proportion of alpha-helix than rhodanese. The renaturation of sulfane sulfurtransferase denatured in 6 M guanidine was shown to be rapid and complete provided that the enzyme had not been oxidized while in the denatured state. Sulfane sulfurtransferase, like rhodanese, catalyzes the transfer of sulfur from thiosulfate to cyanide via a persulfide intermediate, and displays remarkably similar kinetics in this process (Aird, B.A., Heinrikson, R.L. and Westley, J. (1987) J. Biol. Chem 262, 17327-17335). In light of this, the results of the structural studies with sulfane sulfurtransferase are compared and contrasted to data from similar experiments with rhodanese in hopes that they would provide insight about which phenomena observed with rhodanese are intrinsic to the process of transferring sulfur atoms.  相似文献   

12.
Physical and kinetic studies have been used to explore hysteretic effects that are observed in rhodanese catalysis at pH 5 and also at neutral pH when the ionic strength of the medium is high. Experiments that involve observation of changes in intrinsic protein fluorescence of the enzyme and kinetic investigation of its interactions with product thiocyanate anion at pH 5 have implicated enzyme isomerization as the cause of hysteresis. Taken all together, the data indicate that the conformations of enzyme forms in the catalytic cycle are dynamically determined, depending on the relative rates of conformational relaxation and catalysis as influenced by the concentrations of substrates and products.  相似文献   

13.
Unfolded bovine rhodanese, a sulfurtransferase, does not regain full activity upon refolding due to the formation of aggregates and disulfide-linked misfolded states unless a large excess of reductant such as 200 mM -ME and 5 mg/ml detergent are present [Tandon and Horowitz (1990), J. Biol. Chem. 265, 5967]. Even then, refolding is incomplete. We have studied the unfolding and refolding of three rhodanese forms whose crystal structures are known: ES, containing the transferred sulfur as a persulfide; E, without the transferred sulfur, and carboxymethylated rhodanese (CMR), in which the active site was blocked by chemical modification. The X-ray structures of ES, E, and CMR are virtually the same, but their tertiary structures in solution differ somewhat as revealed by near-UV CD. Among these three, CMR is the only form of rhodanese that folds reversibly, requiring 1 mM DTT. A minimum three-state folding model of CMR (NIU) followed by fluorescence at 363 nm, (NI) by fluorescence at 318 nm, and CD (IU) is consistent with the presence of a thermodynamically stable molten globule intermediate in 5–6 M urea. We conclude that the active-site sulfhydryl group in the persulfide form is very reactive; therefore, its modification leads to the successful refolding of urea-denatured rhodanese even in the absence of a large excess of reductant and detergent. The requirement for DTT for complete reversibility of CMR suggests that oxidation among the three non-active-site SH groups can represent a minor trap for refolding through species that can be easily reduced.  相似文献   

14.
For the first time, the enzyme rhodanese has been refolded after denaturation in guanidinium chloride (GdmHCl). Renaturation was by either (a) direct dilution into the assay, (b) intermediate dilution into buffer, or (c) dialysis followed by concentration and centrifugation. Method (c) preferentially retained active enzyme whose specific activity was 1140 IU/mg, which fell to 898 IU/mg after 6 days. The specific activity of native enzyme is 710 IU/mg. Progress curves were linear for the dialyzed enzyme, and kinetic analysis showed it had the same Km for thiosulfate as the native enzyme, but apparently displayed a higher turnover number. Progress curves for denatured enzyme directly diluted into assay mix showed as many as three phases: a lag during which no product formed; a first order reactivation; and an apparently linear steady state. An induction period was determined by extrapolating the steady-state line to the time axis. The percent reactivation fell to 7% (t1/2 = 10 min) as the time increased between GdmHCl dilution and the start of the assay, independent of the presence of thiosulfate. The induction period, which decreased to zero as the incubation time increased, was retained in the presence of thiosulfate. There were no observable differences between native and renatured protein by electrophoresis or fluorescence spectroscopy. Previous reports of some refolding of urea-denatured rhodanese (Stellwagen, E. (1979) J. Mol. Biol. 135, 217-229) were confirmed, extended, and compared with results using GdmHCl. A working hypothesis is that rhodanese refolding involves intermediates that partition into active and inactive products. These intermediates may result from nucleation of the two rhodanese domains, which exposes hydrophobic surfaces that become the interdomain interface in the correctly folded protein.  相似文献   

15.
Hur O  Niks D  Casino P  Dunn MF 《Biochemistry》2002,41(31):9991-10001
Reactions catalyzed by the beta-subunits of the tryptophan synthase alpha(2)beta(2) complex involve multiple covalent transformations facilitated by proton transfers between the coenzyme, the reacting substrates, and acid-base catalytic groups of the enzyme. However, the UV/Vis absorbance spectra of covalent intermediates formed between the pyridoxal 5'-phosphate coenzyme (PLP) and the reacting substrate are remarkably pH-independent. Furthermore, the alpha-aminoacrylate Schiff base intermediate, E(A-A), formed between L-Ser and enzyme-bound PLP has an unusual spectrum with lambda(max) = 350 nm and a shoulder extending to greater than 500 nm. Other PLP enzymes that form E(A-A) species exhibit intense bands with lambda(max) approximately 460-470 nm. To further investigate this unusual tryptophan synthase E(A-A) species, these studies examine the kinetics of H(+) release in the reaction of L-Ser with the enzyme using rapid kinetics and the H(+) indicator phenol red in solutions weakly buffered by substrate L-serine. This work establishes that the reaction of L-Ser with tryptophan synthase gives an H(+) release when the external aldimine of L-Ser, E(Aex(1)), is converted to E(A-A). This same H(+) release occurs in the reaction of L-Ser plus the indole analogue, aniline, in a step that is rate-determining for the appearance of E(Q)(Aniline). We propose that the kinetic and spectroscopic properties of the L-Ser reaction with tryptophan synthase reflect a mechanism wherein the kinetically detected proton release arises from conversion of an E(Aex(1)) species protonated at the Schiff base nitrogen to an E(A-A) species with a neutral Schiff base nitrogen. The mechanistic and conformational implications of this transformation are discussed.  相似文献   

16.
For the first time, the enzyme rhodanese has been proteolytically cleaved to give species that most likely correspond to individual domains. This indicates cleavage can occur in the interdomain tether. Further, the conditions for cleavage show that availability of the susceptible bond(s) depends on conformational changes triggered by oxidative inactivation. Rhodanese, without persulfide sulfur (E), was oxidized consequent to incubation with phenylglyoxal, NADH, or hydrogen peroxide. The oxidized enzyme (Eox) was probed using the proteolytic enzymes endoproteinase glutamate C (V8), trypsin, chymotrypsin, or subtilisin. The proteolytic susceptibility of Eox, formed using hydrogen peroxide, was compared with that of E and the form of the enzyme containing transferred sulfur, ES. ES was totally refractory to proteolysis, while E was only clipped to a small extent by trypsin or V8 and not at all by chymotrypsin or subtilisin. Eox was susceptible to proteolysis by all the proteases used, and, although there were some differences among the proteolytic patterns, there was always a band on sodium dodecyl sulfate-polyacrylamide gel electrophoresis corresponding to Mr = 16,500. This was the only band observed in addition to the parent species (Mr = 33,000) when Eox was digested with chymotrypsin, and conservation of total protein was observed after digestion up to 90 min. No additional species were observable on silver staining, although there was some indication that the band at 16,500 might be a doublet. The results are consistent with the occurrence of a conformational change after oxidation that results in increased exposure and/or flexibility of the interdomain tether which contains residues that meet the specificity requirements of the proteases used.  相似文献   

17.
K L Wun  A Gih  C Sutherland 《Biochemistry》1977,16(5):921-924
The photoreactivating enzyme, PRE, monomerizes pyrimidine dimers in DNA in a light requiring reaction (lambda greater than 300 nm). However, the purified PRE from E. coli has no well-defined absorption band for lambda greater than 300 nm. Using absorption difference spectroscopy, we show that when PRE is mixed with ultraviolet-irradiated DNA, new absorption appears in the spectral region required for catalysis. There is a concomitant decrease in the absorption of the mixture for wavelength less than 300 nm. The hyperchromicity for lambda greater than 300 nm is true absorption, not an artifact due to light scattering. Both the hyperchromicity (lambda greater than 300 nm) and hypochromicity (lambda less than 300 nm) can be reversed by irradiation of 365 nm with identical first-order kinetics. We estimate the molar extinction coefficient of the new absorption to be 6900 +/- 1400 at 350 nm. We conclude that the PRE from E. coli does not possess a distinct "chromophore" which by itself is entirely responsible for the absorption of photoreactivating light. Instead, new absorption results when PRE binds its substrate, dimer-containing DNA.  相似文献   

18.
Reaction of glycolaldehyde with the binary E-NADP complex of bovine kidney aldose reductase (ALR2) produces an enzyme-bound chromophore whose absorbance (lambd max 341 nm) and fluorescence (lambda ex max 341 nm; lambda emit max 421 nm) properties are distinct from those of NADPH or E.NADPH yet are consistent with the proposed covalent adduct structure [1,4-dihydro-4-(1-hydroxy-2-oxoethyl)nicotinamide adenine dinucleotide phosphate]. The kinetics of adduct formation, both in solution and at the enzyme active site, support a mechanism involving rate-determining enolization of glycolaldehyde at high [NADP+] or [E.NADP]. At low [NADP+] or [E.NADP] the reaction is second-order overall, but the ALR2-mediated reaction displays saturation by glycolaldehyde due to competition of the aldehyde (plus hydrate) and enol for E.NADP. Measurement of the pre-steady-state burst of E-adduct formation confirms that glycolaldehyde enol is the reactive species and gives a value of 1.3 x 10(-6) for Kenol = [enol]/[( aldehyde] + [hydrate]), similar to that determined by trapping the enol with I3-. At the ALR2 active site, the rate of adduct formation is enhanced 79,000-fold and the adduct is stabilized greater than or equal to 13,000-fold relative to the reaction with NADP+ in solution. A portion of this enhancement is ascribed to specific interaction of NADP+ with the enzyme since the 3-acetylpyridine analogue, (AP)ADP+, gives values that are 15-200-fold lower. Additional evidence for strong interaction of ALR2 with both NADP+ and NADPH is reported. Yet, because dissociation of adduct is slow, catalysis of the overall adduct formation reaction by ALR2 is less than or equal to 67-fold.  相似文献   

19.
Beef liver rhodanese can be modified covalently at the active site (Cys-247) either reversibly or irreversibly by sulfur, selenium, iodoacetate, and hydrogen peroxide. Each derivative shows an intrinsic fluorescence lower than that of the free enzyme. The reaction of rhodanese with iodoacetate or hydrogen peroxide is time-dependent and accompanied by enzyme inactivation, by the loss of one or two sulfhydryl groups, respectively, by quenching and bathochromic shift of fluorescence, and by an absorbance perturbation in the near UV. The latter findings are indicative for a displacement of some tryptophyl side chains from hydrophobic to hydrophilic environment. The fluorescence decays of the various rhodanese derivatives can be fitted by a double-exponential function with two lifetimes: a shorter one of 1-1.7 ns and a longer one of 2.8-4.6 ns. The S-loaded and Se-loaded rhodanese samples have proportionally shorter lifetimes and lower quantum yields. No such proportionality was observed for the iodoacetate-treated and for the hydrogen peroxide treated enzyme. These findings indicate that two different quenching mechanisms are operating in rhodanese derivatives, a long-range energy transfer from tryptophan to persulfide (or sulfoselenide) group and a static quenching accompanying a conformational change of the protein after modification of the active site.  相似文献   

20.
The notion of "ground-state destabilization" has been well documented in enzymology. It is the unfavourable interaction (strain) in the enzyme-substrate complex, and increases the k(cat) value without changing the k(cat)/K(m) value. During the course of the investigation on the reaction mechanism of aspartate aminotransferase (AAT), we found another type of strain that is crucial for catalysis: the strain of the distorted internal aldimine in the unliganded enzyme. This strain raises the energy level of the starting state (E+S), thereby reducing the energy gap between E+S and ES(++) and increasing the k(cat)/K(m) value. Further analysis on the reaction intermediates showed that the Michaelis complex of AAT with aspartate contains strain energy due to an unfavourable interaction between the main chain carbonyl oxygen and the Tyr225-aldimine hydrogen-bonding network. This belongs to the classical type of strain. In each case, the strain is reflected in the pK(a) value of the internal aldimine. In the historical explanation of the reaction mechanism of AAT, the shifts in the aldimine pK(a) have been considered to be the driving forces for the proton transfer during catalysis. However, the above findings indicate that the true driving forces are the strain energy inherent to the respective intermediates. We describe here how these strain energies are generated and are used for catalysis, and show that variations in the aldimine pK(a) during catalysis are no more than phenomenological results of adjusting the energy levels of the reaction intermediates for efficient catalysis.  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