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 703 毫秒
1.
A new procedure that employs a one-pot, oxidative Hofmann rearrangement to incorporate a urea linkage into peptide backbones is detailed herein. This methodology was used to replace the scissile peptide bonds of [Leu5]enkephalin and a hexapeptide HIV-1 protease substrate. The [Leu5]enkephalin analogue was found to inhibit cleavage of hippurylhistidylleucine (HHL) by porcine kidney angiotensin-converting enzyme (PK-ACE) with a 0.88 mM IC50 value, comparable to the Michaelis constant of [Leu5]enkephalin with the same enzyme. The HIV-1 protease substrate analogue was shown to inhibit HIV-1 protease with an IC50=34 microM.  相似文献   

2.
Single-crystal X-ray studies of a manganese(II) derivative of guanosine 5'-monophosphate, [Mn(5'-GMP)(H(2)O)(5)],3H(2)O, have shown that it is isostructural with its nickel analogue. The manganese atom therefore is bonded to five water molecules with the remaining octahedral co-ordination site being occupied by N-7 of the nucleotide base. No direct metal-phosphate bonding is involved, but there are structure-stabilizing intramolecular hydrogen bonds between two phosphate oxygen atoms and co-ordinated water molecules.  相似文献   

3.
The compound P1-(5'-adenosyl)-P2-N-(2-mercaptoethyl)diphosphoramidate (AMEDA) was synthesized as an ATP analogue for in situ reaction with the 4-nitro-2,1,3-[14C]benzoxadiazolyl group (NBD) in the labeled F1-ATPase (F1). AMEDA was found to reactivate O-[14C]NBD-F1 via a dual-path mechanism. The principal path involves the binding of AMEDA at a site in F1 with Kd = 14.5 microM and subsequent reaction with the [14C]NBD label. The second slower path involves the direct biomolecular reaction of AMEDA with the radioactive label on F1. The rate of reactivation of O-[14C]NBD-F1 by AMEDA was decreased by ADP or ATP which competes with the ATP analogue for binding to the labeled enzyme. The reaction product was found to contain one adenine group, two phosphate groups, and one [14C]NBD label per molecule as expected from the structure of the compound AMEDA-[14C]NBD. Purified AMEDA-[14C]NBD was found to bind to unlabeled F1 with Kd = 2 microM. These observations demonstrate the in situ reaction of bound AMEDA with the nearby [14C]NBD label attached to Tyr-beta 311 and support the assumed presence of Tyr-beta 311 near the phosphate groups of ATP bound at the hydrolytic site of F1-ATPase. The possible locations of Tyr-beta 364, His-beta 427, and Tyr-beta 345 relative to Tyr-beta 311 in F1 are discussed, and the validity of the previously proposed model for F1-ATPase with one hydrolytic site assisted by two auxiliary sites is examined and compared with that of the widely accepted alternating sites model.  相似文献   

4.
The influence of substrate inhibition on xanthine oxidase-intramolecular electron transport was studied by steady-state kinetic analysis. Experiments with hypoxanthine and xanthine up to 900 microM indicated an inhibition pattern which fitted an equation of the general form nu 0 = nu max . [S]/(Km + a[S] + b[S]2/Ki). Univalent electron flux to oxygen was favored at substrate concentrations above 50 microM. This augmentation of univalent flux percentage that appeared at a high substrate concentration was greater for hypoxanthine that xanthine and at pH 8.3 than at 9.5. Our results support a mechanism of inhibition in which a substrate-reduced enzyme, non-productive Michaelis complex was formed. It is possible that this non-productive complex favored the univalent pathway of enzyme reoxidation (superoxide production) by increasing the midpoint redox potential of the molybdenum active site.  相似文献   

5.
Binding of chromomycin A3 (CRA) to calf thymus DNA was investigated in the presence of divalent cations using visible absorption and 1H-nmr spectroscopies. An apparent equilibrium binding constant (approximately 10(11) M-1) was obtained from metal competition experiments using EDTA to remove the metal cation from the DNA-M-CRA (M: metal) complex. The large binding constant of the drug to DNA enabled us to obtain essentially complete complexation of CRA to the short homogeneous d(ATGCAT)2 duplex using stoichiometric amounts of the metal cation. Large induced chemical shifts were observed in the 1H-nmr spectrum of the above complex using the paramagnetic Co2+ cation, indicating that the metal occupies a unique binding site. Since no induced 1H-nmr chemical shifts were observed for the drug-Co2+ mixture, it was concluded that no metal-drug complex is formed. In addition, it was found that bound CRA is negatively charged at physiological pH and binding to the DNA could be affected only by using metal cations whose ionic radius size (less than 0.85 A) and charge (2+) were simultaneously satisfied. Stringent metal cation selectivity for the DNA-M-CRA complex may be intimately connected with the antitumor selectivity of CRA, since different types of cells generally possess widely differing molar concentrations of metal cations.  相似文献   

6.
Aree T  Chaichit N 《Carbohydrate research》2003,338(15):1581-1589
A new crystal form of beta-cyclodextrin (beta-CD)[bond]ethanol[bond]dodecahydrate inclusion complex [(C(6)H(10)O(5))(7).0.3C(2)H(5)OH.12H(2)O] belongs to monoclinic space group C2 (form II) with unit cell constants a=19.292(1), b=24.691(1), c=15.884(1) A, beta=109.35(1) degrees. The beta-CD macrocycle is more circular than that of the complex in space group P2(1) [form I: J. Am. Chem. Soc. 113 (1991) 5676]. In form II, a disordered ethanol molecule (occupancy 0.3) is placed in the upper part of beta-CD cavity (above the O-4 plane) and is sustained by hydrogen bonding to water site W-2. In form I, an ethanol molecule located below the O-4-plane is well ordered because it hydrogen bonds to surrounding O-3[bond]H, O-6[bond]H groups of the symmetry-related beta-CD molecules. In the crystal lattice of form I, beta-CD macrocycles are stacked in a typical herringbone cage structure. By contrast, the packing structure of form II is a head-to-head channel that is stabilized at both O-2/O-3 and O-6 sides of each beta-CD by direct O(CD)...O(CD) and indirect O(CD)...O(W)...(O(W))...O(CD) hydrogen bonds. The 12 water molecules are disordered in 18 positions both inside the channel-like cavity of beta-CD dimer (W-1[bond]W-6) and in the interstices between the beta-CD macrocycles (W-7[bond]W-18). The latter forms a cluster that is hydrogen bonded together and to the neighboring beta-CD O[bond]H groups.  相似文献   

7.
Acetohydroxy acid isomeroreductase (EC 1.1.1.86), the second enzyme of the parallel branched chain amino acid pathway, is a homodimer with an Mr of approximately 114000 which in the presence of Mg2+ ions catalyzes an unusual alkyl migration followed by an NADPH-dependent reduction. Prior binding of NADPH and Mg2+ to the enzyme was shown to be required for substrate or competitive inhibitor [N-hydroxy-N-isopropyloxamate (IpOHA)] binding [Dumas, R., et al. (1994) Biochem. J. 301, 813-820]. Moreover, crystallographic data for the enzyme-NADPH-Mg2+-IpOHA complex [Biou, V., et al. (1997) EMBO J. 16, 3405-3415] have shown that IpOHA was completely buried inside the active site. These observations raised the question of how the reaction intermediate analogue inhibitor can reach the active site and implied that conformational changes occurred during the binding process. With a view of characterizing these conformational changes, H-D exchange experiments combined with mass spectrometry were performed. Results demonstrated that Mg2+ ions and NADPH binding led to an initial conformational change at the interface of the two domains of each monomer. Binding of the two cofactors to isomeroreductase alters the structure of the active site to promote inhibitor (substrate) binding, in agreement with the ordered mechanism of the enzyme. Structural changes remote from the active site were also found. They were interpreted as long-range structural effects on the two domains and on the two monomers in the time course of the ligand binding process.  相似文献   

8.
The alpha(2)beta(2) tetrameric E1 component of the branched-chain 2-oxo acid (BCOA) dehydrogenase multienzyme complex is a thiamin diphosphate (ThDP)-dependent enzyme. E1 catalyzes the decarboxylation of a BCOA concomitant with the formation of the alpha-carbanion/enamine intermediate, 2-(1-hydroxyalkyl)-ThDP, followed by transfer of the 1-hydroxyalkyl group to the distal sulfur atom on the lipoamide of the E2 component. In order to elucidate the catalytic mechanism of E1, the alpha- and beta-subunits of E1 from Thermus thermophilus HB8 have been co-expressed in Escherichia coli, purified and crystallized as a stable complex, and the following crystal structures have been analyzed: the apoenzyme (E1(apo)), the holoenzyme (E1(holo)), E1(holo) in complex with the substrate analogue 4-methylpentanoate (MPA) as an ES complex model, and E1(holo) in complex with 4-methyl-2-oxopentanoate (MOPA) as the alpha-carbanion/enamine intermediate (E1(ceim)). Binding of cofactors to E1(apo) induces a disorder-order transition in two loops adjacent to the active site. Furthermore, upon binding of MPA to E1(holo), the loop comprised of Gly121beta-Gln131beta moves close to the active site and interacts with MPA. The carboxylate group of MPA is recognized mainly by Tyr86beta and N4' of ThDP. The hydrophobic moiety of MPA is recognized by Phe66alpha, Tyr95alpha, Met128alpha and His131alpha. As an intermediate, MOPA is decarboxylated and covalently linked to ThDP, and the conformation of the protein loop is almost the same as in the substrate-free (holoenzyme) form. These results suggest that E1 undergoes an open-closed conformational change upon formation of the ES complex with a BCOA, and the mobile region participates in the recognition of the carboxylate group of the BCOA. ES complex models of E1(holo).MOPA and of E1(ceim).lipoamide built from the above structures suggest that His273alpha and His129beta' are potential proton donors to the carbonyl group of a BCOA and to the proximal sulfur atom on the lipoamide, respectively.  相似文献   

9.
Enolase from rabbit muscle (betabeta-enolase) is inactivated by NaClO(4). Enolase free of divalent cations is more susceptible to inactivation by NaClO(4) than is enolase in the presence of Mg(2+). We find that substrate protects apo-enolase against inactivation, indicating that substrate can bind to enolase in the absence of a divalent cation. This binding is not due to contamination by trace levels of divalent cations since (1) it occurs even in the presence of EDTA or EGTA and (2) metal analysis by ICP (inductively coupled plasma) mass spectrometry did not reveal sufficient contamination to account for the protection. The binding of PGA to apo-enolase did require Na(+). When TMAClO(4) was used instead of NaClO(4), there was no protection by PGA. Protection was restored when TMAClO(4) plus NaCl were used. The inactivation of apo-enolase by NaClO(4) is due to dissociation into inactive monomers. We conclude that Na(+) binds to apo-enolase, permitting substrate to then bind. Of the three known Me(2+) binding sites on enolase, we believe the most likely binding site for Na(+) is the carboxylate cluster of site 1, the highest affinity site of enolase.  相似文献   

10.
The carbon-13 NMR spectrum of oxaloacetate bound in the active site of citrate synthase has been obtained at 90.56 MHz. In the binary complex with enzyme, the positions of the resonances of oxaloacetate are shifted relative to those of the free ligand as follows: C-1 (carboxylate), -2.5 ppm; C-2 (carbonyl), +4.3 ppm; C-3 (methylene), -0.6 ppm; C-4 (carboxylate), +1.3 ppm. The change observed in the carbonyl chemical shift is successively increased in ternary complexes with the product [coenzyme A (CoA)], a substrate analogue (S-acetonyl-CoA), and an acetyl-CoA enolate analogue (carboxymethyl-CoA), reaching a value of +6.8 ppm from the free carbonyl resonance. Binary complexes are in intermediate to fast exchange on the NMR time scale with free oxaloacetate; ternary complexes are in slow exchange. Line widths of the methylene resonance in the ternary complexes suggest complete immobilization of oxaloacetate in the active site. Analysis of line widths in the binary complex suggests the existence of a dynamic equilibrium between two or more forms of bound oxaloacetate, primarily involving C-4. The changes in chemical shifts of the carbonyl carbon indicate strong polarization of the carbonyl bond or protonation of the carbonyl oxygen. Some of this carbonyl polarization occurs even in the binary complex. Development of positive charge on the carbonyl carbon enhances reactivity toward condensation with the carbanion/enolate of acetyl-CoA in the mechanism which has been postulated for this enzyme. The very large change in the chemical shift of the reacting carbonyl in the presence of an analogue of the enolate of acetyl-CoA supports this interpretation.  相似文献   

11.
Chloride salts of Li+, Na+, K+, Mg2+, Ca2+, Cr3+, Mn2+, Fe2+, and Fe3+ had no effect on [3H]diazepam binding. Chloride salts of Co2+, Ni2+, Cu2+, and Zn2+ increased [3H]diazepam binding by 34 to 68% in a concentration-dependent fashion. Since these divalent cations potentiated the GABA-enhanced [3H]diazepam binding and the effect of each divalent cation was nearly additive with GABA, these cations probably act at a site different from the GABA recognition site in the benzodiazepine-receptor complex. Scatchard plots of [3H]diazepam binding without an effective divalent cation showed a single class of binding, with a Kd value of 5.3 mM. In the presence of 1 mM Co2+, Ni2+, Cu2+, or Zn2+, two distinct binding sites were evident with apparent Kd values of 1.0 nM and 5.7 nM. The higher-affinity binding was not detected in the absence of an effective divalent cation and is probably a novel, super-high-affinity binding site.  相似文献   

12.
T G Chu  M Orlowski 《Biochemistry》1984,23(16):3598-3603
A soluble metalloendopeptidase isolated from rat brain preferentially cleaves bonds in peptides having aromatic residues in the P1 and P2 position. An additional aromatic residue in the P3' position greatly increases the binding affinity of the substrate, suggesting the presence of an extended substrate recognition site in the enzyme, capable of binding a minimum of five amino acid residues [Orlowski, M., Michaud, C., & Chu, T.G. (1983) Eur. J. Biochem. 135, 81-88]. A series of N-carboxymethyl peptide derivatives structurally related to model substrates and containing a carboxylate group capable of coordinating with the active site zinc atom were synthesized and tested as potential inhibitors. One of these inhibitors, N-[1(RS)-carboxy-2-phenylethyl]-Ala-Ala-Phe-p-aminobenzoate, was found to be a potent competitive inhibitor of the enzyme with a Ki of 1.94 microM. The two diastereomers of this inhibitor were separated by high-pressure liquid chromatography. The more potent diastereomer had a Ki of 0.81 microM. The inhibitory potency of the less active diastereomer was lower by 1 order of magnitude. Decreasing the hydrophobicity of the residue binding the S1 subsite of the enzyme by, for example, replacement of the phenylethyl group with a methyl residue decreased the inhibitory potency by almost 2 orders of magnitude. Deletion of the carboxylate group decreased the inhibitory potency by more than 3 orders of magnitude. Shortening the inhibitor chain by a single alanine residue had a similar effect. Binding of the inhibitor to the enzyme increased its thermal stability.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The "family 9 polysaccharide lyase" pectate lyase L (Pel9A) from Erwinia chrysanthemi comprises a 10-coil parallel beta-helix domain with distinct structural features including an asparagine ladder and aromatic stack at novel positions within the superhelical structure. Pel9A has a single high affinity calcium-binding site strikingly similar to the "primary" calcium-binding site described previously for the family Pel1A pectate lyases, and there is strong evidence for a common second calcium ion that binds between enzyme and substrate in the "Michaelis" complex. Although the primary calcium ion binds substrate in subsite -1, it is the second calcium ion, whose binding site is formed by the coming together of enzyme and substrate, that facilitates abstraction of the C5 proton from the sacharride in subsite +1. The role of the second calcium is to withdraw electrons from the C6 carboxylate of the substrate, thereby acidifying the C5 proton facilitating its abstraction and resulting in an E1cb-like anti-beta-elimination mechanism. The active site geometries and mechanism of Pel1A and Pel9A are closely similar, but the catalytic base is a lysine in the Pel9A enzymes as opposed to an arginine in the Pel1A enzymes.  相似文献   

14.
Little information is available on the role of Na+, K+, and Cl- in the initial event of uptake of substrates by the dopamine transporter, i.e., the recognition step. In this study, substrate recognition was studied via the inhibition of binding of [3H]WIN 35,428 [2beta-carbomethoxy-3beta-(4-fluorophenyl)[3H]tropane], a cocaine analogue, to the human dopamine transporter in human embryonic kidney 293 cells. D-Amphetamine was the most potent inhibitor, followed by p-tyramine and, finally, dl-octopamine; respective affinities at 150 mM Na+ and 140 mM Cl- were 5.5, 26, and 220 microM. For each substrate, the decrease in the affinity with increasing [K+] could be fitted to a competitive model involving the same inhibitory cation site (site 1) overlapping with the substrate domain as reported by us previously for dopamine. K+ binds to this site with an apparent affinity, averaged across substrates, of 9, 24, 66, 99, and 134 mM at 2, 10, 60, 150, and 300 mM Na+, respectively. In general, increasing [Na+] attenuated the inhibitory effect of K+ in a manner that deviated from linearity, which could be modeled by a distal site for Na+, linked to site 1 by negative allosterism. The presence of Cl- did not affect the binding of K+ to site 1. Models assuming low binding of substrate in the absence of Na+ did not provide fits as good as models in which substrate binds in the absence of Na+ with appreciable affinity. The binding of dl-octopamine and p-tyramine was strongly inhibited by Na+, and stimulated by Cl- only at high [Na+] (300 mM), consonant with a stimulatory action of Cl- occurring through Na+ disinhibition.  相似文献   

15.
We have investigated the binding of 3-[125I]iododizocilpine ([125I]iodo-MK-801) to the N-methyl-D-aspartate (NMDA) receptor in well-washed rat brain membranes. [125I]Iododizocipline binding was displaced by the following: dizocilpine greater than thienylphencyclidine greater than phencyclidine greater than ketamine. Binding of [125I]iododizocilpine was enhanced by glutamate, glycine, and spermidine, whose actions could be reversed by CGS-19755, 7-chlorokynurenate, and arcaine, respectively. [125I]Iododizocilpine binding was also enhanced by a number of divalent cations, including Ba2+, Ca2+, Mg2+, Mn2+, and Sr2+, and several monovalent cations, including Na+ and K+. These cations enhanced [125I]iododizocilpine binding by an action at the polyamine site. In addition, the inhibitory effects associated with high concentrations of these cations was markedly reduced compared to those found in previous studies with [3H]dizocilpine. Analysis of the ability of spermidine, Mg2+, and Sr2+ to alter the inhibition of [125I]iododizocilpine by arcaine gave pA2 values of 5.41, 4.47, and 4.93, corresponding to EC50 concentrations of 3.9, 34.7, and 12.0 microM, respectively, suggesting that physiological concentrations of Mg2+ may occupy the polyamine site. These results demonstrate that [125I]iododizocilpine is a useful probe for the NMDA receptor. Moreover, its high specific activity and relative insensitivity to the inhibitory actions of divalent cations should make [125I]iododizocilpine a valuable ligand for the study of NMDA receptors in intact cellular systems.  相似文献   

16.
H Kim  W N Lipscomb 《Biochemistry》1990,29(23):5546-5555
O-[[(1R)-[[N-(Phenylmethoxycarbonyl)-L-alanyl]amino]ethyl] hydroxyphosphinyl]-L-3-phenyllacetate [ZAAP(O)F], an analogue of (benzyloxycarbonyl)-Ala-Ala-Phe or (benzyloxycarbonyl)-Ala-Ala-phenyllactate, binds to carboxypeptidase A with great affinity (Ki = 3 pM). Similar phosphonates have been shown to be transition-state analogues of the CPA-catalyzed hydrolysis [Hanson, J. E., Kaplan, A. P., & Bartlett, P. A. (1989) Biochemistry 28, 6294-6305]. In the present study, the structure of the complex of this phosphonate with carboxypeptidase A has been determined by X-ray crystallography to a resolution of 2.0 A. The complex crystallizes in the space group P2(1)2(1)2(1) with cell dimensions a = 61.9 A, b = 67.2 A, and c = 76.2 A. The structure of the complex was solved by molecular replacement. Refinement of the structure against 20,776 unique reflections between 10.0 and 2.0 A yields a crystallographic residual of 0.193, including 140 water molecules. The two phosphinyl oxygens of the inhibitor bind to the active-site zinc at 2.2 A on the electrophilic (Arg-127) side and 3.1 A on the nucleophilic (Glu-270) side. Various features of the binding mode of this phosphonate inhibitor are consistent with the hypothesis that carboxypeptidase A catalyzed hydrolysis proceeds through a general-base mechanism in which the carbonyl carbon of the substrate is attacked by Zn-hydroxyl (or Zn-water). An unexpected feature of the bound inhibitor, the cis carbamoyl ester bond at the benzyloxycarbonyl linkage to alanine, allows the benzyloxycarbonyl phenyl ring of the inhibitor to interact favorably with Tyr-198. This complex structure is compared with previous structures of carboxypeptidase A, including the complexes with the potato inhibitor, a hydrated keto methylene substrate analogue, and a phosphonamidate inhibitor. Comparisons are also made with the complexes of thermolysin with some phosphonamidate inhibitors.  相似文献   

17.
Weyand M  Schlichting I 《Biochemistry》1999,38(50):16469-16480
We used freeze trapping to stabilize the Michaelis complex of wild-type tryptophan synthase and the alpha-subunit substrate indole-3-glycerol phosphate (IGP) and determined its structure to 1. 8 A resolution. In addition, we determined the 1.4 A resolution structure of the complex with indole-3-propanole phosphate (IPP), a noncleavable IGP analogue. The interaction of the 3'-hydroxyl of IGP with the catalytic alphaGlu49 leads to a twisting of the propane chain and to a repositioning of the indole ring compared to IPP. Concomitantly, the catalytic alphaAsp60 rotates resulting in a translocation of the COMM domain [betaGly102-betaGly189, for definition see Schneider et al. (1998) Biochemistry 37, 5394-5406] in a direction opposite to the one in the IPP complex. This results in loss of the allosteric sodium ion bound at the beta-subunit and an opening of the beta-active site, thereby making the cofactor pyridoxal 5'-phosphate (PLP) accessible to solvent and thus serine binding. These findings form the structural basis for the information transfer from the alpha- to the beta-subunit and may explain the affinity increase of the beta-active site for serine upon IGP binding.  相似文献   

18.
Malic enzymes catalyze the oxidative decarboxylation of L-malate to pyruvate and CO(2) with the reduction of the NAD(P)(+) cofactor in the presence of divalent cations. We report the crystal structures at up to 2.1 A resolution of human mitochondrial NAD(P)(+)-dependent malic enzyme in different pentary complexes with the natural substrate malate or pyruvate, the dinucleotide cofactor NAD(+) or NADH, the divalent cation Mn(2+), and the allosteric activator fumarate. Malate is bound deep in the active site, providing two ligands for the cation, and its C4 carboxylate group is out of plane with the C1-C2-C3 atoms, facilitating decarboxylation. The divalent cation is positioned optimally to catalyze the entire reaction. Lys183 is the general base for the oxidation step, extracting the proton from the C2 hydroxyl of malate. Tyr112-Lys183 functions as the general acid-base pair to catalyze the tautomerization of the enolpyruvate product from decarboxylation to pyruvate.  相似文献   

19.
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway.  相似文献   

20.
The cephalosporin derivatives L 658758 [1-[[3-(acetoxymethyl)-7 alpha-methoxy-8-oxo-5-thia-1-azabicyclo [4.2.0]oct-2-en-2-yl]carbonyl]proline S,S-dioxide] and L 659286 [1-[[7 alpha-methoxy-8-oxo-3-[[(1,2,5,6-tetrahydro-2-methyl-5,6-dioxo- 1,2,4-triazin-3-yl)thio]methyl]-5-thia-1-aza-(6R)-bicyclo[4.2.0]-o ct-2-en-2-yl]carbonyl]pyrrolidine S,S-dioxide] are mechanism based inhibitors of human leukocyte elastase (HLE). The mechanism involves initial formation of a Michaelis complex followed by acylation of the active site serine. The group on the 3'-methylene is liberated during the course of these reactions, followed by partitioning of an intermediate between hydrolysis to regenerate active enzyme and further modification to produce a stable HLE-inhibitor complex. The partition ratio of 2.0 obtained for the reaction with L 658758 approaches that of an optimal inhibitor. These compounds are functionally irreversible inhibitors as the recovery of activity after inactivation is slow. The half-lives at 37 degrees C of the L 658758 and L 659286 derived HLE-I complexes were 9 and 6.5 h, respectively. The complexes produced by both inhibitors are similar chemically since the thermodynamic parameters for activation to regenerate active enzyme are essentially identical. The free energy of activation for this process is dominated primarily by the enthalpy term. The stability of the final complexes likely arises from Michael addition on the active site histidine to the 3'-methylene.  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