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.
The glutamate dehydrogenase-NADPH-alpha-ketoglutarate complex, an active intermediate on the reaction pathway has a number of unusual properties: 1) it is the only blue-shifted natural complex of this enzyme; 2) it has an anomalously slow rate of dissociation; 3) its off-rate shows a substantial pH-independent D2O solvent isotope effect not exhibited by any other ternary complex of this enzyme; and 4) it has an unusually large enthalpy of interaction parameter. These properties must be ascribable to at least one of the two possibilities conferred on the complex by the presence of the alpha-carbonyl group of alpha-ketoglutarate; the ability to engage in carbonyl addition reactions; and/or the ability to form a specific hydrogen bond. Oxalylglycine, a competitive inhibitor of alpha-ketoglutarate in this enzyme-catalyzed reaction, provides a means of discriminating between these two modes of action. The structure of oxalylglycine provides a dicarboxylic compound which has the same intercarboxylate proton distance and has a carbonyl group in a position spatially analogous to that of alpha-ketoglutarate. Its carbonyl group, however, is that of an amide group and cannot, therefore, engage in carbonyl addition reactions, but can hydrogen bond. Therefore, any effects observed with both oxalylglycine and alpha-ketoglutarate must be ascribed to formation of specific alpha-carbonyl hydrogen bonding, whereas any effects observed with alpha-ketoglutarate alone must be due to an alpha-carbonyl addition reaction. We have used this logic to test the source of the four phenomena listed above. In each case, oxalylglycine and alpha-ketoglutarate showed the same effect. Therefore, we conclude that all four phenomena are in fact due to the formation of a specific alpha-carbonyl hydrogen bond and that the specific carbonyl addition reaction between alpha-ketoglutarate and an enzyme lysine group, postulated in one proposed catalytic mechanism, does not occur.  相似文献   

2.
The reductive amination of alpha-ketoglutarate, catalyzed by bovine liver glutamate dehydrogenase, is inhibited by various anions. Formate and acetate ions are competitive with alpha-ketoglutarate. The pH dependence of the pKi profiles for these anions reveals that they bind to the enzyme-NADPH complex only when an enzymatic residue of pK 8.0 +/- 0.1 in the binary complex is protonated. The ionization of this residue has a delta Hion of 15 +/- 4 kcal/mol. These pK and delta Hion values are not significantly different from those observed in the same complex for the enzyme group which binds the gamma-CO2- of alpha-ketoglutarate and oxalylglycine. It is concluded that formate and acetate also bind to the gamma-carboxylate site in enzyme-NADPH. The Ki values for formate and acetate in a buffer containing 0.1 M phosphate are 20 +/- 4 and 32 +/- 5 mM, respectively, when the pK 8.0 group is fully protonated. Phosphate and trifluoroacetate also show an inhibitory effect, while valerate and sulfate have little effect on the reductive amination rates. The results suggest that specific anions can bind to the gamma-carboxylate site by ionic interactions and alter the kinetic and thermodynamic parameters of the glutamate dehydrogenase-NADPH complex in significant ways.  相似文献   

3.
Human estrogenic 17beta-hydroxysteroid dehydrogenase (17beta-HSD1), a member of the short chain dehydrogenase/reductase (SDR) family, is responsible for the biosynthesis of all active estrogens. The crystal structures of two C19-steroid ternary complexes (17beta-HSD1-androstanedione-NADP and 17beta-HSD1-androstenedione-NADP) reveal the critical role of Leu149 in regulating the substrate specificity and provide novel insight into the different fates of a conserved glutamate residue in the estrogen-specific proteins upon the binding of the keto and hydroxyl groups of steroids. The whole NADP molecule can be unambiguously defined in the NADP binary complex, whereas both ternary complexes show that the nicotinamide moiety of NADP cannot be located in the density maps. In both ternary complexes, the expected position of carboxamide oxygen of NADP is occupied by a water molecule, which makes a bifurcated hydrogen bond with the O3 of C19-steroid and the main chain nitrogen of Val188. These results demonstrate that the hydrogen bonding interaction between the main chain amide group and the carboxamide group of NAD(P)(H) plays an important role in anchoring the nicotinamide ring to the enzyme. This finding is substantiated by structural analyses of all 33 NAD(P)(H) complexes of different SDR proteins, because 29 structures of 33 show this interaction. This common feature reveals a general mechanism among the SDR family, providing a rational basis for inhibitor design against biologically relevant SDR targets.  相似文献   

4.
Arginine kinase (AK) is a key enzyme for energetic balance in invertebrates. Although AK is a well-studied system that provides fast energy to invertebrates using the phosphagen phospho-arginine, the structural details on the AK-arginine binary complex interaction remain unclear. Herein, we determined two crystal structures of the Pacific whiteleg shrimp (Litopenaeus vannamei) arginine kinase, one in binary complex with arginine (LvAK-Arg) and a ternary transition state analog complex (TSAC). We found that the arginine guanidinium group makes ionic contacts with Glu225, Cys271 and a network of ordered water molecules. On the zwitterionic side of the amino acid, the backbone amide nitrogens of Gly64 and Val65 coordinate the arginine carboxylate. Glu314, one of proposed acid–base catalytic residues, did not interact with arginine in the binary complex. This residue is located in the flexible loop 310–320 that covers the active site and only stabilizes in the LvAK-TSAC. This is the first binary complex crystal structure of a guanidine kinase in complex with the guanidine substrate and could give insights into the nature of the early steps of phosphagen biosynthesis.  相似文献   

5.
6.
Binding experiments indicate that mitochondrial aspartate aminotransferase can associate with the alpha-ketoglutarate dehydrogenase complex and that mitochondrial malate dehydrogenase can associate with this binary complex to form a ternary complex. Formation of this ternary complex enables low levels of the alpha-ketoglutarate dehydrogenase complex, in the presence of the aminotransferase, to reverse inhibition of malate oxidation by glutamate. Thus, glutamate can react with the aminotransferase in this complex without glutamate inhibiting production of oxalacetate by the malate dehydrogenase in the complex. The conversion of glutamate to alpha-ketoglutarate could also be facilitated because in the trienzyme complex, oxalacetate might be directly transferred from malate dehydrogenase to the aminotransferase. In addition, association of malate dehydrogenase with these other two enzymes enhances malate dehydrogenase activity due to a marked decrease in the Km of malate. The potential ability of the aminotransferase to transfer directly alpha-ketoglutarate to the alpha-ketoglutarate dehydrogenase complex in this multienzyme system plus the ability of succinyl-CoA, a product of this transfer, to inhibit citrate synthase could play a role in preventing alpha-ketoglutarate and citrate from accumulating in high levels. This would maintain the catalytic activity of the multienzyme system because alpha-ketoglutarate and citrate allosterically inhibit malate dehydrogenase and dissociate this enzyme from the multienzyme system. In addition, citrate also competitively inhibits fumarase. Consequently, when the levels of alpha-ketoglutarate and citrate are high and the multienzyme system is not required to convert glutamate to alpha-ketoglutarate, it is inactive. However, control by citrate would be expected to be absent in rapidly dividing tumors which characteristically have low mitochondrial levels of citrate.  相似文献   

7.
Malate dehydrogenase specifically oxidizes malate to oxaloacetate. The specificity arises from three arginines in the active site pocket that coordinate the carboxyl groups of the substrate and stabilize the newly forming hydroxyl/keto group during catalysis. Here, the role of Arg-153 in distinguishing substrate specificity is examined by the mutant R153C. The x-ray structure of the NAD binary complex at 2.1 A reveals two sulfate ions bound in the closed form of the active site. The sulfate that occupies the substrate binding site has been translated approximately 2 A toward the opening of the active site cavity. Its new location suggests that the low catalytic turnover observed in the R153C mutant may be due to misalignment of the hydroxyl or ketone group of the substrate with the appropriate catalytic residues. In the NAD.pyruvate ternary complex, the monocarboxylic inhibitor is bound in the open conformation of the active site. The pyruvate is coordinated not by the active site arginines, but through weak hydrogen bonds to the amide backbone. Energy minimized molecular models of unnatural analogues of R153C (Wright, S. K., and Viola, R. E. (2001) J. Biol. Chem. 276, 31151-31155) reveal that the regenerated amino and amido side chains can form favorable hydrogen-bonding interactions with the substrate, although a return to native enzymatic activity is not observed. The low activity of the modified R153C enzymes suggests that precise positioning of the guanidino side chain is essential for optimal orientation of the substrate.  相似文献   

8.
We have solved the x-ray structures of the binary horseradish peroxidase C-ferulic acid complex and the ternary horseradish peroxidase C-cyanide-ferulic acid complex to 2.0 and 1.45 A, respectively. Ferulic acid is a naturally occurring phenolic compound found in the plant cell wall and is an in vivo substrate for plant peroxidases. The x-ray structures demonstrate the flexibility and dynamic character of the aromatic donor binding site in horseradish peroxidase and emphasize the role of the distal arginine (Arg(38)) in both substrate oxidation and ligand binding. Arg(38) hydrogen bonds to bound cyanide, thereby contributing to the stabilization of the horseradish peroxidase-cyanide complex and suggesting that the distal arginine will be able to contribute with a similar interaction during stabilization of a bound peroxy transition state and subsequent O-O bond cleavage. The catalytic arginine is additionally engaged in an extensive hydrogen bonding network, which also includes the catalytic distal histidine, a water molecule and Pro(139), a proline residue conserved within the plant peroxidase superfamily. Based on the observed hydrogen bonding network and previous spectroscopic and kinetic work, a general mechanism of peroxidase substrate oxidation is proposed.  相似文献   

9.
Interactions between the alpha-helix peptide dipoles and charged groups close to the ends of the helix were found to be an important determinant of alpha-helix stability in a previous study. The charge on the N-terminal residue of the C-peptide from ribonuclease A was varied chiefly by changing the alpha-NH2 blocking group, and the correlation of helix stability with N-terminal charge was demonstrated. An alternative explanation for some of those results is that the succinyl and acetyl blocking groups stabilize the helix by hydrogen bonding to an unsatisfied main-chain NH group. The helix dipole model is tested here with peptides that contain either a free alpha-NH3+ or alpha-COO- group, and no other charged groups that would titrate with similar pKa's. This model predicts that alpha-NH3+ and alpha-COO- groups are helix-destabilizing and that the destabilizing interactions are electrostatic in origin. The hydrogen bonding model predicts that alpha-NH3+ and alpha-COO- groups are not themselves helix-destabilizing, but that an acetyl or amide blocking group at the N- or C-terminus, respectively, stabilizes the helix by hydrogen bonding to an unsatisfied main-chain NH or CO group. The results are as follows: (1) Removal of the charge from alpha-NH3+ and alpha-COO- groups by pH titration stabilizes an alpha-helix. (2) The increase in helix stability on pH titration of these groups is close to the increase produced by adding an acetyl or amide blocking group.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The cooperativity between hydrogen and halogen bonds in XY···HNC···XY (X, Y = F, Cl, Br) complexes was studied at the MP2/aug-cc-pVTZ level. Two hydrogen-bonded dimers, five hydrogen-bonded dimers, and ten trimers were obtained. The hydrogen- and halogen-bonded interaction energies in the trimers were larger than those in the dimers, indicating that both the hydrogen bonding interaction and the halogen bonding interaction are enhanced. The binary halogen bonding interaction plays the most important role in the ternary system. The hydrogen donor molecule influences the magnitude of the halogen bonding interaction much more than the hydrogen bonding interaction in the trimers with respect to the dimers. Our calculations are consistent with the conclusion that the stronger noncovalent interaction has a bigger effect on the weaker one. The variation in the vibrational frequency in the HNC molecule was considered. The NH antisymmetry vibration frequency has a blue shift, whereas the symmetry vibration frequency has a red shift. A dipole moment enhancement is observed upon formation of the trimers. The variation in topological properties at bond critical points was obtained using the atoms in molecules method, and was consistent with the results of the interaction energy analysis.  相似文献   

11.
12.
Hydrogen bonding stabilizes globular proteins.   总被引:8,自引:1,他引:7       下载免费PDF全文
It is clear that intramolecular hydrogen bonds are essential to the structure and stability of globular proteins. It is not clear, however, whether they make a net favorable contribution to this stability. Experimental and theoretical studies are at odds over this important question. Measurements of the change in conformational stability, delta (delta G), for the mutation of a hydrogen bonded residue to one incapable of hydrogen bonding suggest a stabilization of 1.0 kcal/mol per hydrogen bond. If the delta (delta G) values are corrected for differences in side-chain hydrophobicity and conformational entropy, then the estimated stabilization becomes 2.2 kcal/mol per hydrogen bond. These and other experimental studies discussed here are consistent and compelling: hydrogen bonding stabilizes globular proteins.  相似文献   

13.
Andi B  Xu H  Cook PF  West AH 《Biochemistry》2007,46(44):12512-12521
Three structures of saccharopine dehydrogenase (l-lysine-forming) (SDH) have been determined in the presence of sulfate, adenosine monophosphate (AMP), and oxalylglycine (OxGly). In the sulfate-bound structure, a sulfate ion binds in a cleft between the two domains of SDH, occupies one of the substrate carboxylate binding sites, and results in partial closure of the active site of the enzyme due to a domain rotation of almost 12 degrees in comparison to the apoenzyme structure. In the second structure, AMP binds to the active site in an area where the NAD+ cofactor is expected to bind. All of the AMP moieties (adenine ring, ribose, and phosphate) interact with specific residues of the enzyme. In the OxGly-bound structure, carboxylates of OxGly interact with arginine residues representative of the manner in which substrate (alpha-ketoglutarate and saccharopine) may bind. The alpha-keto group of OxGly interacts with Lys77 and His96, which are candidates for acid-base catalysis. Analysis of ligand-enzyme interactions, comparative structural analysis, corroboration with kinetic data, and discussion of a ternary complex model are presented in this study.  相似文献   

14.
Different modes of binding of transition state mimics: amide, phosphonate and difluoro ketone, to human synovial fluid phospholipase A2 (HSF PLA2) are studies by molecular dynamics simulations computed in solvent. The results are analysed in the light of primary binding sites. Hydrogen bonding interaction plays an important role for amino acids such as Gly32, Val30, and Glu55, apart from the well known active site residues viz Asp48, Gly25, Gly29, Gly31, His27, His47, Lys62, Phe23, Asn114 and Tyr112. In addition, the hydrogen bonding interaction between Sn-1 tetrahedral phosphonate group of amide and difluoro ketone inhibitors and crystallographic water molecules (H2O 523, H2O 524 and H2O 401) seems to have a significant role. Many of the active site charged residues display considerable movement upon ligand binding. The structural effects of ligand binding were analyzed from RMS deviations of Cα in the resulting energy-minimized average structures of the receptor–ligand complexes. The values of the RMS deviations differ among the HSF PLA2s, in a pattern that is not the same for the three complexes. This suggests that ligands with different pharmacological efficacies induce different types of conformational changes of the receptor. Our active-orientation model is, at least qualitatively, consistent with experimental data and should be useful for the rational design of more potent inhibitors.  相似文献   

15.
16.
Values of K, delta G(o), delta H(o), delta S(o) and delta C(po) for the binding reaction of small organic ligands forming 1:1 complexes with either alpha- or beta-cyclodextrin were obtained by titration calorimetry from 15 degrees C to 45 degrees C. A hydrogen bond or hydrophobic interaction was introduced by adding a single functional group to the ligand. The thermodynamics of binding with and without the added group are compared to estimate the contribution of the hydrogen bond or hydrophobic interaction. A change in the environment of a functional group is required to influence the binding thermodynamics, but molecular size-dependent solute-solvent interactions have no effect. For phenolic O-H-O hydrogen bond formation, delta H(o) varies from -2 to -1.4 kcal mol(-1) from 15 degrees C to 45 degrees C, and delta C(p) is increased by 18 cal K(-1) mol(-1). The hydrophobic interaction has an opposite effect: in alpha-cyclodextrin, delta C(po) = -13.3 cal K(-1) mol(-1) per ligand -CH(2)-, identical to values found for the transfer of a -CH(2)-group from water to a nonpolar environment. At room temperature, the hydrogen bond and the -CH(2)-interaction each contribute about -600 cal mol(-1) to the stability (delta G(o)) of the complex. With increased temperature, the hydrogen bond stability decreases (i.e., hydrogen bonds "melt"), but the stability of the hydrophobic interaction remains essentially constant.  相似文献   

17.
Cosgrove S  Rogers L  Hewage CM  Malthouse JP 《Biochemistry》2007,46(39):11205-11215
Z-Ala-Ala-Phe-glyoxal (where Z is benzyloxycarbonyl) has been shown to be a competitive inhibitor of pepsin with a Ki = 89 +/- 24 nM at pH 2.0 and 25 degrees C. Both the ketone carbon (R13COCHO) and the aldehyde carbon (RCO13CHO) of the glyoxal group of Z-Ala-Ala-Phe-glyoxal have been 13C-enriched. Using 13C NMR, it has been shown that when the inhibitor is bound to pepsin, the glyoxal keto and aldehyde carbons give signals at 98.8 and 90.9 ppm, respectively. This demonstrates that pepsin binds and preferentially stabilizes the fully hydrated form of the glyoxal inhibitor Z-Ala-Ala-Phe-glyoxal. From 13C NMR pH studies with glyoxal inhibitor, we obtain no evidence for its hemiketal or hemiacetal hydroxyl groups ionizing to give oxyanions. We conclude that if an oxyanion is formed its pKa must be >8.0. Using 1H NMR, we observe four hydrogen bonds in free pepsin and in pepsin/Z-Ala-Ala-Phe-glyoxal complexes. In the pepsin/pepstatin complex an additional hydrogen bond is formed. We examine the effect of pH on hydrogen bond formation, but we do not find any evidence for low-barrier hydrogen bond formation in the inhibitor complexes. We conclude that the primary role of hydrogen bonding to catalytic tetrahedral intermediates in the aspartyl proteases is to correctly orientate the tetrahedral intermediate for catalysis.  相似文献   

18.
Carbon assimilation in plants is regulated by the reduction of specific protein disulfides by light and their re-oxidation in the dark. The redox switch CP12 is an intrinsically disordered protein that can form two disulfide bridges. In the dark oxidized CP12 forms an inactive supramolecular complex with glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and phosphoribulokinase, two enzymes of the carbon assimilation cycle. Here we show that binding of CP12 to GAPDH, the first step of ternary complex formation, follows an integrated mechanism that combines conformational selection with induced folding steps. Initially, a CP12 conformation characterized by a circular structural motif including the C-terminal disulfide is selected by GAPDH. Subsequently, the induced folding of the flexible C-terminal tail of CP12 in the active site of GAPDH stabilizes the binary complex. Formation of several hydrogen bonds compensates the entropic cost of CP12 fixation and terminates the interaction mechanism that contributes to carbon assimilation control.  相似文献   

19.
All-atom simulation data are presented for ternary mixtures of palmitoyl sphingomyelin (PSM), cholesterol, and either palmitoyl oleoyl phosphatidyl choline or dioleoyl phosphatidyl choline (DOPC). For comparison, data for a mixture of dipalmitoyl phosphatidyl choline (DPPC), cholesterol, and DOPC are also presented. Compositions corresponding to the liquid-ordered phase, the liquid-disordered phase, and coexistence of the two phases are simulated for each mixture. Within the liquid-ordered phase, cholesterol is preferentially solvated by DOPC if it is available, but if DOPC is replaced by POPC, cholesterol is preferentially solvated by PSM. In the DPPC mixtures, cholesterol interacts preferentially with the saturated chains via its smooth face, whereas in the PSM mixtures, cholesterol interacts preferentially with PSM via its rough face. Interactions between cholesterol and PSM have a very particular character: hydrogen bonding between cholesterol and the amide of PSM rotates the tilt of the amide plane, which primes it for more robust hydrogen bonding with other PSM. Cholesterol-PSM hydrogen bonding also locally modifies the hexagonal packing of hydrocarbon chains in the liquid-ordered phase of PSM mixtures.  相似文献   

20.
M R Eftink  K Bystr?m 《Biochemistry》1986,25(21):6624-6630
The association of the coenzyme NAD+ to liver alcohol dehydrogenase (LADH) is known to be pH dependent, with the binding being linked to the shift in the pK of some group on the protein from a value of 9-10, in the free enzyme, to 7.5-8 in the LADH-NAD+ binary complex. We have further characterized the nature of this linkage between NAD+ binding and proton dissociation by studying the pH dependence (pH range 6-10) of the proton release, delta n, and enthalpy change, delta Ho(app), for formation of both binary (LADH-NAD+) and ternary (LADH-NAD+-I, where I is pyrazole or trifluoroethanol) complexes. The pH dependence of both delta n and delta Ho(app) is found to be consistent with linkage to a single acid dissociating group, whose pK is perturbed from 9.5 to 8.0 upon NAD+ binding and is further perturbed to approximately 6.0 upon ternary complex formation. The apparent enthalpy change for NAD+ binding is endothermic between pH 7 and pH 10, with a maximum at pH 8.5-9.0. The pH dependence of the delta Ho(app) for both binary and ternary complex formation is consistent with a heat of protonation of -7.5 kcal/mol for the coupled acid dissociating group. The intrinsic enthalpy changes for NAD+ binding and NAD+ plus pyrazole binding to LADH are determined to be approximately 0 and -11.0 kcal/mol, respectively. Enthalpy change data are also presented for the binding of the NAD+ analogues adenosine 5'-diphosphoribose and 3-acetylpyridine adenine dinucleotide.  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