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.
Equilibrium and calorimetric studies of substrate binding to turkey egg white (TEW) lysozyme were carried out at 30degrees as a function of pH (2 to 9) and ligand size (monosaccharide to hexasaccharide of N-acetylglucosamine). Steady state kinetic measurements using the N-acetylglucosamine hexasaccharide were carried out as a function of pH (2 to 9) and temperature (20-60degrees). These experiments allow comparison of the properties of TEW lysozyme with those of the hen egg white (HEW) enzyme reported previously (Banerjee, S. K., Holler, E., Hess, G. P., and Rupley, J. A. (1975) J. Biol. Chem. 250, 4355-4367, and references therein). The free energies and enthalpies of oligosaccharide binding are the same for TEW and HEW lysozymes at pH 2 but are less negative for TEW lysozyme at pH 5. The pH dependence of the binding of (GlcNAc)3 and higher oligomers to TEW lysozyme is like that for the binding of beta-methyl-N-acetylglucosaminide to TEW lysozyme. These data indicate that oligosaccharide ligands bind identically with HEW and TEW lysozymes, except for the interactions of residue 101, which is aspartic acid in the HEW protein and glycine in the TEW protein (Larue, J. N., and Speck, J. C., Jr. (1970) J. Biol. Chem. 245, 1985-1991). The pH dependence of kcat is described by apparent pK values of 3.9 and 6.8 and a maximum value of kcat of 0.135 s-1. A value of 21.0 kcal/mol was calculated for deltaH from the temperature dependence of kcat. These values and the dependence of the transglycosylation reaction on acceptor concentration are within experimental error the same as those for HEW lysozyme. The more acid pK seen in the pH rate profile reflects the ionization of Asp-52 in the lysozyme-(GlcNAc)6 complex. The pK of Asp-52 in the free protein is 0.3 pK unit lower. The essential identity of the active sites of the HEW and TEW enzymes, except for the Asp-101 interactions, allows estimation of the thermodynamic properties associated with formation of the two hydrogen bonds between Asp-101 and substrate as deltaG0 = -1.2 kcal/mol, DeltaH0 = -3.6 kcal/mol, and deltaS0 = -7.9 e.u.  相似文献   

2.
Enthalpy changes of alpha-chymotrypsin acylation by 3-(2-furyl)acryloylimidazole (FAI) were calorimetrically determined as a function of pH. By observing the functional dependence of acylation enthalpies on buffer ionization heats, a complex pH profile was obtained describing proton release accompanying formation of acyl-enzyme. A pKa of 4.0 for FAI ionization and apparent pKa values of 6.8, 7.55 and 8.8 on the enzyme were used to account for the proton release data. A model which accounts for the proton release behavior was used to fit the acylation enthalpy data and values for the apparent dissociation enthalpies of the groups involved were obtained along with a pH-independent intrinsic enthalpy of acylation. This model suggests a group with an apparent pK = 6.8 and delta Hion = 8.7 kcal/mol which is perturbed to a pK of 7.55 and delta Hion = 7.6 kcal/mol on attachment of the acyl moiety to the enzyme. The apparent ionization enthalpy change for the active-inactive transition (pK3 = 8.8; delta H = 3.0 kcal/mol) corresponds with that calculated from the data of Fersht (J. Mol. Biol. 64 (1972) 497). The pH-independent intrinsic enthalpy of acylation (delta H = -7.9 kcal/mol) is corrected for group ionizations linked to the acylation process. Consequently, it more closely reflects molecular processes of interest such as substrate binding, covalent bond rearrangement, and product release.  相似文献   

3.
We constructed the complexes between HEWL and (GlcNAc)6 oligomer in order to investigate the amino acid residues related to substrate binding in the productive and nonproductive complexes, and the relationship between the distortion of the GlcNAc residue D and the formation of the productive complexes. We obtained 49 HEWL-(GlcNAc)6 complexes by a systematic conformational search and classified the each one to the three binding modes; left side, center, or right side. Furthermore we performed the molecular dynamics simulation against 20 HEWL-(GlcNAc)6 complexes (8: chair model, 12 : half-chair model) selected from the 49 complexes to investigate the interaction between HEWL and (GlcNAc)6. As results, we confirmed that it is necessary for GlcNAc residue D to be half-chaired form to bind toward the right side to form productive complexes. We found newly that eight amino acid residues interact with the (GlcNAc)6 oligomer, as follows, Arg73, Gly102, Asn103 for GlcNAc residue A; Asn103 for GlcNAc residues B and C; Leu56, Ala107, Val109 for GlcNAc residue D; Ala110 for GlcNAc residue E; and Lys33 for GlcNAc residue F. We also indicated that GlcNAc residue F does not interact with Thr47 and rarely interacts with Phe34 and Asn37.  相似文献   

4.
M Fuxreiter  A Warshel  R Osman 《Biochemistry》1999,38(30):9577-9589
T4 Endonuclease V (EndoV) is a base excision repair enzyme that removes thymine dimers (TD) from damaged DNA. To elucidate the role of the active site residues in catalysis, their pK(a)'s were evaluated using the semimicroscopic version of the protein dipoles-Langevin dipoles method (PDLD/S). Contributions of different effects to the pK(a) such as charge-charge interactions, conformational rearrangement, protein relaxation, and DNA binding were analyzed in detail. Charging of the active site residues was found to be less favorable in the complex than in the free enzyme. The pK(a) of the N-terminus decreased from 8.01 in the free enzyme to 6.52 in the complex, while the pK(a) of Glu-23 increased from 1. 52 to 7.82, which indicates that the key residues are neutral in the reactant state of the glycosylase step. These pK(a)'s are in agreement with the optimal pH range of the reaction and support the N-terminus acting as a nucleophile. The Glu-23 in its protonated form is hydrogen bonded to O4' of the sugar of 5' TD and can play a role in increasing the positive charge of C1' and, hence, accelerating the nucleophilic substitution. Furthermore, the neutral Glu-23 is a likely candidate to protonate O4' to induce ring opening required to complete the glycosylase step of EndoV. The positively charged Arg-22 and Arg-26 provide an electrostatically favorable environment for the leaving base. To distinguish between S(N)1 and S(N)2 mechanisms of the glycosylase step the energetics of protonating O2 of 5' TD was calculated. The enzyme was found to stabilize the neutral thymine by approximately 3.6 kcal/mol, whereas it destabilizes the protonated thymine by approximately 6.6 kcal/mol with respect to an aqueous environment. Consequently, the formation of a protonated thymine intermediate is unlikely, indicating an S(N)2 reaction mechanism for the glycosylase step.  相似文献   

5.
The pH dependence of Vmax and Vmax/Km for hydrolysis of Dnp-Pro-Leu-Gly-Leu-Trp-Ala-D-Arg-NH2 at the Gly-Leu bond by porcine synovial collagenase and gelatinase was determined in the pH range 5-10. Both enzymes exhibited bell-shaped dependencies on pH for these two kinetic parameters, indicating that activity is dependent on at least two ionizable groups, one of which must be unprotonated and the other protonated. For collagenase, Vmax/Km data indicate that in the substrate-free enzyme, these groups have apparent pK values of 7.0 and 9.5, while the Vmax profile indicates similar pK values of 6.8 and 10.1 for the enzyme-substrate complex. The corresponding pH profiles of gelatinase were similar to those of collagenase, indicating the importance of groups with apparent pK values of 5.9 and 10.0 for the free enzyme and 5.9 and 11.1 for the enzyme-substrate complex. When these kinetic constants were determined in D2O using the peptide substrate, there was no significant effect on Vmax or Km for collagenase or Km for gelatinase. However, there was a deuterium isotope effect of approximately 1.5 on Vmax for gelatinase. These results indicate that a proton transfer step is not involved in the rate-limiting step for collagenase, but may be limiting with gelatinase. The Arrhenius activation energies for peptide bond hydrolysis of the synthetic peptide as well as the natural substrates were also determined for both enzymes. The activation energy (81 kcal) for hydrolysis of collagen by collagenase was nine times greater than that determined for the synthetic substrate (9.2 kcal). In contrast, the activation energy for hydrolysis of gelatin by gelatinase (26.3 kcal) was only 2.4 times greater than that for the synthetic substrate (11 kcal).  相似文献   

6.
LexA repressor of Escherichia coli and phage lambda repressor are inactivated in vivo and in vitro by specific cleavage of an Ala-Gly peptide bond in reactions requiring RecA protein. At mildly alkaline pH, the in vitro cleavage reaction also proceeds spontaneously, suggesting that peptide bond hydrolysis is an activity of the repressors rather than of RecA. The spontaneous cleavage reaction, termed "autodigestion", has been characterized for the LexA and lambda repressors. The results show that the reaction is intramolecular. The rate of LexA autodigestion was studied over the pH range 7.15-11.77 and over the temperature range 4-46 degrees C. The logarithm of the rate constant increased linearly with pH and reached a plateau value (2.5 X 10(-3) s-1 at 37 degrees C) at pH above 10. The data closely followed a model in which a single residue side chain (apparent pK = 9.8 at 37 degrees C) must be deprotonated for the protein to show activity. Analysis of the temperature dependence gave the heat of proton dissociation as 19.9 kcal/mol and the heat of activation for hydrolysis as 15.3 kcal/mol at 25 degrees C. Autodigestion of lambda repressor, studied over the pH range 8.65-10.70 at 37 degrees C, was similar to the LexA reaction in its pH dependence, yielding a pK of 9.8. The maximum rate at 37 degrees C for lambda repressor, 6.1 X 10(-5) s-1, was 40 times slower than for LexA, a difference similar to that previously observed in vivo and in vitro for RecA-dependent cleavage reactions. There was no significant solvent deuterium isotope effect on the autodigestion of LexA. Changes in buffer composition, including high concentrations of glycine for lambda repressor and of imidazole or hydroxylamine for LexA, indicated that solvent components other than water do not participate in the rate-determining step. Removal or addition of metal ions did not significantly affect LexA autodigestion. These and other observations suggest that the deprotonated form of an amino acid side chain plays a central role in the chemistry of the cleavage reaction. The above observations establish repressor autodigestion as a member of an emerging set of biologically important self-processing reactions.  相似文献   

7.
Laser Raman spectroscopy has been employed to study the structure of the hen egg yolk protein phosvitin in H2O and D2O solutions at neutral and acidic pH (pD) and in the solid state. The Raman data indicate an unusual conformation for phosvitin in neutral aqueous solution, which is deficient in both alpha-helix and conventional beta-sheet conformations. This unusual pH 7 structure is, however, largely converted to a beta-sheet conformation in strongly acidic media (pH less than 2). beta-Sheet is also the predominant secondary structure for phosvitin in the solid state, obtained by lyophilization of the protein from aqueous solution at neutral pH. The imidazolium rings of histidyl residues remain significantly protonated near neutrality, which suggests substantial elevation of the pK for imidazolium ring ionizations of phosvitin in aqueous solution. This may result from extensive ion-pair interactions involving positively charged histidines and negatively charged phosphoserines, which are prevalent in the phosvitin sequence. The present results suggest that antiparallel beta-sheets may not be the secondary structure most characteristic of native phosvitin (physiological pH), even though beta-sheet is the predominant conformation for phosvitin in acidic solutions (pH 1.5) and in the lyophilized solid. Phosvitin appears to be the first protein for which the major component to the Raman amide I band is centered near 1685 cm-1, which is 10-40 cm-1 higher than proteins heretofore examined in aqueous solution by Raman spectroscopy.  相似文献   

8.
The interaction of Escherichia coli RNA polymerase with poly[d(A-T)] and poly[d-(I-C)] was studied by difference absorption spectroscopy at temperatures, from 5 to 45 degrees C in the absence and presence of Mg2+. The effect of KCl concentration, at a fixed temperature, was studied from 12.5 to 400 mM. Difference absorption experiments permitted calculation of the extent of DNA opening induced by RNA polymerase and estimation of the equilibrium constant associated with the isomerization from a closed to an open RNA polymerase-DNA complex. delta H0 and delta S0 for the closed-to-open transition with poly[d(A-T)] or poly[d(I-C)] complexed with RNA polymerase are significantly lower than the values associated with the helix-to-coil transition for the free polynucleotides. For the RNA polymerase complexes with poly[d(A-T)] and poly[d(I-C)] in 50 mM KCl, delta H0 approximately 15-16 kcal/mol (63-67 kJ/mol) and delta S0 approximately 50-57 cal/K per mol (209-239 J/K per mol). The presence of Mg2+ does not change these parameters appreciably for the RNA polymerase-poly[d(A-T)] complex, but for the RNA polymerase-poly[d(I-C)] complex in the presence of Mg2+, the delta H0 and delta S0 values are larger and temperature-dependent, with delta H0 approximately 22 kcal/mol (92 kJ/mol) and delta S0 approximately 72 cal/K per mol (approx. 300 J/K per mol) at 25 degrees C, and delta Cp0 approximately 2 kcal/K per mol (approx. 8.3 kJ/K per mol). The circular dichroism (CD) changes observed for helix opening induced by RNA polymerase are qualitatively consistent with the thermally induced changes observed for the free polynucleotides, supporting the difference absorption method. The salt-dependent studies indicate that two monovalent cations are released upon helix opening. For poly[d(A-T)], the temperature-dependence of enzyme activity correlates well with the helix opening, implying this step to be the rate-determining step. In the case of poly[d(I-C)], the same is not true, and so the rate-determining step must be a process subsequent to helix opening.  相似文献   

9.
Toteva MM  Silvaggi NR  Allen KN  Richard JP 《Biochemistry》2011,50(46):10170-10181
D-Xylose isomerase (XI) and triosephosphate isomerase (TIM) catalyze the aldose-ketose isomerization reactions of D-xylose and d-glyceraldehyde 3-phosphate (DGAP), respectively. D-Glyceraldehyde (DGA) is the triose fragment common to the substrates for XI and TIM. The XI-catalyzed isomerization of DGA to give dihydroxyacetone (DHA) in D(2)O was monitored by (1)H nuclear magnetic resonance spectroscopy, and a k(cat)/K(m) of 0.034 M(-1) s(-1) was determined for this isomerization at pD 7.0. This is similar to the k(cat)/K(m) of 0.017 M(-1) s(-1) for the TIM-catalyzed carbon deprotonation reaction of DGA in D(2)O at pD 7.0 [Amyes, T. L., O'Donoghue, A. C., and Richard, J. P. (2001) J. Am. Chem. Soc. 123, 11325-11326]. The much larger activation barrier for XI-catalyzed isomerization of D-xylose (k(cat)/K(m) = 490 M(-1) s(-1)) versus that for the TIM-catalyzed isomerization of DGAP (k(cat)/K(m) = 9.6 × 10(6) M(-1) s(-1)) is due to (i) the barrier to conversion of cyclic d-xylose to the reactive linear sugar (5.4 kcal/mol) being larger than that for conversion of DGAP hydrate to the free aldehyde (1.7 kcal/mol) and (ii) the intrinsic binding energy [Jencks, W. P. (1975) Adv. Enzymol. Relat. Areas Mol. Biol. 43, 219-410] of the terminal ethylene glycol fragment of D-xylose (9.3 kcal/mol) being smaller than that of the phosphodianion group of DGAP (~12 kcal/mol). The XI-catalyzed isomerization of DGA in D(2)O at pD 7.0 gives a 90% yield of [1-(1)H]DHA and a 10% yield of [1-(2)H]DHA, the product of isomerization with incorporation of deuterium from solvent D(2)O. By comparison, the transfer of (3)H from the labeled hexose substrate to solvent is observed only once in every 10(9) turnovers for the XI-catalyzed isomerization of [2-(3)H]glucose in H(2)O [Allen, K. N., Lavie, A., Farber, G. K., Glasfeld, A., Petsko, G. A., and Ringe, D. (1994) Biochemistry 33, 1481-1487]. We propose that truncation of the terminal ethylene glycol fragment of d-xylose to give DGA results in a large decrease in the rate of XI-catalyzed isomerization with hydride transfer compared with that for proton transfer. An ultra-high-resolution (0.97 ?) X-ray crystal structure was determined for the complex obtained by soaking crystals of XI with 50 mM DGA. The triose binds to XI as the unreactive hydrate, but ligand binding induces metal cofactor movement and conformational changes in active site residues similar to those observed for XI·sugar complexes.  相似文献   

10.
A Bertazzon  T Y Tsong 《Biochemistry》1989,28(25):9784-9790
High-resolution differential scanning calorimetry (DSC) has been employed to study the thermal stability of myosin, its major constitutive fragments (S-1, light chains, and rod), and also reconstituted thick filaments. The thermal denaturation of soluble myosin was complex and was characterized by a multistep endothermic process for the temperature range from 41 to 60 degrees C. The shape of the endotherm was highly dependent on the pH and the ionic strength of the solution, although the delta Hcal (calorimetric enthalpy) of denaturation (1715 +/- 75 kcal/mol) was insensitive to these changes for the solvent conditions used in this study. This value also agrees, within experimental error, with the sum of the denaturation enthalpies obtained for isolated fragments (1724 +/- 79 kcal/mol). In identical conditions of ionic strength, pH, and heating rate, the computer-calculated differential endotherms of domains belonging to S-1 and light chains were superimposable with those of the isolated fragments. Their responses to changes in the solvent condition were also similar. We suggest that the observed functional independence of the major domains in myosin reflects also the independence of their structural stability. The thermal unfolding of the isolated rod was multiphasic and readily reversible (95%). It occurred between 41 and 60 degrees C, with an delta Hcal of 1058 +/- 59 kcal/mol. The melting of S-1 showed a single peak at 46.3 +/- 0.1 degrees C with an delta Hcal of 255 +/- 12 kcal/mol. Light chains melted at 51.0 +/- 0.2 degrees C with an delta Hcal of 85 +/- 15 kcal/mol.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The two most buried carboxyl groups in ribonuclease Sa (RNase Sa) are Asp33 (99% buried; pK 2.4) and Asp79 (85% buried; pK 7.4). Above these pK values, the stability of the D33A variant is 6kcal/mol less than wild-type RNase Sa, and the stability of the D79A variant is 3.3kcal/mol greater than wild-type RNase Sa. The key structural difference between the carboxyl groups is that Asp33 forms three intramolecular hydrogen bonds, and Asp79 forms no intramolecular hydrogen bond. Here, we focus on Asp79 and describe studies of 11 Asp79 variants. Most of the variants were at least 2kcal/mol more stable than wild-type RNase Sa, and the most interesting was D79F. At pH 3, below the pK of Asp79, RNase Sa is 0.3kcal/mol more stable than the D79F variant. At pH 8.5, above the pK of Asp79, RNase Sa is 3.7kcal/mol less stable than the D79F variant. The unfavorable contribution of Asp79 to the stability appears to result from the Born self-energy of burying the charge and, more importantly, from unfavorable charge-charge interactions. To counteract the effect of the negative charge on Asp79, we prepared the Q94K variant and the crystal structure showed that the amino group of the Lys formed a hydrogen-bonded ion pair (distance, 2.71A; angle, 100 degrees ) with the carboxyl group of Asp79. The stability of the Q94K variant was about the same as the wild-type at pH 3, where Asp79 is uncharged, but 1kcal/mol greater than that of wild-type RNase Sa at pH 8.5, where Asp79 is charged. Differences in hydrophobicity, steric strain, Born self-energy, and electrostatic interactions all appear to contribute to the range of stabilities observed in the variants. When it is possible, replacing buried, non-hydrogen bonded, ionizable side-chains with non-polar side-chains is an excellent means of increasing protein stability.  相似文献   

12.
Prabhakar R  Vreven T  Morokuma K  Musaev DG 《Biochemistry》2005,44(35):11864-11871
The mechanism of the hydrogen peroxide reduction by two molecules of glutathione catalyzed by the selenoprotein glutatione peroxidase (GPx) has been computationally studied. It has been shown that the first elementary reaction of this process, (E-SeH) + H(2)O(2) --> (E-SeOH) + H(2)O (1), proceeds via a stepwise pathway with the overall barrier of 17.1 kcal/mol, which is in good agreement with the experimental barrier of 14.9 kcal/mol. During reaction 1, the Gln83 residue has been found to play a key role as a proton acceptor, which is consistent with experiments. The second elementary reaction, (E-SeOH) + GSH --> (E-Se-SG) + HOH (2), proceeds with the barrier of 17.9 kcal/mol. The last elementary reaction, (E-Se-SG) + GSH --> (E-SeH) + GS-SG (3), is initiated with the coordination of the second glutathione molecule. The calculations clearly suggest that the amide backbone of the Gly50 residue directly participates in this reaction and the presence of two water molecules is absolutely vital for the reaction to occur. This reaction proceeds with the barrier of 21.5 kcal/mol and is suggested to be a rate-determining step of the entire GPx-catalyzed reaction H(2)O(2) + 2GSH --> GS-SG + 2H(2)O. The results discussed in the present study provide intricate details of every step of the catalytic mechanism of the GPx enzyme and are in good general agreement with experimental findings and suggestions.  相似文献   

13.
N Ahn  J P Klinman 《Biochemistry》1983,22(13):3096-3106
Dopamine beta-monooxygenase catalyzes a reaction in which 2 mol of protons are consumed for each turnover of substrate. Studies of the pH dependence of initial rate parameters (Vmax and Vmax/Km) and their primary hydrogen isotope effects show that at least two ionizable residues are involved in catalysis. One residue (B1, pK = 5.6-5.8) must be protonated prior to the carbon-hydrogen bond cleavage step, implying a role for general-acid catalysis in substrate activation. A second protonated residue (B2, pK = 5.2-5.4) facilitates, but is not required for, product release. Recent measurement of the intrinsic isotope effect for dopamine beta-monoxygenase [Miller, S. M., & Klinman, J. P. (1983) Biochemistry (preceding paper in this issue)] allows an analysis of the pH dependence of rate constant ratios and in selected instances individual rate constants. We demonstrate large changes in the rate-determining step as well as an unprecedented inversion in the kinetic order of substrate release from ternary complex over an interval of 2 pH units. Previously, fumarate has been used in dopamine beta-monooxygenase assays because of its property of enzyme activation. Studies of the pH behavior in the presence of saturating concentrations of fumarate have shown two causes of the activation: (i) fumarate perturbs the pK of B1 to pK = 6.6-6.8 such that the residue remains protonated and the enzyme optimally active over a wider pH range; (ii) fumarate decreases the rate of dopamine release from the ternary enzyme-substrate complex, increasing the equilibrium association constant for dopamine binding. Both effects are consistent with a simple electrostatic stabilization of bound cationic charges by the dianionic form of fumarate.  相似文献   

14.
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   

15.
Formation and thermodynamic characteristics of C-H ... O hydrogen bonding of methylated uracils and caffeine have been studied by nmr along two lines. 1. The concentration and temperature dependencies of the PMR spectra of 1,3-dimethyluracil (m2 1,3Ura), 1,3-dimethylthymine (m2 1,3Thy), and 1,3,6-trimethyluracil (m3 1,3,6Ura) in chloroform at high concentrations of base analogs indicated the self-association of m2 1,3Ura and m2 1,3Thy via C(6)H ... O hydrogen bonding and the competitive formation of C-H ... O bonds between carbonyl oxygens and chloroform. The intermolecular interaction energy and the arrangement of molecules in the local minima of various m2 1,3Ura dimers were calculated by the method of atom-atom potentials. The deepest minimum for the m2 1,3Ura coplanar dimer corresponds to a C(6)-H ... O hydrogen-bond formation. 2. At low concentration of m2 1,3Ura and caffeine in CCl4, C(6)-H ... O bonding for m2 1,3Ura and C(8)-H ... O bonding for caffeine with oxygens of dimethyl sulfoxide (DMSO) and acetone were observed. The association constants of these complexes were obtained at different temperatures. The enthalpies delta H, of the m2 1,3Ura-DMSO, m2 1,3Ura-accetone, caffeine-DMSO, and caffeine-acetone complexes were -2 +/- 0.1 kcal/mol. The calculations showed that the deepest minimum of the caffeine-acetone coplanar complex corresponds to C(8)-H ... O bonding with energy of -3.5 kcal/mol and that of the m2 1,3Ura-acetone complexes corresponds to C(6)-H ... O bonding with energy of -3.4 kcal/mol. The approximate correction for the solvent effect provides good agreement of the experimental data with the calculations.  相似文献   

16.
The structure and phospholipid-binding properties of human plasma high density apolipoprotein A-I (apoA-I) has been studied at pH 7.4 and 3.1 by microcalorimetry, circular dichroism and density gradient ultracentrifugation. At pH values of 7.4 and 3.1, apoA-I binds to dimyristoyl phosphatidylcholine (DMPC) to form complexes of similar composition (molar ratio of DMPC/apoA-I of 100) and helical content (67%). At pH 7.4, the lipid-protein association is accompanied by an increase in helical content from 58 to 67% and an exothermic enthalpy of binding (deltaHB) of -90 kcal/mol apoA-I. At pH 3.1, the helical content of apoA-I is increased from 48 to 67% on binding to DMPC and the enthalpy of binding was -170 kcal/mol. We suggest that the difference in the enthalpies of binding (-80 kcal/mol) at pH 3.1 compared to 7.4 is due to the greater coil leads to helix transition at the lower pH.  相似文献   

17.
The pH dependence of the maximum velocity (V) for the phosphorylation of glucose, the V/Kglucose and the V/KMgATP have been obtained in H2O and 2H2O. In H2O, V decreases below a pK of 5.8, V/Kglucose decreases below a pK of 6.1 and V/KMgATP decreases below a pK of 6.7. In 2H2O, complex behavior is observed for these parameters as a function of pD. The ratios of the parameters in H2O and 2H2O above their respective pK values give solvent deuterium isotope effects of about 1.5-1.7 for all three parameters. When 1,5-anhydromannitol is used as an alternative substrate, an isotope effect different than unity is obtained only for V/K1,5-anhydromannitol which gives a value of about 0.7. Both the complex pH profiles and the relative magnitude of the isotope effects are interpreted in terms of a pH-dependent change in the E X glucose complex.  相似文献   

18.
The temperature dependence of the oxygen equilibrium of tadpole hemoglobin has been determined between 0 degrees and 32 degrees for the unfractionated but phosphate-free lysate and between 12 degrees and 32 degrees for each of the four isolated components between pH 6 and 10 in 0.05 M cacodylate, Tris, or glycine buffers containing 0.1 M NaCl and 1 mM EDTA. Under these conditions the Bohr effect (defined as deltalog p50/deltapH) of the unfractionated lysate is positive at low temperatures between pH 6 and 8.5 and is negative above pH 8.5 to 8.8 at any temperature. As the temperature rises the Bohr effect below pH 8.5 changes greatly. In the interval pH 7.0 to 7.5, the magnitude of the Bohr effect decreases from + 0.28 at 0 degrees to zero at about 24 degrees and becomes negative, as in mammalian hemoglobins, above this temperature. Measurements with the isolated components show that the temperature dependence of oxygen binding for Components I and II and for Components III and IV is very similar. For both sets of components the apparent overall enthalpy of oxygenation at pH 7.5 is about -16.4 kcal/mol and -12.6 kcal/mol at pH 9.5. The measured enthalpies include contributions from the active Bohr groups, the buffer ions themselves, the hemoglobin groups contributing buffering, and any pH-dependent, oxygenation-dependent binding of ions such as chloride by the hemoglobin. The apportioning of the total enthalpy among these various processes remains to be determined. Between pH 8 and 10.5 tadpole oxyhemoglobin undergoes a pH-dependent dissociation from tetramer to dimer. The pH dependence of the apparent tetramer-dimer dissociation constant indicates that at pH 9.5 the dissociation of each tetramer is accompanied by the release of approximately 2 protons. In this pH range the oxygen equilibrium measurements indicate that about 0.5 proton is released for each oxygen molecule bound. The results are consistent with the conclusion that one acid group per alphabeta dimer changes its pK from about 10 to 8 or below upon dissociation of the tetramer.  相似文献   

19.
(1) In the pH range between 5.0 and 8.0, the rate constants for the reaction of ferrocytochrome c with both the high- and low-affinity sites on cytochrome aa3 increase by a factor of approx. 2 per pH unit. (2) The pre-steady-state reaction between ferrocytochrome c and cytochrome aa3 did not cause a change in the pH of an unbuffered medium. Furthermore, it was found that this reaction and the steady-state reaction are equally fast in H2O and 2H2O. From these results it was concluded that no protons are directly involved in a rate-determining reaction step. (3) Arrhenius plots show that the reaction between ferrocytochrome c and cytochrome aa3 requires a higher enthalpy of activation at temperatures below 20°C (15–16 kcal/mol) as compared to that at higher temperature (9 kcal/mol). We found no effect of ionic strength on the activation enthalpy of the pre-steady-state reaction, nor on that of the steady-state reaction. This suggests that ionic strength does not change the character of these reactions, but merely affects the electrostatic interaction between both cytochromes.  相似文献   

20.
The solvent kinetic isotope effects (SKIE) on the yeast alpha-glucosidase-catalyzed hydrolysis of p-nitrophenyl and methyl-d-glucopyranoside were measured at 25 degrees C. With p-nitrophenyl-D-glucopyranoside (pNPG), the dependence of k(cat)/K(m) on pH (pD) revealed an unusually large (for glycohydrolases) solvent isotope effect on the pL-independent second-order rate constant, (DOD)(k(cat)/K(m)), of 1.9 (+/-0.3). The two pK(a)s characterizing the pH profile were increased in D(2)O. The shift in pK(a2) of 0.6 units is typical of acids of comparable acidity (pK(a)=6.5), but the increase in pK(a1) (=5.7) of 0.1 unit in going from H(2)O to D(2)O is unusually small. The initial velocities show substrate inhibition (K(is)/K(m) approximately 200) with a small solvent isotope effect on the inhibition constant [(DOD)K(is)=1.1 (+/-0.2)]. The solvent equilibrium isotope effects on the K(is) for the competitive inhibitors D-glucose and alpha-methyl D-glucoside are somewhat higher [(DOD)K(i)=1.5 (+/-0.1)]. Methyl glucoside is much less reactive than pNPG, with k(cat) 230 times lower and k(cat)/K(m) 5 x 10(4) times lower. The solvent isotope effect on k(cat) for this substrate [=1.11 (+/-0. 02)] is lower than that for pNPG [=1.67 (+/-0.07)], consistent with more extensive proton transfer in the transition state for the deglucosylation step than for the glucosylation step.  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