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 78 毫秒
1.
A beta-glycosidase (M(r) 50000) from Spodoptera frugiperda larval midgut was purified, cloned and sequenced. It is active on aryl and alkyl beta-glucosides and cellodextrins that are all hydrolyzed at the same active site, as inferred from experiments of competition between substrates. Enzyme activity is dependent on two ionizable groups (pK(a1)=4.9 and pK(a2)=7.5). Effect of pH on carbodiimide inactivation indicates that the pK(a) 7.5 group is a carboxyl. k(cat) and K(m) values were obtained for different p-nitrophenyl beta-glycosides and K(i) values were determined for a range of alkyl beta-glucosides and cellodextrins, revealing that the aglycone site has three subsites. Binding data, sequence alignments and literature beta-glycosidase 3D data supported the following conclusions: (1) the groups involved in catalysis were E(187) (proton donor) and E(399) (nucleophile); (2) the glycone moiety is stabilized in the transition state by a hydrophobic region around the C-6 hydroxyl and by hydrogen bonds with the other equatorial hydroxyls; (3) the aglycone site is a cleft made up of hydrophobic amino acids with a polar amino acid only at its first (+1) subsite.  相似文献   

2.
Karsten WE  Liu D  Rao GS  Harris BG  Cook PF 《Biochemistry》2005,44(9):3626-3635
The pH dependence of kinetic parameters of several active site mutants of the Ascaris suum NAD-malic enzyme was investigated to determine the role of amino acid residues likely involved in catalysis on the basis of three-dimensional structures of malic enzyme. Lysine 199 is positioned to act as the general base that accepts a proton from the 2-hydroxyl of malate during the hydride transfer step. The pH dependence of V/K(malate) for the K199R mutant enzyme reveals a pK of 5.3 for an enzymatic group required to be unprotonated for activity and a second pK of 6.3 that leads to a 10-fold loss in activity above the pK of 6.3 to a new constant value up to pH 10. The V profile for K199R is pH independent from pH 5.5 to pH 10 and decreases below a pK of 4.9. Tyrosine 126 is positioned to act as the general acid that donates a proton to the enolpyruvate intermediate to form pyruvate. The pH dependence of V/K(malate) for the Y126F mutant is qualitatively similar to K199R, with a requirement for a group to be unprotonated for activity with a pK of 5.6 and a partial activity loss of about 3-fold above a pK of 6.7 to a new constant value. The Y126F mutant enzyme is about 60000-fold less active than the wild-type enzyme. In contrast to K199R, the V rate profile for Y126F also shows a partial activity loss above pH 6.6. The wild-type pH profiles were reinvestigated in light of the discovery of the partial activity change for the mutant enzymes. The wild-type V/K(malate) pH-rate profile exhibits the requirement for a group to be unprotonated for catalysis with a pK of 5.6 and also shows the partial activity loss above a pK of 6.4. The wild-type V pH-rate profile decreases below a pK of 5.2 and is pH independent from pH 5.5 to pH 10. Aspartate 294 is within hydrogen-bonding distance to K199 in the open and closed forms of malic enzyme. D294A is about 13000-fold less active than the wild-type enzyme, and the pH-rate profile for V/K(malate) indicates the mutant is only active above pH 9. The data suggest that the pK present at about pH 5.6 in all of the pH profiles represents D294, and during catalysis D294 accepts a proton from K199 to allow K199 to act as a general base in the reaction. The pK for the general acid in the reaction is not observed, consistent with rapid tautomerization of enolpyruvate. No other ionizable group in the active site is likely responsible for the partial activity change observed in the pH profiles, and thus the group responsible is probably remote from the active site and the effect on activity is transmitted through the protein by a conformational change.  相似文献   

3.
The phospholipase A2 from the venom of A. halys blomhoffii was titrated with micellar n-hexadecylphosphorylcholine (an analog of lysolecithin) by following the tryptophyl fluorescence change at 25 degrees C and ionic strength 0.1. The data were analyzed by assuming that the micellar surface has multiple binding sites for the enzyme and that these sites are identical and mutually independent. The enzyme binding site was found to accommodate a constant number of the substrate (monomer) molecules, N = 10.0 and 6.7 for the apoenzyme and its Ca2+ complex, respectively. The binding constant of the enzyme to the substrate micelle was found to be enhanced by Ca2+ binding to the enzyme. The pH dependence of the binding constant of the apoenzyme to the micelle was well interpreted in terms of pK shifts of two ionizable groups from 5.16 to 5.67 and from 6.45 to 6.6. The pH-dependence curve for the enzyme-Ca2+ complex, which lacked the former transition, was interpreted in terms of the pK shift of a single ionizable group from 5.55 to 5.76. The former ionizable group was assigned as Asp 49, to which Ca2+ ion can coordinate, and the latter as His 48 in the active site. No participation of the alpha-amino group with a pK value of 7.30 was observed. The binding constant of the enzyme to the substrate micelle, Kmic = 0.45-2.3 X 10(6) M-1, was found to be far greater than that to the monomeric substrate, Kmon = 0.2-1.0 X 10(4) M-1. This was interpreted in terms of the presence of an additional weak substrate-binding site in the enzyme molecule.  相似文献   

4.
Biosorption of copper ions by an industrial algal waste, from agar extraction industry has been studied in a batch system. This biosorbent was compared with the algae Gelidium itself, which is the raw material for agar extraction, and the industrial waste immobilized with polyacrylonitrile (composite material). The effects of contact time, pH, ionic strength (IS) and temperature on the biosorption process have been studied. Equilibrium data follow both Langmuir and Langmuir-Freundlich models. The parameters of Langmuir equilibrium model were: q(max)=33.0mgg(-1), K(L)=0.015mgl(-1); q(max)=16.7mgg(-1), K(L)=0.028mgl(-1) and q(max)=10.3mgg(-1), K(L)=0.160mgl(-1) respectively for Gelidium, algal waste and composite material at pH=5.3, T=20 degrees C and IS=0.001M. Increasing the pH, the number of deprotonated active sites increases and so the uptake capacity of copper ions. In the case of high ionic strengths, the contribution of the electrostatic component to the overall binding decreases, and so the uptake capacity. The temperature has little influence on the uptake capacity principally for low equilibrium copper concentrations. Changes in standard enthalpy, Gibbs energy and entropy during biosorption were determined. Kinetic data at different solution pH (3, 4 and 5.3) were fitted to pseudo-first-order and pseudo-second-order models. The adsorptive behaviour of biosorbent particles was modelled using a batch reactor mass transfer kinetic model, which successfully predicts Cu(II) concentration profiles.  相似文献   

5.
The chromophore of octopus rhodopsin is 11-cis retinal, linked via a protonated Schiff base to the protein backbone. Its stable photoproduct, metarhodopsin, has all-trans retinal as its chromphore. The Schiff base of acid metarhodopsin (lambda max = 510 nm) is protonated, whereas that of alkaline metarhodopsin (lambda max = 376 nm) is unprotonated. Metarhodopsin in photoreceptor membranes was titrated and the apparent pK of the Schiff base was measured at different ionic strengths. From these salt-dependent pKs the surface charge density of the octopus photoreceptor membranes and the intrinsic Schiff base pK of metarhodopsin were obtained. The surface charge density is sigma = -1.6 +/- 0.1 electronic charges per 1,000 A2. Comparison of the measured surface charge density with values from octopus rhodopsin model structures suggests that the measured value is for the extracellular surface and so the Schiff base in metarhodopsin is freely accessible to protons from the extracellular side of the membrane. The intrinsic Schiff base pK of metarhodopsin is 8.44 +/- 0.12, whereas that of rhodopsin is found to be 10.65 +/- 0.10 in 4.0 M KCl. These pK values are significantly higher than the pK value around 7.0 for a retinal Schiff base in a polar solvent; we suggest that a plausible mechanism to increase the pK of the retinal pigments is the preorganization of their chromophore-binding sites. The preorganized site stabilizes the protonated Schiff base with respect to the unprotonated one. The difference in the pK for the octopus rhodopsin compared with metarhodopsin is attributed to the relative freedom of the latter's chromophore-binding site to rearrange itself after deprotonation of the Schiff base.  相似文献   

6.
1. Kinetic studies of lipoamide dehydrogenase and its modified enzymes catalyzing lipoamide oxidoreduction and ancillary reactions at various pH are compared. 2. The asymptotic kinetics of lipoamide oxidoreductions switch between the ping pong and ordered mechanisms by varying pH of the reactions. 3. pH-rate profiles of these reactions are bell-shaped suggesting the participation of 2 ionizable residues with pK values of 6.6 +/- 0.5 and above 8 respectively. 4. The unusually high pK value for the catalytic site histidine is attributed to its involvement in an ion-pair formation. 5. In the absence of the catalytic site histidine, the pH-rate profile for the lipoamide reduction of the photooxidized enzyme is no longer bell-shaped but it is similar to those of the transhydrogenation and NADH-oxidation of the native enzyme. 6. This implies the participation of a low-pK protonated group in these reactions.  相似文献   

7.
Bacterial L-asparaginases (E.C. 3.5.1.1) have been used as therapeutic agents in the treatment of acute childhood lymphoblastic leukaemia. L-asparaginase from Erwinia carotovora NCYC 1526 (ErA) was cloned and expressed in E. coli. The enzyme was purified to homogeneity by a two-step procedure comprising cation-exchange chromatography and affinity chromatography on immobilised L-asparagine. The enzymatic properties of the recombinant enzyme were investigated and the kinetic parameters (K(m), k(cat)) for a number of substrates were determined. Molecular modelling studies were also employed to create a model of ErA, based on the known structure of the Erwinia chrysanthemi enzyme. The molecular model was used to help interpret biochemical data concerning substrate specificity and catalytic mechanism of the enzyme. The kinetic parameters of selected substrates were determined at various pH values, and the pH-dependence profiles of V(max) and V(max)/K(m) were analyzed. The pH-dependence of V(max) shows one transition in the acidic pH range with pK(a)=5.4, and the pH-dependence of V(max)/K(m) exhibits two transitions with pK(a)=5.4 and 8.5. Based on analysis of alternative substrates and molecular modelling studies, it was concluded that the pK(a) at the acidic pH range corresponds to the active site residues Asp115 or Glu82, whereas the pK(a) observed at the alkaline pH range is not due to substrate amino group ionisation, but rather is the result of enzyme ionisation. The effect of temperature and viscosity on the catalytic activity of the enzyme was also investigated and it was concluded that the rate-limiting step of the catalytic reaction is relevant to structural transitions of the protein. Thermodynamic analysis of the activity data showed that the activation energies are dependent on the substrate, and entropy changes appear to be the main determinant contributing to substrate specificity.  相似文献   

8.
Co-solvent and temperature effects on the pK of histidine (imidazolium) residue 46 of trypsin, as well as of weak electrolytes (buffers), which have been reported in two preceding papers, can be satisfactorily explained in terms of enthalpy-entropy compensation patterns. Such patterns have been generated for various mixed solvents between 20 degrees and minus 20 degrees and minus 50 degrees. Under these conditions compensation temperature, T-c, is strongly dependent on the nature of the ionizable group studied: 240 plus or minus 10 K for neutral acids and 310 plus or minus 5 K for cationic acids. This work focuses on the possibilities offered and on the problems raised by the use of this methodology as a tool in the investigation of the active site of enzymes. Furthermore, it is shown in the case of histidine residue 46 of trypsin that the co-solvent effect vanishes at the compensation temperature, a result of great practical significance if applicable to any ionizable group at the active site of enzymes.  相似文献   

9.
1. The effects of varying pH, ionic strength and temperature on the parameters K(m) and V(max.) for a purified alkaline phosphatase from calf intestinal mucosa with a new fluorogenic substrate, 4-methylumbelliferyl phosphate monoester disodium salt, and an ammediol-hydrochloric acid buffer system were determined. 2. It was found that, under varying conditions, a relationship exists between K(m) and V(max.) such that V(max.)=beta/(1+alpha/K(m)), where alpha and beta are constants, temperature- and ionic strength-dependent, but pH-independent. It is shown that this relationship accounts satisfactorily for the well-known effect of varying substrate concentration on optimum pH and velocity. 3. The various results are interpreted in terms of a pH-dependent conformational equilibrium between two forms of the enzyme, E(1) and E(2). Only E(1) combines with substrate, and only E(2) reacts to give inorganic phosphate. 4. To account for the pH-variation of K(m) and V(max.) in terms of this theory, it is postulated that the conformational change is associated with a change in pK of two basic groups in the enzyme.  相似文献   

10.
The optimisation of enzymes for particular application or conditions remains an important target in all protein engineering endeavours. Here, we report a successful strategy for altering the pH-profile of kinetic parameters and to define in detail the molecular mechanism of maize glutathione S-transferase I (GST I). To accomplish this, selected residues from the glutathione binding site (His40, Ser11, Lys41, Asn49, Gln53 and Ser67) were mutated to Ala, and the pH-dependence of the catalytic parameters V(max), and V(max)/K(GSH)(m) of the mutated forms were analysed. The pH-dependence of V(max) for the wild-type enzyme exhibits two transitions in the acidic pH range with pK(a1) of 5.7 and pK(a2) of 6.6. Based on thermodynamic data, site-directed mutagenesis and UV deference spectroscopy, it was concluded that pK(a1) corresponds to GSH carboxylates, whereas the pK(a2) has a conformational origin of the protein. The pH-dependence of V(max)/K(GSH)(m) for the wild-type enzyme exhibits a single transition with pK(a) of 6.28 which was attributed to the thiol ionisation of bound GSH. These findings complement the conclusions about the catalytic mechanism deduced from the crystal structure of the enzyme and provide the basis for rationally designing engineered forms of GST I with valuable properties.  相似文献   

11.
Gadda G  Fitzpatrick PF 《Biochemistry》2000,39(6):1406-1410
Nitroalkane oxidase catalyzes the oxidation of nitroalkanes to aldehydes or ketones with production of nitrite and hydrogen peroxide. pH and kinetic isotope effects with [1, 1-(2)H(2)]nitroethane have been used to study the mechanism of this enzyme. The V/K(ne) pH profile is bell-shaped. A group with a pK(a) value of about 7 must be unprotonated and one with a pK(a) value of 9.5 must be protonated for catalysis. The lower pK(a) value is seen also in the pK(is) profile for the competitive inhibitor valerate, indicating that nitroethane has no significant external commitments to catalysis. The (D)(V/K)(ne) value is pH-independent with a value of 7.5, whereas the (D)V(max) value increases from 1.4 at pH 8.2 to a limiting value of 7.4 below pH 5. The V(max) pH profile decreases at low and high pH, with pK(a) values of 6.6 and 9.5, respectively. Imidazole, which activates the enzyme, affects the V(max) but not the V/K(ne) pH profile. In the presence of imidazole at pH 7 the (D)V(max) value increases to a value close to the intrinsic value, consistent with cleavage of the carbon-hydrogen bond of the substrate being fully rate-limiting for catalysis in the presence of imidazole.  相似文献   

12.
The effect of pH on the kinetic parameters (Km and Ki) for extracellular acid Penicillium brevicompactum RNAse (pH max 4.7+/-0.1), non-specific to the chemical nature of nucleic bases, was studied. The pKm--pH dependence curve showed bends within the following intervals of pH: 3.5--4.0 and 5.6--6.0 (upward side concavity) and 6.2--6.8 (downward side concavity). The pKi--pH dependence for adenosine-3'-monophosphate as an inhibitor is identical to the pH dependence on pKm for the substrate. On the other hand, the pKi--pH dependence curves obtained for the base-free inhibitors (ribose-5'-monophosphate, or phosphate (adenosine) show no bends within the pH intervals of 3.0--4.0 and 5.6--7.0 respectively. A possibility is discussed of the presence of a carboxylic (pK 3.58+/-0.1) and two imidazole groups (pK 6.42+/-0.1--a weakly protonated and 5.8+/-+/-0.1--a strongly protonated group) in the RNAse active site and their participation in the formation of the RNAse-nucleotide (RNAse-substrate) complex.  相似文献   

13.
The flavoprotein oxidase Fms1 from Saccharomyces cerevisiae catalyzes the oxidation of spermine and N(1)-acetylspermine to spermidine and 3-aminopropanal or N-acetyl-3-aminopropanal. Within the active site of Fms1, His67 is positioned to form hydrogen bonds with the polyamine substrate. This residue is also conserved in other polyamine oxidases. The catalytic properties of H67Q, H67N, and H67A Fms1 have been characterized to evaluate the role of this residue in catalysis. With both spermine and N(1)-acetylspermine as the amine substrate, the value of the first-order rate constant for flavin reduction decreases 2-3 orders of magnitude, with the H67Q mutation having the smallest effect and H67N the largest. The k(cat)/K(O2) value changes very little upon mutation with N(1)-acetylspermine as the amine substrate and decreases only an order of magnitude with spermine. The k(cat)/K(M)-pH profiles with N(1)-acetylspermine are bell-shaped for all the mutants; the similarity to the profile of the wild-type enzyme rules out His67 as being responsible for either of the pK(a) values. The pH profiles for the rate constant for flavin reduction for all the mutant enzymes similarly show the same pK(a) as wild-type Fms1, about ~7.4; this pK(a) is assigned to the substrate N4. The k(cat)/K(O2)-pH profiles for wild-type Fms1 and the H67A enzyme both show a pK(a) of about ~6.9; this suggests His67 is not responsible for this pH behavior. With the H67Q, H67N, and H67A enzymes the k(cat) value decreases when a single residue is protonated, as is the case with the wild-type enzyme. The structure of H67Q Fms1 has been determined at a resolution of 2.4 ?. The structure shows that the mutation disrupts a hydrogen bond network in the active site, suggesting that His67 is important both for direct interactions with the substrate and to maintain the overall active site structure.  相似文献   

14.
F Guerlesquin  J C Sari  M Bruschi 《Biochemistry》1987,26(23):7438-7443
The complex formation between cytochrome c3 and ferredoxin I from Desulfovibrio desulfuricans Norway was studied by microcalorimetric and pH-stat titration measurements. The stoichiometry of the complex was found to be one molecule of cytochrome c3 per monomer of ferredoxin I. The association constant determined at T = 283 K in tris(hydroxymethyl)aminomethane hydrochloride (Tris-HCl) buffer, 10(-2) M and pH 7.7, was KA = 1.3 X 10(6) M-1. Though the enthalpy (delta H = 19 +/- 1 kJ.mol-1) and the entropy (delta S = 183 J.K-1.mol-1) were positive and consistent with a hydrophobic process involved in the interaction, the analysis of ionic strength dependence exhibited an important electrostatic effect on the association. The use of both Tris-HCl and phosphate buffers during microcalorimetric experiments showed proton release at pH 6.6. The pH-stat study of proton release indicated that one of the charged groups involved in the interacting site underwent a pK shift from 7.35 to 6.05.  相似文献   

15.
A speciation study was carried out in aqueous solution of the anti-inflammatory drug tenoxicam (Htenox), under quasi-physiological conditions (temperature of 37 degrees C and ionic strength 0.15 M NaCl) in order to determine the acidity constants from spectrophotometric studies, the pK(a) values found being pK(1)=1.143+/-0.008 and pK(2)=4.970+/-0.004. Subsequently, the spectrophotometrical speciation of the different complexes of Cu(II) with the drug was performed under the same conditions of temperature and ionic strength, observing the formation of Cu(Htenox)(2)(2+) with log beta(212)=20.05+/-0.01, Cu(tenox)(2) with log beta(012)=13.6+/-0.1, Cu(Htenox)(2+) with log beta(111)=10.52+/-0.08, as well as Cu(tenox)(+) with log beta(011)=7.0+/-0.2, all of them in solution, and solid species Cu(tenox)(2)(s) with an estimated value of log beta(012)(s) approximately 18.7. The crystalline structure of the complex [Cu(tenox)(2)(py)(2)]. EtOH, was also determined, and it was observed that tenoxicam employs the oxygen of the amide group and the pyridyl nitrogen to bond to the cation.  相似文献   

16.
Paracoccus pantotrophus grown anaerobically under denitrifying conditions expressed similar levels of the periplasmic nitrate reductase (NAP) when cultured in molybdate- or tungstate-containing media. A native PAGE gel stained for nitrate reductase activity revealed that only NapA from molybdate-grown cells displayed readily detectable nitrate reductase activity. Further kinetic analysis showed that the periplasmic fraction from cells grown on molybdate (3 microM) reduced nitrate at a rate of V(max)=3.41+/-0.16 micromol [NO(3)(-)] min(-1) mg(-1) with an affinity for nitrate of K(m)=0.24+/-0.05 mM and was heat-stable up to 50 degrees C. In contrast, the periplasmic fraction obtained from cells cultured in media supplemented with tungstate (100 microM) reduced nitrate at a much slower rate, with much lower affinity (V(max)=0.05+/-0.002 micromol [NO(3)(-)] min(-1) mg(-1) and K(m)=3.91+/-0.45 mM) and was labile during prolonged incubation at >20 degrees C. Nitrate-dependent growth of Escherichia coli strains expressing only nitrate reductase A was inhibited by sub-mM concentrations of tungstate in the medium. In contrast, a strain expressing only NAP was only partially inhibited by 10 mM tungstate. However, none of the above experimental approaches revealed evidence that tungsten could replace molybdenum at the active site of E. coli NapA. The combined data show that tungsten can function at the active site of some, but not all, molybdoenzymes from mesophilic bacteria.  相似文献   

17.
The hydrolysis of 1,2-dihexanoyl-sn-glycero-3-phosphorylcholine (diC6PC), catalyzed by the phospholipase A2 from the venom of Agkistrodon halys blomhoffii, was studied at 25 degrees C and the ionic strength of 0.1 in the presence of 3-33.3 mM Ca2+, which can saturate the Ca2+-binding site of the enzyme. The initial velocity data, obtained at various concentrations of the substrate below the critical micelle concentration (cmc), were analyzed according to the Michaelis-Menten equation. The pH-dependence curve of the Km value exhibited only one transition below pH 8. The analytical results indicated that the pK value of 6.30 of an ionizable group changed to 6.54 on the binding of the monodispersed substrate. This ionizable group was assigned as the alpha-amino group on the basis of its pK value, which had been determined from the pH dependence of the binding constant of monodispersed n-dodecylphosphorylcholine (n-C12PC) (Ikeda and Samejima (1981) J. Biochem. 90, 799-804, and Haruki et al. (1986) J. Biochem. 99, 99-109). The pH-dependence curve of the kcat value exhibited two transitions, below pH 6.5 and above pH 9.5. The analytical results indicated the participation of two ionizable groups with pK values of 5.55 and 10.50. Deprotonation of the former and protonation of the latter group were found to be essential for the catalysis. The former ionizable group was assigned as His 48 in the active site on the basis of its pK value, which had been determined from the pH dependence of the binding constant of Ca2+ (Ikeda et al. (1981) J. Biochem. 90, 1125-1130).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Mixed-solvent systems of methanol and other alcohols and water were used to study the properties of bovine phenylethanolamine-N-methyltransferase. The presence of methanol decreased the binding affinity of the enzyme for its amine substrate but did not alter the maximum velocity. The change in binding was accompanied by an alkaline shift in the pK of an ionizable group in the active site. The well-known property of enzyme inhibition by substrate was also alleviated. Increasing the pH of the medium, in the presence or absence of methanol, increased the maximum velocity but did not alter substrate inhibition. It is proposed that substrate inhibition is due in part to the ionic state of a single unidentified ionizable group in the active site of the enzyme and to a slow release of product. Evidence that an essential, pH-dependent sulfhydryl modulates product release is presented. The properties of phenylethanolamine-N-methyl-transferase are quite responsive to changes in pH, ionic strength, and water content so that the enzyme may well be regulated at the microenvironmental level.  相似文献   

19.
1. To identify the functional groups that are involved in the conversion of beta-glycerophosphate by alkaline phosphatase (EC 3.1.3.1) from pig kidney, the kinetics of alkaline phosphatase were investigated in the pH range 6.6-10.3 at substrate concentrations of 3 muM-30 mM. From the plots of log VH+ against pH and log VH+/KH+m against pH one functional group with pK = 7.0 and two functional groups with pK = 9.1 were identified. These groups are involved in substrate binding. Another group with pK = 8.8 was found, which in its unprotonated form catalyses substrate conversion. 2. GSH inhibits the alkaline phosphatase reversibly and non-competitively by attacking the bound Zn(II). 3. The influence of the H+ concentration on the activation by Mg2+ ions of alkaline pig kidney phosphate was investigated between pH 8.4 and 10.0. The binding of substrate and activating Mg2+ ions occurs independently at all pH values between 8.4 and 10.0. The activation mechanism is not affected by the H+ concentration. The Mg2+ ions are bound by a functional group with a pK of 10.15. 4. A scheme is proposed for the reaction between enzyme, substrate, Mg2+ and H+ and the overall rate equation is derived. 5. The mechanism of substrate binding and splitting by the functional groups of the active centre is discussed on the basis of a model. Mg2+ seems to play a role as an autosteric effector.  相似文献   

20.
The enzyme glycinamide ribonucleotide transformylase (GART) catalyzes the transfer of a formyl group from formyl tetrahydrofolate (fTHF) to glycinamide ribonucleotide (GAR), a process that is pH-dependent with pK(a) of approximately 8. Experimental studies of pH-rate profiles of wild-type and site-directed mutants of GART have led to the proposal that His108, Asp144, and GAR are involved in catalysis, with His108 being an acid catalyst, while forming a salt bridge with Asp144, and GAR being a nucleophile to attack the formyl group of fTHF. This model implied a protonated histidine with pK(a) of 9.7 and a neutral GAR with pK(a) of 6.8. These proposed unusual pK(a)s have led us to investigate the electrostatic environment of the active site of GART. We have used Poisson-Boltzmann-based electrostatic methods to calculate the pK(a)s of all ionizable groups, using the crystallographic structure of a ternary complex of GART involving the pseudosubstrate 5-deaza-5,6,7,8-THF (5dTHF) and substrate GAR. Theoretical mutation and deletion analogs have been constructed to elucidate pairwise electrostatic interactions between key ionizable sites within the catalytic site. Also, a construct of a more realistic catalytic site including a reconstructed pseudocofactor with an attached formyl group, in an environment with optimal local van der Waals interactions (locally minimized) that imitates closely the catalytic reactants, has been used for pK(a) calculations. Strong electrostatic coupling among catalytic residues His108, Asp144, and substrate GAR was observed, which is extremely sensitive to the initial protonation and imidazole ring flip state of His108 and small structural changes. We show that a proton can be exchanged between GAR and His108, depending on their relative geometry and their distance to Asp144, and when the proton is attached on His108, catalysis could be possible. Using the formylated locally minimized construct of GART, a high pK(a) for His108 was calculated, indicating a protonated histidine, and a low pK(a) for GAR(NH(2)) was calculated, indicating that GAR is in neutral form. Our results are in qualitative agreement with the current mechanistic picture of the catalytic process of GART deduced from the experimental data, but they do not reproduce the absolute magnitude of the pK(a)s extracted from fits of k(cat)-pH profiles, possibly because the static time-averaged crystallographic structure does not describe adequately the dynamic nature of the catalytic site during binding and catalysis. In addition, a strong effect on the pK(a) of GAR(NH(2)) is produced by the theoretical mutations of His108Ala and Asp144Ala, which is not in agreement with the observed insensitivity of the pK(a) of GAR(NH(2)) modeled from the experimental data using similar mutations. Finally, we show that important three-way electrostatic interactions between highly conserved His137, with His108 and Asp144, are responsible for stabilizing the electrostatic microenvironment of the catalytic site. In conclusion, our data suggest that further detailed computational and experimental work is necessary.  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