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.
Guha S  Sahu K  Roy D  Mondal SK  Roy S  Bhattacharyya K 《Biochemistry》2005,44(25):8940-8947
Solvation dynamics at the active site of an enzyme, glutaminyl-tRNA synthetase (GlnRS), was studied using a fluorescence probe, acrylodan, site-specifically attached at cysteine residue C229, near the active site. The picosecond time-dependent fluorescence Stokes shift indicates slow solvation dynamics at the active site of the enzyme, in the absence of any substrate. The solvation dynamics becomes still slower when the substrate (glutamine or tRNA(Gln)) binds to the enzyme. A mutant Y211H-GlnRS was constructed in which the glutamine binding site is disrupted. The mutant Y211H-GlnRS labeled at C229 with acrylodan exhibited significantly different solvent relaxation, thus demonstrating that the slow dynamics is indeed associated with the active site. Implications for catalysis and specificity have been discussed.  相似文献   

2.
In living systems, tungsten is exclusively found in microbial enzymes coordinated by the pyranopterin cofactor, with additional metal coordination provided by oxygen and/or sulfur, and/or selenium atoms in diverse arrangements. Prominent examples are formate dehydrogenase, formylmethanofuran dehydrogenase, and aldehyde oxidoreductase all of which catalyze redox reactions. The bacterial enzyme acetylene hydratase (AH) stands out of its class as it catalyzes the conversion of acetylene to acetaldehyde, clearly a non-redox reaction and a reaction distinct from the reduction of acetylene to ethylene by nitrogenase. AH harbors two pyranopterins bound to W, and a [4Fe–4S] cluster. W is coordinated by four dithiolene sulfur atoms, one cysteine sulfur, and one oxygen ligand. AH activity requires a strong reductant suggesting W(IV) as the active oxidation state. Two different types of reaction pathways have been proposed. The 1.26 Å structure reveals a water molecule coordinated to W which could gain a partially positive net charge by the adjacent protonated Asp-13, enabling a direct attack of C2H2. To access the W–Asp site, a substrate channel was evolved distant from where it is found in other members of the DMSOR family. Computational studies of this second shell mechanism led to unrealistically high energy barriers, and alternative pathways were proposed where C2H2 binds directly to W. The architecture of the catalytic cavity, the specificity for C2H2 and the results from site-directed mutagenesis do not support this first shell mechanism. More investigations including structural information on the binding of C2H2 are needed to present a conclusive answer.  相似文献   

3.
The soluble tungsten, iron-sulfur enzyme acetylene hydratase (AH) from mesophilic Pelobacter acetylenicus is a member of the dimethyl sulfoxide (DMSO) reductase family. It stands out from its class as it catalyzes a nonredox reaction, the addition of H2O to acetylene (H—C☰C—H) to form acetaldehyde (CH3CHO). Caught in its active W(IV) state, the high-resolution three-dimensional structure of AH offers an excellent starting point to tackle its unique chemistry and to identify catalytic amino acid residues within the active site cavity: Asp13 close to W(IV) coordinated to two molybdopterin-guanosine-dinucleotide ligands, Lys48 which couples the [4Fe-4S] cluster to the W site, and Ile142 as part of a hydrophobic ring at the end of the substrate access channel designed to accommodate the substrate acetylene. A protocol was developed to express AH in Escherichia coli and to produce active-site variants which were characterized with regard to activity and occupancy of the tungsten and iron-sulfur centers. By this means, fusion of the N-terminal chaperone binding site of the E. coli nitrate reductase NarG to the AH gene improved the yield and activity of AH and its variants significantly. Results from site-directed mutagenesis of three key residues, Asp13, Lys48, and Ile142, document their important role in catalysis of this unusual tungsten enzyme.Molybdenum and tungsten are the only transition metals of the second (Mo) and third (W) row of the periodic table of elements with known biological functions (7). In their biologically active form, both metals are bound to the cofactor molybdopterin (Moco), which is present in all molybdenum and tungsten enzymes with the exception of nitrogenase, where molybdenum is coordinated to a large iron-sulfur cluster, MoFe7S9 (9). Virtually all organisms including plants and mammals use either molybdenum or tungsten proteins in important metabolic pathways (35). Microorganisms carry a wide variety of molybdenum enzymes, such as nitrate reductase (NAR), formate dehydrogenase (FDH), dimethyl sulfoxide reductase (DMSOR), or trimethylamine N-oxide reductase (TMAOR) (7). These enzymes are involved in either oxygen atom transfer reactions or in reductive hydroxylations. By this means, the metal shuttles between the oxidation states +IV and +VI (16). Notably, the tungsten, iron-sulfur enzyme acetylene hydratase ([AH] EC 4.2.1.112), isolated from the soluble fraction of the mesophilic anaerobe Pelobacter acetylenicus, is an exception (26). It catalyzes the hydration of acetylene to acetaldehyde via an enol intermediate as an initial step for the fermentation of acetylene by P. acetylenicus, clearly a nonredox reaction (equation 1): Except for nitrogenase, which reduces acetylene to ethylene (H2C=CH2), AH is the only enzyme known to accept acetylene as a substrate. However, acetylene is well known to act as an inhibitor for numerous metal-dependent enzymes (10). AH is a member of the DMSOR family and carries one [4Fe-4S] cluster and two molybdopterin-guanosine-dinucleotide (referred to as P- and Q-MGD) ligands that coordinate the tungsten atom (Fig. (Fig.1)1) (18). The enzyme is sensitive toward dioxygen, and its [4Fe-4S] cluster is converted to a truncated [3Fe-4S] cluster upon exposure to air, as shown by electron paramagnetic resonance (EPR) (18). In AH prepared under the exclusion of dioxygen, the EPR signal of the [3Fe-4S] cluster was absent, and reaction with sodium dithionite led to a rhombic EPR signal (gz of 2.048, gy of 1.939, and gx of 1.920) originating from a [4Fe-4S]+ cluster. Upon oxidation with hexacyanoferrate(III), a new signal appeared (gx of 2.007, gy of 2.019, and gz of 2.048; average g value [gav] of 2.022), which was assigned to a W(V) center (18).Open in a separate windowFIG. 1.Acetylene hydratase from P. acetylenicus. (Left) Overall structure, with the [4Fe-4S] cluster and W(MGD)2 buried inside the protein; the peptide backbone is shown in dark blue at the N-terminal end, and continues as light blue, cyan, green, and yellow to orange at the C-terminal end. (Right) [4Fe-4S] cluster and W(MGD)2 center. C is shown in gray, N in blue, O in red, P in orange, S in yellow, Fe in brown, and W in cyan (Protein Data Bank [PDB] code 2E7Z).For catalytic activity, AH requires a strong reductant, such as sodium dithionite or titanium(III) citrate (18). Recently, the X-ray structure of AH in the reduced state could be solved at 1.26-Å resolution (28) which gave a first view of its active site: W(IV) is coordinated by four sulfur atoms delivered by the two dithiolene ligands (MGD), one cysteinyl sulfur (Cys141), and one oxygen ligand at a distance of 2.04 Å (Fig. (Fig.1).1). Mechanistically, the nature of this oxygen ligand is critical. The observed W-O distance of 2.04 Å is right between the values expected for a hydroxide ligand (1.9 to 2.1 Å) and a coordinated water (2.0 to 2.3 Å), thus not allowing an unequivocal assignment of the sixth ligand of the WS5O core. Two different reaction mechanisms have been proposed: (i) nucleophilic attack of the hydroxide group and (ii) electrophilic attack of a polarized water molecule, on the C,C triple bond of acetylene (28). As a consequence of theoretical calculations, and in agreement with the observed bond distances, the active W(IV) state should favor a water ligand and therefore an electrophilic addition mechanism (28). Acetylene can access the tungsten ion through a well-defined channel close to the N-terminal domain that harbors the [4Fe-4S] cluster. One residue, Asp13, interacts with the oxygen ligand bound to the W ion, forming a short hydrogen bond of 2.4 Å. Above the W ion and the coordinated H2O molecule, the substrate channel ends in a ring of six hydrophobic residues. These residues build a cavity with dimensions perfect for accommodating acetylene. Experiments to bind the substrate acetylene, ethylene, the inhibitor propargyl alcohol (H—C☰C—CH2OH), and dinitrogen or carbon monoxide have failed thus far to produce a complex in the crystal. However, computer docking of one acetylene molecule at this position led to a reasonable fit, positioning the two carbon atoms of the substrate exactly above the H2O molecule coordinated to tungsten (28).To gain further information about the reaction mechanism of AH, we initiated a study by site-directed mutagenesis and exchanged several amino acids which have been suggested to be important for catalysis at the active site cavity. To achieve this goal, we had to develop a suitable procedure for the heterologous expression of AH in Escherichia coli. Notably, E. coli uses a chaperone system for the insertion of Moco into its enzymes (6). These chaperones of the TorD superfamily act in two ways. First, they bind at the N-terminal signal sequence, similar to the sequence of the TAT export system, thereby delaying the folding of the newly synthesized molybdenum enzyme until Moco has been properly inserted (25). Second, they actively facilitate the incorporation of the molybdopterin cofactor by binding to a second, yet unknown site (13). To improve the assembly of the metal sites as well as to increase the enzymatic activity of the recombinant AH, the N-terminal chaperone binding site of the nitrate reductase, NarG, from E. coli was fused to the AH gene in the expression vector.  相似文献   

4.
《BBA》2022,1863(3):148523
The Mo-dependent enzyme YiiM enzyme from Escherichia coli is a member of the sulfite oxidase family and shares many similarities with the well-studied human mitochondrial amidoxime reducing component (mARC). We have investigated YiiM catalysis using electrochemical and spectroscopic methods. EPR monitored redox potentiometry found the active site redox potentials to be MoVI/V –0.02 V and MoV/IV –0.12 V vs NHE at pH 7.2. In the presence of methyl viologen as an electrochemically reduced electron donor, YiiM catalysis was studied with a range of potential substrates. YiiM preferentially reduces N-hydroxylated compounds such as hydroxylamines, amidoximes, N-hydroxypurines and N-hydroxyureas but shows little or no activity against amine-oxides or sulfoxides. The pH optimum for catalysis was 7.1 and a bell-shaped pH profile was found with pKa values of 6.2 and 8.1 either side of this optimum that are associated with protonation/deprotonations that modulate activity. Simulation of the experimental voltammetry elucidated kinetic parameters associated with YiiM catalysis with the substrates 6–hydroxyaminopurine and benzamidoxime.  相似文献   

5.
Enzymes enhance chemical reaction rates by lowering the activation energy, the energy barrier of the reaction leading to products. This occurs because enzymes bind the high-energy intermediate of the reaction (the transition state) more strongly than the substrate. We studied details of this process by determining the substrate binding energy (DeltaG(s), calculated from K(m) values) and the activation energy (DeltaG(T), determined from k(cat)/K(m) values) for the trypsin-catalyzed hydrolysis of oligopeptides. Plots of DeltaG(T) versus DeltaG(s) for oligopeptides with 15 amino acid replacements at each of the positions P(1)', P(1), and P(2) were straight lines, as predicted by a derived equation that relates DeltaG(T) and DeltaG(s). The data led to the conclusion that the trypsin active site has subsites that bind moieties of substrate and of transition state in characteristic ratios, whichever substrate is used. This was unexpected and means that each subsite characteristically favors substrate binding or catalysis.  相似文献   

6.
The hairpin ribozyme is a small catalytic RNA that accelerates reversible cleavage of a phosphodiester bond. Structural and mechanistic studies suggest that divalent metals stabilize the functional structure but do not participate directly in catalysis. Instead, two active site nucleobases, G8 and A38, appear to participate in catalytic chemistry. The features of A38 that are important for active site structure and chemistry were investigated by comparing cleavage and ligation reactions of ribozyme variants with A38 modifications. An abasic substitution of A38 reduced cleavage and ligation activity by 14,000-fold and 370,000-fold, respectively, highlighting the critical role of this nucleobase in ribozyme function. Cleavage and ligation activity of unmodified ribozymes increased with increasing pH, evidence that deprotonation of some functional group with an apparent pK(a) value near 6 is important for activity. The pH-dependent transition in activity shifted by several pH units in the basic direction when A38 was substituted with an abasic residue, or with nucleobase analogs with very high or low pK(a) values that are expected to retain the same protonation state throughout the experimental pH range. Certain exogenous nucleobases that share the amidine group of adenine restored activity to abasic ribozyme variants that lack A38. The pH dependence of chemical rescue reactions also changed according to the intrinsic basicity of the rescuing nucleobase, providing further evidence that the protonation state of the N1 position of purine analogs is important for rescue activity. These results are consistent with models of the hairpin ribozyme catalytic mechanism in which interactions with A38 provide electrostatic stabilization to the transition state.  相似文献   

7.
Use of radioactive molybdenum demonstrates that the tetrathionate reductase of Salmonella typhimurium is a molydenum containing enzyme. It is proposed that this enzyme shares with other molybdo-proteins, such as nitrate reductase, a common molybdenum containing cofactor the defect of which leads to the loss of the tetrathionate reductase and nitrate reductase activities.  相似文献   

8.
Thirty-seven sets of crystallographic enzyme data were selected from Protein Data Bank (PDB, 1995). The average temperature factors (B) of the critical residues at the active site and the whole molecule of those enzymes were calculated respectively. The statistical results showed that the critical residues at the active site of most of the enzymes had lowerB factors than did the whole molecules, indicating that in the crystalline state the critical residues at the active site of the natural enzymes possess more stable conformation than do the whole molecules. The flexibility of the active site during the unfolding by denaturing was also discussed.  相似文献   

9.
The ATP-competitive inhibitors dasatinib and nilotinib, which bind to catalytically different conformations of the Abl kinase domain, have recently been approved for the treatment of imatinib-resistant CML. These two new drugs, albeit very efficient against most of the imatinib-resistant mutants of Bcr–Abl, fail to effectively suppress the Bcr–Abl activity of the T315I (or gatekeeper) mutation. Generating new ATP site-binding drugs that target the T315I in Abl has been hampered, amongst others, by target selectivity, which is frequently an issue when developing ATP-competitive inhibitors. Recently, using an unbiased cellular screening approach, GNF-2, a non-ATP-competitive inhibitor, has been identified that demonstrates cellular activity against Bcr–Abl transformed cells. The exquisite selectivity of GNF-2 is due to the finding that it targets the myristate binding site located near the C-terminus of the Abl kinase domain, as demonstrated by genetic approaches, solution NMR and X-ray crystallography. GNF-2, like myristate, is able to induce and/or stabilize the clamped inactive conformation of Abl analogous to the SH2-Y527 interaction of Src. The molecular mechanism for allosteric inhibition by the GNF-2 inhibitor class, and the combined effects with ATP-competitive inhibitors such as nilotinib and imatinib on wild-type Abl and imatinib-resistant mutants, in particular the T315I gatekeeper mutant, are reviewed.  相似文献   

10.
Fatty acid synthetase from goose uropygial gland was inactivated by treatment with pyridoxal 5′-phosphate. Malonyl-CoA and acetyl-CoA did not protect the enzyme whereas NADPH provided about 70% protection against this inactivation. 2′-Monophospho-ADP-ribose was nearly as effective as NADPH while 2′-AMP, 5′-AMP, ADP-ribose, and NADH were ineffective suggesting that pyridoxal 5′-phosphate modified a group that interacts with the 5′-pyrophosphoryl group of NADPH and that the 2′-phosphate is necessary for the binding of the coenzyme to the enzyme. Of the seven component activities catalyzed by fatty acid synthetase only the enoyl-CoA reductase activity was inhibited. Inactivation of both the overall activity and enoyl-CoA reductase of fatty acid synthetase by this compound was reversed by dialysis or dilution but not after reduction with NaBH4. The modified protein showed a characteristic Schiff base absorption (maximum at 425 nm) that disappeared on reduction with NaBH4 resulting in a new absorption spectrum with a maximum at 325 nm. After reduction the protein showed a fluorescence spectrum with a maximum at 394 nm. Reduction of pyridoxal phosphate-treated protein with NaB3H4 resulted in incorporation of 3H into the protein and paper chromatography of the acid hydrolysate of the modified protein showed only one fluorescent spot which was labeled and ninhydrin positive and had an Rf identical to that of authentic N6-pyridoxyllysine. When [4-3H]pyridoxal phosphate was used all of the 3H, incorporated into the protein, was found in pyridoxyllysine. All of these results strongly suggest that pyridoxal phosphate inhibited fatty acid synthetase by forming a Schiff base with the ?-amino group of lysine in the enoyl-CoA reductase domain of the enzyme. The number of lysine residues modified was estimated with [4-3H]pyridoxal-5′-phosphate/NaBH4 and by pyridoxal-5′-phosphate/NaB3H4. Scatchard analysis showed that modification of two lysine residues per subunit resulted in complete inactivation of the overall activity and enoyl-CoA reductase of fatty acid synthetase. NADPH prevented the inactivation of the enzyme by protecting one of these two lysine residues from modification. The present results are consistent with the hypothesis that each subunit of the enzyme contains an enoyl-CoA reductase domain in which a lysine residue, at or near the active site, interacts with NADPH.  相似文献   

11.
A series of chemical modification reactions have been carried out with rabbit pulmonary angiotensin converting enzyme (dipeptidyl carboxypeptidase, EC 3.4.15.1) in order to identify amino acid residues essential for its catalytic activity. The enzyme is rapidly inactivated by nitration with tetranitromethane and by O-acetylation with N-acetylimidazole. Deacylation with hydroxylamine restores activity to the acetylated enzyme, while the inhibitor, β-phenylpropionyl-L-phenylalanine, protects against acetylimidazole inactivation. These results indicate the presence of functional tyrosyl residues at the active site of the enzyme. Reaction with butanedione decreases activity, an effect that is markedly enhanced by the presence of borate, indicating essential arginyl residues. In addition, activity is diminished by the carboxyl reagent, cyclohexylmorpholinoethyl carbodiimide. Thus, the three functional residues long known to be components of the active site of bovine carboxypeptidase A, tyrosyl, arginyl, and glutamyl, have counterparts in the angiotensin converting enzyme. The effects of pyridoxal phosphate and a number of other reagents demonstrate that the converting enzyme also contains an important lysyl residue.  相似文献   

12.
The program GRID was used to design potential inhibitors of human L-xylulose reductase based on a model of the holoenzyme in complex with n-butyric acid. The inclusion of phosphate or carboxylate functional groups in the ligand suggested an increase in the net binding energy of the complex up to 2.8- and 4.0-fold, respectively. This study may be useful in the development of potent and specific inhibitors of the enzyme.  相似文献   

13.
Crystallographic B factor of critical residues at enzyme active site   总被引:1,自引:0,他引:1  
Thirty-seven sets of crystallographic enzyme data were selected from Protein Data Bank (PDB, 1995). The average temperature factors (B) of the critical residues at the active site and the whole molecule of those enzymes were calculated respectively. The statistical results showed that the critical residues at the active site of most of the enzymes had lower B factors than did the whole molecules, indicating that in the crystalline state the critical residues at the active site of the natural enzymes possess more stable conformation than do the whole molecules. The flexibility of the active site during the unfolding by denaturing was also discussed.  相似文献   

14.
Conditions for heterologous expression of Rhodobacter sphaeroides biotin sulfoxide reductase in Escherichia coli were modified, resulting in a significant improvement in the yield of recombinant enzyme and enabling structural studies of the molybdenum center. Quantitation of the guanine and the molybdenum as compared to that found in R. sphaeroides DMSO reductase demonstrated the presence of the bis(MGD)molybdenum cofactor. UV-visible absorption spectra were obtained for the oxidized, NADPH-reduced, and dithionite-reduced enzyme. EPR spectra were obtained for the Mo(V) state of the enzyme. X-ray absorption spectroscopy at the molybdenum K-edge has been used to probe the molybdenum coordination of the enzyme. The molybdenum site of the oxidized protein possesses a Mo(VI) mono-oxo site (Mo=O at 1.70 A) with additional coordination by approximately four thiolate ligands at 2.41 A and probably one oxygen or nitrogen at 1.95 A. The NADPH- and dithionite-reduced Mo(IV) forms of the enzyme are des-oxo molybdenum sites with approximately four thiolates at 2.33 A and two different Mo-O/N ligands at 2.19 and 1.94 A.  相似文献   

15.
Mercuric reductase, a flavoprotein disulfide oxidoreductase, catalyzes the two-electron reduction of Hg(II) to Hg(0) by NADPH. As with all the members of this class of proteins, the enzyme is a dimer of identical subunits with two active sites per dimer, each composed of one FAD and catalytically essential residues from both subunits. In the enzyme from Tn501, these residues include, at a minimum, FAD and cysteines 135 and 140 from one subunit and cysteines 558' and 559' from the other. With this sort of active site arrangement, the enzyme seems perfectly set up for some type of subunit communication. In this report, we present results from several titrations, as well as kinetics studies, that, taken together, are consistent with the occurrence of subunit communication. In particular, the results indicate that pyridine nucleotide complexed dimers of the enzyme are asymmetric. Since the EH2-NADPH complex of the enzyme is the relevant reductant of Hg(II), these observations suggest that the enzyme may function asymmetrically during catalysis. An alternating sites model is proposed for the catalytic reduction of Hg(II), where both subunits of the dimer function in catalysis, but the steps are staggered and the subunits reverse roles after part of the reaction. An attractive feature of this proposal is that it provides a reasonable solution to the thermodynamic dilemma the enzyme faces in needing to both bind Hg(II) very tightly and reduce it.  相似文献   

16.
A 1.9 A resolution X-ray structure of the apo-form of Corynebacterium 2,5-diketo-d-gluconic acid reductase A (2,5-DKGR A), a member of the aldo-keto reductase superfamily, has been determined by molecular replacement using the NADPH-bound form of the same enzyme as the search model. 2,5-DKGR A catalyzes the NADPH-dependent stereo-specific reduction of 2,5-diketo-d-gluconate (2,5-DKG) to 2-keto-l-gulonate, a precursor in the industrial production of vitamin C. An atomic-resolution structure for the apo-form of the enzyme, in conjunction with our previously reported high-resolution X-ray structure for the holo-enzyme and holo/substrate model, allows a comparative analysis of structural changes that accompany cofactor binding. The results show that regions of the active site undergo coordinated conformational changes of up to 8 A. These conformational changes result in the organization and structural rearrangement of residues associated with substrate binding and catalysis. Thus, NADPH functions not only to provide a hydride ion for catalytic reduction, but is also a critical structural component for formation of a catalytically competent form of DKGR A.  相似文献   

17.
Bray RC  Adams B  Smith AT  Bennett B  Bailey S 《Biochemistry》2000,39(37):11258-11269
Much is unknown concerning the role of thiolate ligands of molybdenum in molybdopterin enzymes. It has been suggested that thiolate dissociation from molybdenum is part of the catalytic mechanism of bis-molybdopterin enzymes of the dimethyl sulfoxide reductase (DMSOR) family. For DMSOR from Rhodobacter capsulatus, thiolate dissociation has therefore been investigated crystallographically, by UV/visible spectroscopy, and by enzyme assays. When crystallized from sodium citrate, all four thiolates of DMSOR are within bonding distance of Mo, but after extended exposure to Na(+)-Hepes, a pair of thiolates dissociates, a mixture of structures being indicated after shorter exposures to this buffer. DMSOR is stable in sodium citrate and other buffers but unstable aerobically although not anaerobically in Na(+)-Hepes. Aerobically in Na(+)-Hepes, a first-order reaction (k = 0.032 hr(-)(1) at 37 degrees C) leads to loss of activity in the backward but not the forward (dimethyl sulfoxide reduction) assay and loss of absorption at lambda > approximately 450 nm. This reaction can be reversed by a cycle of reduction and reoxidation ("redox-cycling"). Slower irreversible loss of activity in the forward assay and cofactor dissociation follow. Spectral analogy with a mono-molybdopterin enzyme supports the conclusion that in the Hepes-modified DMSOR form, only two cofactor dithiolene sulfur atoms are coordinated to molybdenum. Loss of activity provides the first clear evidence that sulfur ligand dissociation is an artifact, not part of the catalytic cycle. Clearly, structural data on DMSOR samples extensively exposed to Hepes is not directly relevant to the native enzyme. The nature of the oxygen ligands detected crystallographically is discussed, as is the specificity of Hepes and the mechanism whereby its effects are achieved. DMSOR forms complexes with Na(+)-Hepes and other buffer ions. For DMSOR crystallized from Hepes, electron density in the substrate binding channel suggests that buffers bind in this site. Like the as-prepared enzyme, the modified form (DMSOR(mod)D), known to arise on extended aerobic exposure to dimethyl sulfide, is susceptible to a further degradative reaction, although this is not buffer-dependent. It involves loss of absorption at lambda > approximately 450 nm and, presumably, dissociation of thiolate ligands. Evidence is presented that, as a result of O(2) damage, DMSOR samples not submitted to redox-cycling may be contaminated with DMSOR(mod)D and with material absorbing in the region of 400 nm, analogous to the Hepes-modified enzyme. Since the latter lacks absorption at lambda > approximately 450 nm, its presence may escape detection.  相似文献   

18.
Inhibition of bioluminescence in Photobacterium phosphoreum by cerulenin has been demonstrated to be due to a specific inactivation of the acyl-CoA reductase subunit of the fatty acid reductase complex required for synthesis of the aldehyde substrate for the luminescent reaction. In contrast, the activities of the other luminescence-related enzymes, acyl-protein synthetase, acyl-transferase, and luciferase, were unaffected by cerulenin. Myristoyl-CoA, but not NADPH, protected the acyl-CoA reductase against cerulenin inhibition. Cerulenin blocked the acylation of the reductase with myristoyl-CoA and the reaction with N-ethylmaleimide. A shift in mobility of the reductase polypeptide on sodium dodecyl sulfate - polyacrylamide gel electrophoresis occurred after reaction with cerulenin, a shift which could be blocked by reaction with N-ethylmaleimide. These results demonstrate that cerulenin blocks aldehyde synthesis by covalent reaction with the acyl-CoA reductase and indicate that the reaction may occur at a cysteine residue involved in the formation of the acyl-reductase intermediate.  相似文献   

19.
Methionine sulfoxide reductase A is an essential enzyme in the antioxidant system which scavenges reactive oxygen species through cyclic oxidation and reduction of methionine and methionine sulfoxide. Recently it has also been shown to catalyze the reverse reaction, oxidizing methionine residues to methionine sulfoxide. A cysteine at the active site of the enzyme is essential for both reductase and oxidase activities. This cysteine has been reported to have a pK(a) of 9.5 in the absence of substrate, decreasing to 5.7 upon binding of substrate. Using three independent methods, we show that the pK(a) of the active site cysteine of mouse methionine sulfoxide reductase is 7.2 even in the absence of substrate. The primary mechanism by which the pK(a) is lowered is hydrogen bonding of the active site Cys-72 to protonated Glu-115. The low pK(a) renders the active site cysteine susceptible to oxidation to sulfenic acid by micromolar concentrations of hydrogen peroxide. This characteristic supports a role for methionine sulfoxide reductase in redox signaling.  相似文献   

20.
Feng J  Goswami S  Howell EE 《Biochemistry》2008,47(2):555-565
R67 dihydrofolate reductase (DHFR) bears no sequence or structural homologies with chromosomal DHFRs. The gene for this enzyme produces subunits that are 78 amino acids long, which assemble into a homotetramer possessing 222 symmetry. More recently, a tandem array of four gene copies linked in-frame was constructed, which produces a monomer containing 312 amino acids named Quad3. Asymmetric mutations in Quad3 have also been constructed to probe the role of Q67 and K32 residues in catalysis. This present study mixes and matches mutations to determine if the Q67H mutation, which tightens binding approximately 100-fold to both dihydrofolate (DHF) and NADPH, can help rescue the K32M mutation. While the latter mutation weakens DHF binding over 60-fold, it concurrently increases kcat by a factor of 5. Two Q67H mutations were added to gene copies 1 and 4 in conjunction with the K32M mutation in gene copies 1 and 3. Addition of these Q67H mutations tightens binding 40-fold, and the catalytic efficiency (kcat/Km(DHF)) of the resulting protein is similar to that of Quad3. Since these Q67H mutations can mostly compensate for the K32M lesion, K32 must not be necessary for DHF binding. Another multimutant combines the K32M mutation in gene copies 1 and 3 with the Q67H mutation in all gene copies. This mutant is inhibited by DHF but not NADPH, indicating that NADPH binds only to the wild type half of the pore, while DHF can bind to either the wild type or mutant half of the pore. This inhibition pattern contrasts with the mutant containing only the Q67H substitution in all four gene copies, which is severely inhibited by both NADPH and substrate. Since gene duplication and divergence are evolutionary tools for gaining function, these constructs are a first step toward building preferences for NADPH and DHF in each half of the active site pore of this primitive enzyme.  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