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.
1. Benzofuroxan (benzofurazan 1-oxide, benzo-2-oxa-1,3-diazole N-oxide) was evaluated as a specific chromophoric oxidizing agent for thiol groups. 2. Aliphatic thiol groups both in low-molecular-weight molecules and in the enzymes papain (EC 3.4.22.2), ficin (EC 3.4.22.3) and bromelain (EC 3.4.22.4) readily reduce benzofuroxan to o-benzoquinone dixime; potential competing reactions of amino groups are negligibly slow. 3. The fate of the thiol depends on its structure: a mechanism is proposed in which the thiol and benzofuroxan form an adduct which, if steric factors permit, reacts with another molecule of thiol to form a disulphide; when the thiol is located in the active site of a thiol proteinase and steric factors preclude enzyme dinner formation, the adduct reacts instead with water or HO- to form a sulphenic acid; attack on the sulphur atom of the adduct by either a sulphur or oxygen nucleophile releases o-benzoquinone dioxine. 4. Benzofuroxan contains n o proton-binding sites with pKa values in the range 3-10 and probably none in the range 0-14; o-benzoquinone dioxine undergoes a one-proton ionization with pKa=6.75.5. o-benzoquinone dioxime absorbs strongly at wavelengths greater than 410nm, where absorption by benzofuroxan, proteins and simple thiol compounds is negligible; 416 nm is an isosbestic point (epsilon 416 = 5110 litre. mol-1-cm-1); epsilon430=3740+[1460/(1+[H+]/Ka)] where pKa=6.75. 6. The possibility of acid-base catalysis of the oxidation by active-centre histidine residues of the thiol proteinases is discussed.  相似文献   

2.
1.2,2'-Dipyridyl disulphide (2-Py-S-S-2-Py) and n-propyl 2-pyridyl disulphide (propyl-S-S-2-Py) were used as two-protonic-state reactivity probes to investigate the active centre of papain (EC 3.4.22.2).2. The existence of a striking rate optimum at pH approx. 4 in the reaction of papain not only with the symmetrical probe but also with the unsymmetrical probe is shown to constitute compelling evidence that the thiolate ion component of the cysteine-25-histidine-159 interactive system of papain possesses appreciable nucleophilic character. It is not a necessary requirement that the probe reagent should engage the imidazolium ion of histidine-159 in hydrogen-bonding for the sulphur atom of the interactive system to display nucleophilic character. The single proton-binding site of propyl-S-S-2-Py cannot simultaneously interrupt the active-centre ion pair and provide for rate enhancement as the pH is lowered towards 4. The possible implication of this for the mechanism of papain-catalysed hydrolysis is discussed. 3. The suspected difference in the active centres of papain and ficin (EC 3.4.22.3), which could be a lack in ficin of a carboxy group conformationally equivalent to that of aspartic acid-158 of papain is confirmed. The reactivity of the papain thiol group towards both probe reagents is controlled by two ionizations with pKa close to 4 that are positively co-operative. 4. In the reaction of papain with 2-Py-S-S-2-Py. the reactivity appears to be controlled also by an addition ionization with pKa approx. 5. Possible origins of this additional ionization are discussed. K. The spectral and ionization characteristics of propyl-S-S-2-Py are reported. 6. The reagent reacts rapidly with thiol groups at the sulphur atom distal from the pyridyl ring to provide, at pH values below 9, stoicheiometric release of 2-thiopyridone. This property, together with the ability of the reagent markedly to increase its electrophilicity consequent on protonation, suggests alkyl-2-pyridyl disulphides in general as valuable two-protonic-state reactivity probes with exceptional specificity for thiol groups.  相似文献   

3.
The active centres of chymopapains A and B (jointly designated EC 3.4.22.6) and papaya (Carica papaya L.) peptidase A were investigated by using 2,2'-dipyridyl disulphide and 5,5'-dithiobis-(2-nitrobenzoic acid) as thiol-specific reactivity probes. Whereas the first active-centre pKa values for chymopapain B and papaya peptidase A are less than 5, is as the case for papain (EC 3.4.22.2) and ficin (EC 3.4.22.3), that for chymopapain A is about 6.8. The reason why the reactions of thiols of pKa approx. 6.5 with 2.2'-dipyridyl disulphide are essentially pH-independent in the pH range around the thiol pKa is delineated. The value of the Brønsted coefficient (beta nuc.) for the reactions of thiolate ions with the 2,2'-dipyridyl disulphide monocation appears to be smaller than its value for the corresponding reactions with the neutral disulphide.  相似文献   

4.
The second-order rate constants (k) for the reactions of 2,2'-dipyridyl disulphide (pKa2,45) with 2-mercaptoethanol (pKa9.6) and with benzimidazol-2-ylmethanethiol (pKa values 5.6 and 8.3) were determined at 25 degrees C at I 0.1 by stopped-flow spectral analysis over a wide range of pH. These were used to calculate the pH-independent second-order rate constants (k) for the reactions of neutral 2,2'-dipyridyl disulphide and of its monocation with the 2-mercaptoethanol thiolate anion (associated pKa9.6) and with the benzimidazol-2-ylmethanethiol zwitterion (associated pKa5.6). For both thiolate ions, the rate-enhancement factor (kmonocation/kneutral disulphide) is about 1.5x10(3). The dependence on pH in acidic media of k for the reaction of 2,2'-dipyridyl disulphide with actinidin, the thiol proteinase from Actinidia chinensis, was shown to differ from the forms of pH-dependence observed for the analogous reactions with papain (EC 3.4.22.2) and ficin (3.4.22.3). The reactivity of the 2,2'-dipyridyl disulphide dication and its apparent sensitivity to the presence and location of a positive charge in the attacking thiol are discussed.  相似文献   

5.
Benzofuroxan reacts with the catalytic-site thiol group of actinidin (EC 3.4.22.14, the cysteine proteinase from Actinidia chinensis) to produce stoicheiometric amounts of the chromophoric reduction product, o-benzoquinone dioxime, and of a catalytically inactive derivative of actinidin that is devoid of thiol and that is assumed to contain, initially at least, the sulphenic acid of cysteine-25. A similar result applies also to papain (EC 3.4.22.2). The rate of o-benzoquinone dioxime formation is neither increased by inclusion of 2-mercaptoethanol or hydroxylamine in the reaction mixture nor decreased by changing the solvent from H2O to 2H2O. The change of solvent was shown to be without effect also on the rate of reaction of benzofuroxan with papain. These results suggest that the reactions of benzofuroxan with both actinidin and papain involve rate-determining attack of the catalytic-site thiol group to produce an intermediate adduct that then reacts rapidly with water to form enzyme sulphenic acid and o-benzoquinone dioxime. The pH-dependence of the second-order rate constant for the reaction of benzofuroxan with actinidin was determined in the pH range 4.3-10.2. In marked contrast with the analogous reaction of papain (reported by Shipton & Brocklehurst [(1977) Biochem. J. 167, 799-810] ) the pH-k profile for the actinidin reaction clearly contains a sigmoidal component with pKa 5.5, in which k increases with decreasing pH. These data together with the molecular pKa values for S-/ImH+ ion-pair formation and decomposition (3.0 and 9.6) suggest that the combined nucleophilic-electrophilic reactivity of the ion-pair of actinidin might be controlled by the state of ionization of another ionizing group, associated with the molecular pKa of 5.5. The pH-dependence of k for the reaction of actinidin with benzofuroxan at 25 degrees C at I 0.1 in aqueous buffers containing 6.7% (v/v) ethanol is probably adequately described by: k = k1/(1 + [H+]/KI + KII/[H+]) + k2/(1 + [H+]/KII + KIII/ [H+] + k3/(1 + [H+]/KIII) in which kI = 2.55 M -1 X s -1, k2 = 1.35 M -1, k3 = 0.93 M -1 X s -1, pKI = 3.0, pKII = 5.5 and pKIII = 9.6. By contrast, the analogous reaction of papain may be described by the same equation but with kI = 0, k2 = 2.2 M -1 X s -1, k3 = 1.3 M -1 X s -1, pKII = 3.6 and pKIII = 9.0.  相似文献   

6.
The characteristics of actinidin (EC 3.4.22.14) and papain (EC 3.4.22.2), two cysteine proteinases whose catalytic-site regions appear to superimpose to a degree that approaches atomic co-ordinate accuracy of both crystal structures, were evaluated by determining (a) the pH-dependence in acid media of the acylation process of the catalytic act (k+2/Ks) using N alpha-benzoyl-L-arginine p-nitroanilide (L-Bz-Arg-Nan) as substrate and (b) the sensitivity of the reactivity of the catalytic-site thiol group and its pH-dependence to structural change in small, thiol-specific, two-protonic-state reactivity probes (2,2'-dipyridyl disulphide and methyl 2-pyridyl disulphide) where enzyme-probe contacts should be restricted to areas close to the catalytic site. Distortion of the catalytic sites of the two enzymes at pH less than 4 was evaluated over time-scales appropriate for both stopped-flow reactivity probe kinetics (less than or equal to 1-2 s) and steady-state substrate catalysis kinetics (3-5 min) by using the 2,2'-dipyridyl disulphide monocation as a titrant for non-distorted catalytic sites. This permitted a lower pH limit to be defined for valid kinetic analysis of both types. The behaviour of the enzymes at pH less than 4 requires a kinetic model in which the apparently biomolecular reaction of enzyme with probe reagent is separated from the process leading to loss of conformational integrity by a potentially reversible step. The acylation of actinidin with L-Bz-Arg-Nan in acidic media occurs in two protonic states, one produced by raising the pH across pKa less than 4 which probably characterizes the formation of -S-/-ImH+ ion pair (pKa approx. 3) and the other, of higher reactivity, produced by raising the pH across pKa 5.5, which may characterize rearrangement of catalytic-site geometry. The pH-dependence of the acylation of papain by L-Bz-Arg-Nan is quite different and is not influenced by protonic dissociation with pKa values in the range 5-6. The earlier conclusion that the acylation of papain depends on two protonic dissociations each with pKa approx. 4 was confirmed. This argument is now more firmly based because titration with 2,2'-dipyridyl disulphide permits the loss of conformational integrity to be taken into account in the analysis of the kinetic data at very low pH. Methyl 2-pyridyl disulphide was synthesized by reaction of pyridine-2-thione with methyl methanethiolsulphonate and its pKa at I = 0.1 was determined by spectral analysis at 307 nm to be 2.8.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
1. 4-Chloro-7-nitrobenzo-2-oxa-1,3-diazole (Nbd chloride) was used as a reactivity probe to characterize the active centres of papin (EC 3.4.22.2), ficin (EC 3.4.22.3) and bromelain (EC 3.4.22.4). 2. In the pH range 0-8 Nbd chloride probably exists mainly as a monocation, possibly with the proton located on N-1 of the oxadiazole ring. 3. Spectroscopic evidence is presented for the intermediacy of Meisenheimer-type adducts in the reaction of Nbd chloride with nucleophiles. 4. The pH-dependence of the second-order rate constants (k) of the reactions of the three enzymes with Nbd chloride was determined at 25 degrees C, I = 0.1 mol/litre in 6.7% (v/v) ethanol in the pH range 2.5-5, where, at least for papain and ficin, the reactions occur specifically with their active-centre thiol groups. The pH-k profile for the papain reaction is bell-shaped (pKaI = 3.24, pKaII = 3.44 and k = 86M(-1)-s(-1), whereas that for ficin is sigmoidal (pKa = 3.6, k = 0.36M(-1)-s(-1), the rate increasing with increasing pH. The profile for the bromelain reaction appears to resemble that for the ficin reaction, but is complicated by amino-group labelling. 5. The bell-shaped profile of the papain reaction is considered to arise from the reaction of the thiolate ion of cysteine-25, maintained in acidic media by interaction with the side chain of histidine-159, with the Nbd chloride monocation hydrogen-bonded at its nitro group to the un-ionized form of the carboxyl group of aspartic acid-158. The lack of acid catalysis in the corresponding reactions of ficin and probably of bromelain suggests that these enzymes may lack carboxyl groups conformationally equivalent to that of aspartic acid-158 of papain. The possible consequences of this for the catalytic sites of these enzymes is discussed.  相似文献   

8.
1. Fully active ficin (EC 3.4.22.3) containing 1 mol of thiol with high reactivity towards 2,2'-dipyridyl disulphide (2-Py-S-S-2-Py) at pH4.5 per mol of protein was prepared from the dried latex of Ficus glabrata by covalent chromatography on a Sepharose-glutathione-2-pyridyl disulphide gel. 2. Ficin thus prepared is a mixture of ficins I-IV and ficin G, in which ficins II and III predominate. The various ficins exhibit similar reactivity characteristics towards 2-Py-S-S-2-Py. 3. Use of 2-Py-S-S-2-Py as a reactivity probe demonstrates (a) that in ficin, as in papain (EC 3.4.22.2), the active-centre thiol and imidazole groups interact to provide a nucleophilic state at pH values of approx. 6 additional to the uncomplicated thiolate ion that predominates at pH values over 9, and (b) a structural difference between ficin and papain that leads to a much higher rate of reaction of 2-Py-S-S-2-Py with ficin than with papain at pH values 3-4. This difference is suggested to include a lack in ficin of a carboxyl group conformationally equivalent to that of aspartic acid-158 in papain. 4. The high electrophilicity of the 2-Py-S-S-2PyH+ monocation allows directly the detection of the exposure of the buried thiol group of ficin at pH values below 4.  相似文献   

9.
Li L  Li Z  Wang C  Xu D  Mariano PS  Guo H  Dunaway-Mariano D 《Biochemistry》2008,47(16):4721-4732
L-arginine deiminase (ADI) catalyzes the hydrolysis of L-arginine to form L-citrulline and ammonia via two partial reactions. A working model of the ADI catalytic mechanism assumes nucleophilic catalysis by a stringently conserved active site Cys and general acid-general base catalysis by a stringently conserved active site His. Accordingly, in the first partial reaction, the Cys attacks the substrate guanidino C zeta atom to form a tetrahedral covalent adduct, which is protonated by the His at the departing ammonia group to facilitate the formation of the Cys- S-alkylthiouronium intermediate. In the second partial reaction, the His activates a water molecule for nucleophilic addition at the thiouronium C zeta atom to form the second tetrahedral intermediate, which eliminates the Cys in formation of the L-citrulline product. The absence of a basic residue near the Cys thiol suggested that the electrostatic environment of the Cys thiol, in the enzyme-substrate complex, stabilizes the Cys thiolate anion. The studies described in this paper explore the mechanism of stabilization of the Cys thiolate. First, the log(k(cat)/K(m)) and log k(cat) pH rate profiles were measured for several structurally divergent ADIs to establish the pH range for ADI catalysis. All ADIs were optimally active at pH 5, which suggested that the Cys pKa is strongly perturbed by the prevailing electrostatics of the ADI active site. The p K a of the Bacillus cereus ADI (BcADI) was determined by UV-pH titration to be 9.6. In contrast, the pKa determined by iodoacetamide Cys alkylation is 6.9. These results suggest that the negative electrostatic field from the two opposing Asp carboxylates perturbs the Cys pKa upward in the apoenzyme and that the binding of the iodoacetamide (a truncated analogue of the citrulline product) between the Cys thiol and the two Asp carboxylates shields the Cys thiol, thereby reducing its pKa. It is hypothesized that the bound positively charged guanidinium group of the L-arginine substrate further stabilizes the Cys thiolate. The so-called "substrate-assisted" Cys ionization, first reported by Fast and co-workers to operate in the related enzyme dimethylarginine dimethylaminohydrolase [Stone, E. M., Costello, A. L., Tierney, D. L., and Fast, W. (2006) Biochemistry 45, 5618-5630], was further explored computationally in ADI by using an ab initio quantum mechanics/molecular mechanics method. The energy profiles for formation of the tetrahedral intermediate in the first partial reaction were calculated for three different reaction scenarios. From these results, we conclude that catalytic turnover commences from the active configuration of the ADI(L-arginine) complex which consists of the Cys thiolate (nucleophile) and His imidazolium ion (general acid) and that the energy barriers for the nucleophilic addition of Cys thiolate to the thiouronium C zeta atom and His imidazolium ion-assisted elimination from the tetrahedral intermediate are small.  相似文献   

10.
1. A convenient method of preparation of jack-bean urease (EC3.5.1.5) involving covalent chromatography by thiol-disulphide interchange is described. 2. Urease thus prepared has specific activity comparable with the highest value yet reported (44.5 +/- 1.47 kat/kg, Km = 3.32 +/- 0.05 mM; kcat. = 2.15 X 10(4) +/- 0.05 X 10(4)s-1 at pH7.0 and 38 degrees C). 3. Titration of the urease thiol groups with 2,2'-dipyridyl disulphide (2-Py-S-S-2-Py) and application of the method of Tsou Chen-Lu [(1962) Sci. Sin. 11, 1535-1558] suggests that the urease molecule (assumed to have mol.wt. 483000 and epsilon280 = 2.84 X 10(5) litre-mol-1-cm-1) contains 24 inessential thiol groups of relatively high reactivity (class-I), six 'essential' thiol groups of low reactivity (class-II) and 54 buried thiol groups (class-III) which are exposed in 6M-guanidinium chloride. 4. The reaction of the class-I thiol groups with 2-Py-S-S-2-Py was studied in the pH range 6-11 at 25 degrees C(I = 0.1 mol/l) by stopped-flow spectrophotometry, and the analogous reaction of the class-II thiol groups by conventional spectrophotometry. 5. The class-I thiol groups consist of at least two sub-classes whose reactions with 2-Py-S-S-2-Py are characterized by (a) pKa = 9.1, k = 1.56 X 10(4)M-1-s-1 and (b) pKa = 8.1, k = 8.05 X 10(2)M-1-s-1 respectively. The reaction of the class-II thiol groups is characterized by pKa = 9.15 and k = 1.60 X 10(2)M-1-s-1. 6. At pH values 7-8 the class-I thiol groups consist of approx. 50% class-Ia groups and 50% class-Ib groups. The ratio class Ia/class Ib decreases an or equal to approx. 9.5, and at high pH the class-I thiol groups consist of at most 25% class-Ia groups and at least 75% class-Ib groups. 7. The reactivity of the class-II thiol groups towards 2-Py-S-S-2-Py is insensitive to the nature of the group used to block the class-I thiols. 8. All the 'essential' thiol groups in urease appear to be eeactive only as uncomplicated thiolate ions. The implications of this for the active-centre chemistry of urease relative to that of the thiol proteinases are discussed.  相似文献   

11.
4-(N-Aminoethyl 4-pyridyl disulphide)-7-nitrobenzo-2-oxa-1,3-diazole was synthesized and evaluted as a two-protonic-state reactivity probe by kinetic study of its reactions with papain (EC 3.4.22.2) and with benzimidazol-2-ylmethanethiol. Evidence is presented to suggest that: (i) both this probe molecule and its 2-pyridyl isomer bind to papain; (ii) the binding is followed by a change in the environment of the thiol group of cysteine-25; (iii) the striking rate maximum in neutral media observed in the reaction of papain with the 2-pyridyl isomer but not with the 4-pyridyl isomer arises from association of the 2-pyridyl leaving group with the imidazolium ion of histidine-159.  相似文献   

12.
Benzofuroxan reacts with the catalytic-site thiol group of cathepsin B (EC 3.4.22.1) to produce stoichiometric amount of the chromophoric reduction product, o-benzoquinone dioxime. In a study of the pH-dependence of the kinetics of this reaction, most data were collected for the bovine spleen enzyme, but the more limited data collected for the rat liver enzyme were closely similar both in the magnitude of the values of the second-order rate constants (k) and in the shape of the pH-k profile. In acidic and weakly alkaline media, the reaction is faster than the reactions of benzofuroxan with some other cysteine proteinases. For example, in the pH region around 5-6, the reaction of cathepsin B is about 10 times faster than that of papain, 15 times faster than that of stem bromelain and 6 times faster than that of ficin. The pH-dependence of k for the reaction of cathepsin B with benzofuroxan was determined in the pH range 2.7-8.3. In marked contrast with the analogous reactions of papain, ficin and stem bromelain [reported by Shipton & Brocklehurst (1977) Biochem. J. 167, 799-810], the pH-k profile for the cathepsin B reaction contains a sigmoidal component with pKa 5.2 in which k increases with decrease in pH. This modulation of the reactivity of the catalytic-site -S-/-ImH+ ion-pair state of cathepsin B (produced by protonic dissociation from -SH/-ImH+ with pKa approx. 3) towards a small, rigid, electrophilic reagent, in a reaction that appears to involve both components of the ion-pair for efficient reaction, suggests that the state of ionization of a group associated with a molecular pKa of approx. 5 may control ion-pair geometry. This might account for the remarkable finding [reported by Willenbrock & Brocklehurst (1984) Biochem. J. 222, 805-814] that, although the ion-pair appears to be generated in cathepsin B as the pH is increased across pKa 3.4, catalytic competence is not generated until the pH is increased across pKa 5-6.  相似文献   

13.
A procedure for the isolation of cathepsin B (EC 3.4.22.1) and of cathepsin H from bovine spleen involving covalent chromatography by thiol-disulphide interchange and ion-exchange chromatography was devised. The stabilities of both cathepsins in alkaline media are markedly temperature-dependent, and reliable kinetic data can be obtained at pH values up to 8 by working at 25 degrees C with a continuous spectrophotometric assay. Both enzyme preparations contain only one type of thiol group as judged by reactivity characteristics towards 2,2'-dipyridyl disulphide at pH values up to 8; in each case this thiol group is essential for catalytic activity. Cathepsin H was characterized by kinetic analysis of the reactions of its thiol group with 2,2'-dipyridyl disulphide in the pH range approx. 2-8 and the analogous study on cathepsin B [Willenbrock & Brocklehurst (1984) Biochem. J. 222, 805-814] was extended to include reaction at pH values up to approx. 8. Cathepsin H, like the other cysteine proteinases, was shown to contain an interactive catalytic-site system in which the nucleophilic character of the sulphur atom is maintained in acidic media. The considerable differences in catalytic site characteristics detected by this two-protonic-state reactivity probe between cathepsin B, cathepsin H, papain (EC 3.4.22.2) and actinidin (EC 3.4.22.14) are discussed. Reaction with 2,2'-dipyridyl disulphide in acidic media, which is known to provide a rapid spectrophotometric active centre titration for many cysteine proteinases, is applicable to cathepsin H. This is useful because other active-centre titrations have proved unsuitable in view of the relatively low reactivity of the thiol group in cathepsin H.  相似文献   

14.
An important aspect of the catalytic mechanism of microsomal glutathione transferase (MGST1) is the activation of the thiol of bound glutathione (GSH). GSH binding to MGST1 as measured by thiolate anion formation, proton release, and Meisenheimer complex formation is a slow process that can be described by a rapid binding step (K(GSH)d = 47 +/- 7 mM) of the peptide followed by slow deprotonation (k2 = 0.42 +/- 0.03 s(-1). Release of the GSH thiolate anion is very slow (apparent first-order rate k(-2) = 0.0006 +/- 0.00002 s(-)(1)) and thus explains the overall tight binding of GSH. It has been known for some time that the turnover (kcat) of MGST1 does not correlate well with the chemical reactivity of the electrophilic substrate. The steady-state kinetic parameters determined for GSH and 1-chloro-2,4-dinitrobenzene (CDNB) are consistent with thiolate anion formation (k2) being largely rate-determining in enzyme turnover (kcat = 0.26 +/- 0.07 s(-1). Thus, the chemical step of thiolate addition is not rate-limiting and can be studied as a burst of product formation on reaction of halo-nitroarene electrophiles with the E.GS- complex. The saturation behavior of the concentration dependence of the product burst with CDNB indicates that the reaction occurs in a two-step process that is characterized by rapid equilibrium binding ( = 0.53 +/- 0.08 mM) to the E.GS- complex and a relatively fast chemical reaction with the thiolate (k3 = 500 +/- 40 s(-1). In a series of substrate analogues, it is observed that log k3 is linearly related (rho value 3.5 +/- 0.3) to second substrate reactivity as described by Hammett sigma- values demonstrating a strong dependence on chemical reactivity that is similar to the nonenzymatic reaction (rho = 3.4). Microsomal glutathione transferase 1 displays the unusual property of being activated by sulfhydryl reagents. When the enzyme is activated by N-ethylmaleimide, the rate of thiolate anion formation is greatly enhanced, demonstrating for the first time the specific step that is activated. This result explains earlier observations that the enzyme is activated only with more reactive substrates. Taken together, the observations show that the kinetic mechanism of MGST1 can be described by slow GSH binding/thiolate formation followed by a chemical step that depends on the reactivity of the electrophilic substrate. As the chemical reactivity of the electrophile becomes lower the rate-determining step shifts from thiolate formation to the chemical reaction.  相似文献   

15.
A K Knap  R F Pratt 《Proteins》1989,6(3):316-323
The RTEM-1 thiol beta-lactamase (Sigal, I.S., Harwood, B.G., Arentzen, R., Proc. Natl. Acad. Sci. U.S.A. 79:7157-7160, 1982) is inactivated by thiol-selective reagents such as iodoacetamide, methyl methanethiosulfonate, and 4,4'-dipyridyldisulfide, which modify the active site thiol group. The pH-rate profiles of these inactivation reactions show that there are two nucleophilic forms of the enzyme, EH2 and EH, both of which, by analogy with the situation with cysteine proteinases, probably contain the active site nucleophile in the thiolate form. The pKa of the active site thiol is therefore shown by the data to be below 4.0. This low pKa is thought to reflect the presence of adjacent functionality which stabilizes the thiolate anion. The low nucleophilicity of the thiolate in both EH2 and EH, with respect to that of cysteine proteinases and model compounds, suggests that the thiolate of the thiol beta-lactamase is stabilized by two hydrogen-bond donors. One of these, of pKa greater than 9.0, is suggested to be the conserved and essential Lys-73 ammonium group, while the identity of the other group, of pKa around 6.7, is less clear, but may be the conserved Glu-166 carboxylic acid. beta-Lactamase activity is associated with the EH2 form, and thus the beta-lactamase active site is proposed to contain one basic or nucleophilic group (the thiolate in the thiol beta-lactamase) and two acidic (hydrogen-bond donor) groups (one of which is likely to be the above-mentioned lysine ammonium group).  相似文献   

16.
1. Whereas the second-order rate constants for the reaction of the thiolate ion of 2-mercaptoethanol with 4,4'-dipyridyl disulphide (k4PDS) and with 5,5'-dithiobis-2-nitrobenzoate dianion increase with decreasing dielectric constant of the solvent, or remain unchanged, the rate constant for the analogous reaction with 2,2'-dipyridyl disulphide (k2PDS) decreases. This anomalous solvent effect and other unusual physicochemical properties of 2,2'-dipyridyl disulphide are discussed. 2. The differential effect of solvent on the reactions of thiolate ion with the 2,2'- and 4,4'-dipyridyl disulphides is shown to provide a method of characterizing solvent environments of thiol groups in proteins by a reactivity-probe method that should not suffer from the usual drawback associated with the existence of steric or binding effects of unknown magnitude. Application of the method to ficin (EC 3.4.22.3) suggests that its active-centre thiol group resides in a relatively hydrophobic environment. 3. The pH-k profile for the reaction of ficin with 4,4'-dipyridyl disulphide is reported.  相似文献   

17.
The dependence on thiol pK of the second-order rate constant (kS) for reaction of thiolate anions with MMTS was shown to follow the Br?nsted equation log kS = log G + beta pK with log G = 1.44 and 3.54 and beta = 0.635 and 0.309 for aryl and alkyl thiols, respectively. The reactivity toward MMTS of the protonated thiol group was found to be negligible in comparison to that of the thiolate anion. For 2-mercaptoethanol the reactivity toward MMTS of the protonated form of the thiol group was shown to be at least 5 X 10(9) smaller than that of the thiolate anion. The pH dependence of the second-order rate constant for reaction of the thiolate group of Cys-25 at the active site of papain was determined and shown to be consistent with the previously determined low pK for Cys-25 and its electrostatic interaction with His-159. The small dependence of the reactivity of Cys-25 on thiol pK (beta approximately 0.09) suggested that the charge-charge interactions that act through space to perturb the pK of the nucleophile at the active site of papain and perhaps other enzymes may serve to increase the fraction of nucleophile present in the reactive basic form without introducing the decrease in nucleophilic reactivity seen in model systems where pK's are lowered primarily by charge-dipole interactions.  相似文献   

18.
Fatty acid nitration by nitric oxide-derived species yields electrophilic products that adduct protein thiols, inducing changes in protein function and distribution. Nitro-fatty acid adducts of protein and reduced glutathione (GSH) are detected in healthy human blood. Kinetic and mass spectrometric analyses reveal that nitroalkene derivatives of oleic acid (OA-NO2) and linoleic acid (LNO2) rapidly react with GSH and Cys via Michael addition reaction. Rates of OA-NO2 and LNO2 reaction with GSH, determined via stopped flow spectrophotometry, displayed second-order rate constants of 183 M(-1)S(-1) and 355 M(-1)S(-1), respectively, at pH 7.4 and 37 degrees C. These reaction rates are significantly greater than those for GSH reaction with hydrogen peroxide and non-nitrated electrophilic fatty acids including 8-iso-prostaglandin A2 and 15-deoxy-Delta(12,14)-prostaglandin J2. Increasing reaction pH from 7.4 to 8.9 enhanced apparent second-order rate constants for the thiol reaction with OA-NO2 and LNO2, showing dependence on the thiolate anion of GSH for reactivity. Rates of nitroalkene reaction with thiols decreased as the pKa of target thiols increased. Increasing concentrations of the detergent octyl-beta-d-glucopyranoside decreased rates of nitroalkene reaction with GSH, indicating that the organization of nitro-fatty acids into micellar or membrane structures can limit Michael reactivity with more polar nucleophilic targets. In aggregate, these results reveal that the reversible adduction of thiols by nitro-fatty acids is a mechanism for reversible post-translational regulation of protein function by nitro-fatty acids.  相似文献   

19.
The nucleophilic substitution reaction between glutathione and 1-chloro-2,4-dinitrobenzene has been studied at temperatures between 4 and 42°C and pH values between 6.99 and 10.80. The apparent enthalpy, entropy and free energy of ionization of the thiol group have been estimated as have the apparent enthalpy, entropy and free energy of activation of the reaction between the glutathione thiolate anion and the aromatic electrophile. The results obtained permit the calculation of values of the second order rate constant governing the reaction at a range of temperatures and pHs. These values are in accord with those reported in the literature from experimental work by others. The major glutathione S-transferase from Galleria mellonella has been studied with respect to its kinetic responses to changes in pH and temperature. There appear to be two kinetically critical ionizations governing the reaction at high pH. These ionization events are characterized by apparent pKa values of 8.61 ± 0.15 and 9.16 ± 0.22. A thermodynamic model of the kinetic behavior of the enzyme permits the prediction of its activity over a range of pH and temperature values. The apparent free energy of activation for the enzyme catalyzed reaction is only 7% lower than that for the non-catalyzed reaction between 1-chloro-2,4-dinitrobenzene and glutathione thiolate anion. This observation is compatible with the suggestion that promotion of the ionization of the glutathione thiol group is the major mechanism of catalysis.  相似文献   

20.
The reaction between 2,2'-dipyridyl disulphide and the thiol group in bovine serum albumin has been studied at pH 1.1-7.9. At pH 5.5-7.9 the reaction rate was second order in dipyridyl disulphide and thiolate ion, as expected for an aliphatic thiol compound. Below pH 5.5 the reaction rate increased and became maximum at pH 2.6. The observed rate constant (110 M-1-s-1) was comparable with that at pH 6.6, although the thiolate ion concentration should be 10(4) times less at the lower pH. The increase in reactivity seemed to be correlated with the conformational change in serum albumin at pH 3.6-4.0. Increased nucleophilicity due to interaction with some suitable functional group might explain the high reactivity of the SH group at acidic pH.  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