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 62 毫秒
1.
Methoxycarbonyl-CoA disulfide has been used as an active-site-directed inhibitor of carnitine acetyltransferase. Stoichiometric addition of methoxycarbonyl-CoA disulfide to carnitine acetyltransferase showed the modification of one sulfhydryl group with concomitant loss of about 80% enzyme activity. The rate of modification of this sulfhydryl group is an order of magnitude faster than that of the remaining sulfhydryl groups in the enzyme. Methoxycarbonyl-CoA disulfide inactivation is biphasic: k1 = 1.09 × 102m?1s?1, k2 = 1.1 × 101m?1s?1. This modification, Enz-SS-CoA is covalent; it can be reversed with either dithioerythritol or thiocholine. Acetyl-carnitine and acetyl-CoA protected the enzyme against methoxycarbonyl-CoA disulfide inactivation; however, carnitine did not. These results indicate the presence of a sulfhydryl group in carnitine acetyltransferase at the site of acetyl group transfer. Titration of carnitine acetyltransferase with nonspecific sulfhydryl reagents, DTNB, and ?-nitrophenoxycarbonyl methyl disulfide, revealed that four sulfhydryl groups were preferentially modified by these reagents. The results also show that seven other sulfhydryl groups are available for modification.  相似文献   

2.
Glucosamine-6-phosphate isomerase deaminase (2-amino-2-deoxy-d-glucose-6-phosphate ketol isomerase (deaminating), EC 5.3.1.10) from Escherichia coli is an hexameric homopolymer that contains five half-cystines per chain. The reaction of the native enzyme with 5′,5′-dithiobis-(2-nitrobenzoate) or methyl iodide revealed two reactive SH groups per subunit, whereas a third one reacted only in the presence of denaturants. Two more sulfhydryls appeared when denatured enzyme was treated with dithiothreitol, suggesting the presence of one disulfide bridge per chain. The enzyme having the exposed and reactive SH groups blocked with 5′-thio-2-nitrobenzoate groups was inactive, but the corresponding alkylated derivative was active and retained its homotropic cooperativity toward the substrate, d-glucosamine 6-phosphate, and the allosteric activation by N-acetyl-d-glucosamine 6-phosphate. Studies of SH reactivity in the presence of enzyme ligands showed that a change in the availability of these groups accompanies the allosteric conformational transition. The results obtained show that sulfhydryls are not essential for catalysis or allosteric behavior of glucosamine-6-phosphate deaminase.  相似文献   

3.
beta-Ketoacyl-CoA thiolase (acyl-CoA:acetyl-CoA C-acyltransferase, EC 2.3.1.16) is known to possess sulfhydryl groups of cysteines at the active site that are essential for its catalytic activity. Other groups at the active site that participate in the catalytic process were identified by using anhydride reagents which covalently modify the protein by specifically reacting with any amino groups potentially present at the active site. Since these reagents may also react with thiol groups, the enzyme's amino groups were modified after masking the cysteine thiols present by an alkylalkane thiosulfonate-type reagent, methyl methanethiol-sulfonate (MMTS), that selectively formed a disulfide bridge, thus generating an inactive thiolmethylated enzyme. When this procedure was followed, the enzyme could be undoubtedly modified at its amino by the anhydride reagent, leading to a doubly modified protein. The thiomethyl group could then be removed by reduction with dithiothreitol, yielding an enzyme modified solely on the amino residues. The amino group could be unblocked in turn by exposure to acidic pH. The different anhydrides inactivated thiolase, but only acetoacetyl coenzyme A (AcAcCoA) provided any protection against inactivation. When thiolmethylcitraconyl thiolase was reduced with dithiothreitol the enzyme remained inactive, but when the doubly modified enzyme was exposed to pH 5 then the reduction led to formation of an active enzyme. These results are interpreted as demonstrating a role for an amino group at the enzyme active site. A catalytic mechanism is proposed for the enzyme which involves the amino group.  相似文献   

4.
NADP-linked malic enzyme from Escherichia coli W contains 7 cysteinyl residues per enzyme subunit. The reactivity of sulfhydryl (SH) groups of the enzyme was examined using several SH reagents, including 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB) and N-ethylmaleimide (NEM). 1. Two SH groups in the native enzyme subunit reacted with DTNB (or NEM) with different reaction rates, accompanied by a complete loss of the enzyme activity. The second-order modification rate constant of the "fast SH group" with DTNB coincided with the second-order inactivation rate constant of the enzyme by the reagent, suggesting that modification of the "fast SH group" is responsible for the inactivation. When the enzyme was denatured in 4 M guanidine HCl, all the SH groups reacted with the two reagents. 2. Althoug the inactivation rate constant was increased by the addition of Mg2+, an essential cofactor in the enzyme reaction, the modification rate constant of the "fast SH group" was unaffected. The relationship between the number of SH groups modified with DTNB or NEM and the residual enzyme activity in the absence of Mg2+ was linear, whereas that in the presence of Mg2+ was concave-upwards. These results suggest that the Mg2+-dependent increase in the inactivation rate constant is not the result of an increase in the rate constant of the "fast FH group" modification. 3. The absorption spectrum of the enzyme in the ultraviolet region was changed by addition of Mg2+. The dissociation constant of the Mg2+-enzyme complex obtained from the Mg2+- dependent increment of the difference absorption coincided with that obtained from the Mg2+- dependent enhancement of NEM inactivation. 4. Both the inactivation rate constant and the modification rate constant of the "fast SH group" were decreased by the addition of NADP+. The protective effect of NADP+ was increased by the addition of Mg2+. Based on the above results, the effects of Mg2+ on the SH-group modification are discussed from the viewpoint of conformational alteration of the enzyme.  相似文献   

5.
Tetrahedral anions bind to a phosphate binding site of 6-phosphogluconate dehydrogenase from Candida utilis, inhibit the enzyme competitively with the 6-phosphogluconate, decrease the reactivity of the SH groups, and mimic the protective effect of 6-phosphogluconate against some inactivating agents. The reaction of the enzyme with butanedione results in the inactivation of the enzyme associated with the modification of a single arginine residue per subunit. This arginine residue may be involved in the binding of the phosphate to the enzyme. Inactivation of the enzyme, upon reaction with permanganate, appears to be due to the oxidation to cysteic acid of a single cysteine residue per enzyme subunit. The reaction of the enzyme with either periodate or hexachloroplatinate causes the loss of the catalytic activity. This inactivation, due to an affinity labeling, is correlated with the oxidation of two SH groups per subunit to an S-S bridge. Photoinactivation of the enzyme by pyridoxal 5′-phosphate is also restricted to the active site of the enzyme. The lysine and the histidine residues involved in this photoinactivation should thus be in the vicinity of the phosphate binding site.  相似文献   

6.
4-Aminobutyrate aminotransferase is inactivated by preincubation with iodosobenzoate at pH 7. The reaction of 2 SH residues/dimer resulted in formation of an oligomeric species of Mr = 100,000 detectable by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The subunits cross-linked via a disulfide bond are dissociated by addition of 2-mercaptoethanol which also restores full catalytic activity (Choi, S. Y., and Churchich, J.E. (1985) J. Biol. Chem. 260, 993-997). The substrate 2-oxoglutarate prevents inactivation of the enzyme by iodosobenzoate and the subsequent formation of one disulfide bond, whereas 4-aminobutyrate has no effect on the reactivity of SH groups with iodosobenzoate. Modified 4-aminobutyrate aminotransferase (containing 1 disulfide bond) catalyzes a half-transamination reaction; but it is unable to react with 2-oxoglutarate to generate the aldimine form of the enzyme. The spectroscopic properties (fluorescence yield and polarization of fluorescence) of PMP bound to the modified enzyme are different from those of pyridoxamine phosphate (PMP) bound to the native enzyme. The polarization of fluorescence values of PMP bound to the cross-linked enzyme, excited over the spectral range 310-370 nm, are greater (25%) than those of the cofactor of the native enzyme. An increase in the polarization values implies that the motion of PMP is restricted when the subunits are cross-linked via a disulfide bond.  相似文献   

7.
The content of free SH groups and disulfide bonds in the purified pig kidney Na+,K+-ATPase was determined by ammetric titration with silver nitrate. In the native enzyme, most of the free SH groups are masked due to their location in the polypeptide chain regions poorly accessible to SH reagents. Denaturation with 5% SDS and 8 M urea makes these regions accessible thus revealing 22 free SH groups/mol of the protein. After complete blocking of free SH groups with silver ions, 8 SH groups/mol of the protein are being released upon sulfitolysis which indicates the presence of four disulfide bonds in the enzyme. At least one disulfide bridge is located in the alpha-subunit whereas the beta-subunit contains three disulfide bonds.  相似文献   

8.
The production of class A beta-lactamases is a major cause of clinical resistance to beta-lactam antibiotics. Some of class A beta-lactamases are known to have a disulfide bridge. Both narrow spectrum and extended spectrum beta-lactamases of TEM and the SHV enzymes possess a disulfide bond between Cys77 and Cys123, and the enzymes with carbapenem-hydrolyzing activity have a well-conserved disulfide bridge between Cys69 and Cys238. We produced A77C/G123C mutant of the extended-spectrum beta-lactamase Toho-1 in order to introduce a disulfide bond between the cysteine residues at positions 77 and 123. The result of 5,5'-dithiobis-2-nitrobenzoic acid (DTNB) titrations confirmed formation of a new disulfide bridge in the mutant. The results of irreversible heat inactivation and circular dichroism (CD) melting experiments indicated that the disulfide bridge stabilized the enzyme significantly. Though kinetic analysis indicated that the catalytic properties of the mutant were quite similar to those of the wild-type enzyme, E. coli producing this mutant showed drug resistance significantly higher than E. coli producing the wild-type enzyme. We speculate that the stability of the enzymes provided by the disulfide bond may explain the wide distribution of TEM and SHV derivatives and explain how various mutations can cause broadened substrate specificity without loss of stability.  相似文献   

9.
The stoichiometry and the reactivity of the sulfhydryl groups of a human liver acid phosphatase have been studied. The smallest (Mr = 14,400) of the three molecular-weight forms of acid phosphatase from human liver, recently purified and characterized in our laboratory, was treated with various sulfhydryl group-specific reagents: p-hydroxymercuribenzoate, p-hydroxymercuriphenylsulfonate, fluorescein mercuriacetate, methyl methanethiosulfonate, p-nitrophenoxycarbonyl methyl disulfide, and thiosulfate. A total loss of enzymatic activity was obtained in each case. By spectrophotometric titration with 5,5′-dithiobis(2-nitrobenzoate) and p-hydroxymercuriphenylsulfonate it was shown that there are six free sulfhydryls per protein molecule, consistent with the amino acid analysis of this enzyme. The same number was deduced as a result of inactivation studies carried out with p-hydroxymercuribenzoate and p-hydroxymercuriphenylsulfonate. A total loss of activity was obtained at reagent to enzyme ratios of 6:1 in both cases. Similar results were obtained upon inactivation by p-nitrophenoxycarbonyl methyl disulfide, where the enzyme was found to possess only 10% residual activity at an inhibitor-to-enzyme ratio of 6:1. With fluorescein mercuriacetate as an inactivator, total loss of activity was found at a 2.5 times molar excess of this reagent over protein. Both the stoichiometry of inactivation and fluorescence titration experiments suggest that fluorescein mercuriacetate can function as a bifunctional sulfhydryl group reagent. The activity of a totally inactivated enzyme preparation obtained following reaction with excess of p-nitrophenoxycarbonyl methyl disulfide or with methyl methanethiolsulfonate could be almost completely restored upon treatment with dithiothreitol. These data are consistent with the interpretation that in each enzyme molecule, there are six free sulfhydryl groups of almost equal reactivity, at least one of which is essential for enzymatic activity.  相似文献   

10.
Escherichia coli mannitol specific EII in membrane vesicles can be inhibited by the action of the oxidizable substrate-reduced phenazine methosulfate (PMS) in a manner similar to E. coli enzyme IIGlc [Robillard, G. T., & Konings, W. (1981) Biochemistry 20, 5025-5032]. The fact that reduced PMS and various oxidizing agents protect the enzyme from inactivation by the sulfhydryl reagents N-ethylmaleimide and bromopyruvate suggests that the active form possesses a dithiol which can be protected by conversion to a disulfide. The sulfhydryl-disulfide distribution has been examined in purified EIImtl by labeling studies with N-[1-14C]ethylmaleimide ( [14C]NEM). EIImtl can be alkylated at three positions per peptide chain. When alkylation takes place in 8 M urea, only two positions are labeled. The third position becomes labeled in urea only after treatment with DTT, suggesting that the native enzyme is composed of two subunits linked by a disulfide bridge. The remaining two sulfhydryl groups per peptide chain appear to undergo changes in oxidation state as indicated by the following results. (1) Treatment of the active enzyme with NEM leads to complete inactivation and incorporation of 1 mol of [14C]NEM per peptide chain. Oxidizing agents protect the activity and prevent labeling presumably by forming a disulfide. (2) Phosphorylating the enzyme (one phosphoryl group per peptide chain) fully protects the activity, but 1 mol of NEM per peptide chain is still incorporated. Subsequent dephosphorylation by adding mannitol causes a second mole of [14C]NEM to be incorporated and results in complete inactivation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Alliinase, an enzyme found in garlic, catalyzes the synthesis of the well-known chemically and therapeutically active compound allicin (diallyl thiosulfinate). The enzyme is a homodimeric glycoprotein that belongs to the fold-type I family of pyridoxal-5′-phosphate-dependent enzymes. There are 10 cysteine residues per alliinase monomer, eight of which form four disulfide bridges and two are free thiols. Cys368 and Cys376 form a S—S bridge located near the C-terminal and plays an important role in maintaining both the rigidity of the catalytic domain and the substrate-cofactor relative orientation. We demonstrated here that the chemical modification of allinase with the colored —SH reagent N-(4-dimethylamino-3,5-dinitrophenyl) maleimide yielded chromophore-bearing peptides and showed that the Cys220 and Cys350 thiol groups are accesible in solution. Moreover, electron paramagnetic resonance kinetic measurements using disulfide containing a stable nitroxyl biradical showed that the accessibilities of the two —SH groups in Cys220 and Cys350 differ. Neither enzyme activity nor protein structure (measured by circular dichroism) were affected by the chemical modification of the free thiols, indicating that alliinase activity does not require free —SH groups. This allowed the oriented conjugation of alliinase, via the —SH groups, with low- or high-molecular-weight molecules as we showed here. Modification of the alliinase thiols with biotin and their subsequent binding to immobilized streptavidin enabled the efficient enzymatic production of allicin.  相似文献   

12.
The contribution of disulfide bridges to the thermostability of a type A feruloyl esterase (AuFaeA) from Aspergillus usamii E001 was studied by introducing an extra disulfide bridge or eliminating a native one from the enzyme. MODIP and DbD, two computational tools that can predict the possible disulfide bridges in proteins for thermostability improvement, and molecular dynamics (MD) simulations were used to design the extra disulfide bridge. One residue pair A126-N152 was chosen, and the respective amino acid residues were mutated to cysteine. The wild-type AuFaeA and its variants were expressed in Pichia pastoris GS115. The temperature optimum of the recombinant (re-) AuFaeAA126C-N152C was increased by 6°C compared to that of re-AuFaeA. The thermal inactivation half-lives of re-AuFaeAA126C-N152C at 55 and 60°C were 188 and 40 min, which were 12.5- and 10-folds longer than those of re-AuFaeA. The catalytic efficiency (kcat/Km) of re-AuFaeAA126C-N152C was similar to that of re-AuFaeA. Additionally, after elimination of each native disulfide bridge in AuFaeA, a great decrease in expression level and at least 10°C decrease in thermal stability of recombinant AuEaeA variants were also observed.  相似文献   

13.
An enzyme widely distributed in rabbit tissues which catalyzes an interchange between N,N-di-dinitrophenyl-L-cystine and oxidized glutathione to form the mixed disulfide is described. D-Penicillamine disulfide can be substituted for oxidized glutathione and the mixed disulfide of cysteine and glutathione can serve as the sole substrate giving as one product of interchange, oxidized glutathione. The enzyme is very labile and only limited purification of it has been achieved. The activity increases with increasing pH above 6.6, the Km for N,N-di-dinitrophenyl-L-cystine is 0.2 mM and for oxidized glutathione 0.8 mM. The enzyme is inhibited by SH reagents with protection against iodoacetamide inactivation provided by N,N-di-dinitrophenyl-L-cystine. Evidence is presented that disulfide-disulfide interchange enzyme is a different activity from the previously described protein disulfide isomerase and thiol transferase.  相似文献   

14.
The apoenzyme of diol dehydrase was inactivated by four sulfhydryl-modifying reagents, p-chloromercuribenzoate, 5,5′-dithiobis(2-nitrobenzoate) (DTNB), iodoacetamide, and N-ethylmaleimide. In each case pseudo-first-order kinetics was observed. p-Chloromercuribenzoate modified two sulfhydryl groups per enzyme molecule and modification of the first one resulted in complete inactivation of the enzyme. DTNB also modified two sulfhydryl groups, but modification of the second one essentially corresponded to the inactivation. In both cases, the inactivation was reversed by incubation with dithiothreitol. Cyanocobalamin, a potent competitive inhibitor of adenosylcobalamin, protected the essential residue, but not the nonessential one, against the modification by these reagents. By resolving the sulfhydryl-modified cyanocobalamin-enzyme complex, the enzyme activity was recovered, irrespective of treatment with dithiothreitol. From these results, we can conclude that diol dehydrase has two reactive sulfhydryl groups, one of which is essential for catalytic activity and located at or in close proximity to the coenzyme binding site. The other is nonessential for activity. Neitherp-chloromercuribenzoate- nor DTNB-modified apoenzyme was able to bind cyanocobalamin, whereas the iodoacetamide- and N-ethylmaleimide-modified apoenzyme only partially lost the ability to bind cyanocobalamin. The inactivation of diol dehydrase by p-chloromercuribenzoate and DTNB did not bring about dissociation of the enzyme into subunits. Total number of the sulfhydryl groups of this enzyme was 14 when determined in the presence of 6 m guanidine hydrochloride. No disulfide bond was detected.  相似文献   

15.
Bacterial and mammalian pyruvate and 2-oxoglutarate dehydrogenase complexes undergo an irreversible inactivation upon accumulation of the dihydrolipoate intermediate. The first component of the complexes, 2-oxoacid dehydrogenase, is affected. Addition of thioredoxin protects from this inactivation, increasing catalytic rates and limiting degrees of the substrate transformation to products, acyl-CoA and NADH. Although the redox active cysteines of thioredoxin are essential for its interplay with the complexes, the effects are observed with both dithiol and disulfide forms of the protein. This indicates that thioredoxin affects an SH/S–S component of the system, which is present in the two redox states. The complex-bound lipoate is concluded to be the thioredoxin target, since (i) both dithiol and disulfide forms of the residue are available during the catalytic cycle and (ii) the thioredoxin reaction with the essential SH/S–S group of the terminal component of the complex, dihydrolipoyl dehydrogenase, is excluded. Thus, the thioredoxin disulfide interacts with the dihydrolipoate intermediate, while the thioredoxin dithiol reacts with the lipoate disulfide. Kinetic consequences of such interplay are consistent with the observed thioredoxin effects. Owing to the essential reactivity of the SH/S–S couple in thioredoxin, the thiol–disulfide exchange between thioredoxin and the lipoate residue is easy reversible, providing both protection (by the mixed disulfide formation) and catalysis (by the appropriate lipoate release). In contrast, non-protein SH/S–S compounds prevent the inactivatory action of dihydrolipoate intermediate only at a high excess over the complex-bound lipoate. This interferes with the catalysis-required release of the residue from its mixed disulfide. Therefore, only thioredoxin is capable to ‘buffer' the steady-state concentration of the reactive dithiol. Such action represents a new thioredoxin function, which may be exploited to protect other enzymes with exposed redox-active thiol intermediates.  相似文献   

16.
Aspartate transaminase (EC 2.6.1.1) from chicken heart cytosol contains 4 thiol groups per subunit. Two of them are fully buried. One exposed SH group is readily modified by iodoacetamide, N-ethylmaleimide, tetranitromethane, 5,5′-dithio-bis(2-nitrobenzoate), 4,4′-dipyridyl disulfide and p-mercuribenzoate. A further SH group is semi-buried: while inaccessible for alkylating reagents and disulfides, it can be blocked by p-mercuribenzoate at pH about 5 (but not at pH 8). Treatment of the enzyme with tetranitromethane in the absence of substrates leads to nitration of maximally 0.8 tyrosine residue per subunit; in the presence of amino and keto substrate 1.65 eq of nitrotyrosine is formed, with a moderate decrease of enzymic activity.  相似文献   

17.
1. (Na+ +K+)-ATPase from rectal gland of Squlus acanthias contains 34 SH groups per mol (Mr 265000). 15 are located on the α subunit (Mr 106 000) and two on the β subunit (Mr 40 000). The β subunit also contains one disulphide bridge. 2. The reaction of (Na+ +K+)-ATPase with N-ethylmaleimide shows the existence of at least three classes of SH groups. Class I contains two SH groups on each α subunit and one on each β subunit. Reaction of these groups with N-methylmaleimide in the presence of 40% glycerol or sucrose does not alter the enzyme activity. Class II contains four SH groups on each α subunit, and the reaction of these groups with 0.1 mM N-ethylmaleimide in the presence of 150 mM K+ leads to an enzyme species with about 16% activity. The remaining enzyme activity can be completely abolished by reaction with 5–10 nM N-ethylmaleimide, indicating a third class of SH groups (Class III). This pattern of inactivation is different from that of the kidney enzyme, where only one class of SH groups essential to activity is observed. 3. It is also shown that N-ethylmaleimide and DTNB inactivate by reacting with the same Class II SH groups. 4. Spin-labelling of the (Na+ +K+)-ATPase with a maleimide derivative shows that Class II groups are mostly buried in the membrane, whereas Class I groups are more exposed. It is also shown that spin label bound to the Class I groups can monitor the difference between the Na+- and K+-forms of the enzyme.  相似文献   

18.
The soybean urease (urea amidohydrolase; EC 3.5.1.5) was investigated to elucidate the presence of sulfhydryl (–SH) groups and their significance in urea catalysis with the help of various –SH group specific reagents. The native urease incubated with 5,5′-dithiobis (2-nitrobenzoic acid) (DTNB) showed exponential increase in the absorbance, thereby revealing the presence of –SH groups. A total of 34 –SH groups per hexamer enzyme molecule were estimated from the absorption studies which represents nearly six –SH groups per subunit. The time-dependent inactivation of urease with DTNB, p-chloromercuribenzoate (p-CMB), N-ethylmaleimide (NEM) and iodoacetamide (IAM) showed biphasic kinetics, where half of the enzyme activity was lost more rapidly than the other half. This study reveals the presence of two categories of “accessible” –SH groups, one category being more reactive than the other. The inactivation of urease by p-CMB was largely reversed on treatment with cysteine, which might be due to unblocking of –SH group by mercaptide exchange reaction. Finally, when NEM inactivated urease was incubated with sodium fluoride, a time-dependent regain of activity was observed with higher concentrations of fluoride ion.  相似文献   

19.
Irreversible inactivation of rat liver tyrosine aminotransferase   总被引:2,自引:0,他引:2  
Homogenates prepared from rat livers irreversibly inactivate tyrosine aminotransferase, both endogenous and purified exogenous enzyme, in the presence of certain compounds which bind to pyridoxal 5′-P. The rate of inactivation ranged from a half-life of 0.72 to greater than 15 hr. The pyridoxal 5′-P binding compounds may be considered to be structural analogs for α-ketoglutarate or l-tyrosine, both of which are substrates for the enzyme. l-Cysteine and l-DOPA are the most effective compounds tested of each of the two structural analog classes, respectively. Absence of the carboxyl group from l-cysteine or l-DOPA has little effect on the half-life of the enzyme, whereas absence or substitution of the amino group results in an increased enzyme half-life. Absence of the —SH group from l-cysteine or of the 3′-OH group from l-DOPA results in little or no inactivation of the enzyme (t12 increased to greater than 15 hr). Semicarbazide and hydroxylamine have little effect on the stability of the enzyme. Addition of pyridoxal 5′-P to homogenates incubated with l-cysteine or l-DOPA inhibits the inactivation of the enzyme. However, the addition of cofactor to inactivated enzyme does not restore lost activity.There is a disappearance of antigenic cross-reacting material during inactivation of the enzyme. This loss of specific cross-reacting material occurs at a slower rate than the loss of enzyme activity, indicating that enzymatic activity is lost prior to loss of antigenic recognition. A three-step proposal is presented to explain the data observed in which the first step is a reversible loss of pyridoxal 5′-P from the enzyme, followed by a specific irreversible inactivation of the enzyme, and ending with nonspecific proteolysis or degradation of the inactivated enzyme molecules.  相似文献   

20.
When air oxidized, partially inactivated rhodanese (EC 2.8.1.1) is treated with dithiothreitol (DTT) to regenerate the reduced essential sulfhydryl group there is an initial reactivation followed by an anomalous slower inactivation. Fully active enzyme shows only inactivation. The inactivated enzyme may be completely reactivated on long incubation with the substrate thiosulfate ion. None of the normal potentialities of DTT appear to be responsible for the inactivation. The results are interpreted in terms of disulfide formation between DTT and an essential enzymic sulfhydryl group with the resulting complex being stabilized by secondary interactions which are particularly favorable due to similarities between DTT and lipoic acid--a normal sulfur acceptor substrate.  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