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 25 毫秒
1.
The second order rate constants for the oxidation of high potential iron-sulfur protein (Hipip) of Chromatium vinosum by ferricyanide were determined as a function of ionic strength and pH. From the ionic strength results, calculations were done to correct the rate constant at each pH for the electrostatic interactions between Hipip and ferricyanide. The electrostatic corrections are necessary since the charge of the protein changes as a function of pH and can mask the ionization of mechanistically important amino acid residues. An apparent pKa 7 was obtained from electrostatically corrected rate-pH profile, indicating the possible participation of histidine. Perturbation difference spectroscopic studies of Hipip as a function of pH also gave apparent pKa values of 6.9 and 6.7 for the reduced and oxidized protein, respectively. That it was indeed His 42 (the only His in the polypeptide) that was responsible for the kinetic and spectroscopic pKa values was demonstrated by modification of His 42 of Hipip by the histidine selective reagent diethylpyrocarbonate. No modification of Tyr 19 could be detected. It is concluded that either deprotonation or modification of His 42 results in the destabilization of the reduced cluster and thus a faster rate of oxidation. This work provides the first experimental evidence of the ‘squeeze effect’ mechanism (Carter, C.W., Jr., Kraut, J., Freer, S.T. and Alden, R.A. (1974) J. Biol. Chem. 249, 6339–6346) in which the polypeptide directly modulates the stability of the iron-sulfur cluster.  相似文献   

2.
Due to contradictions in the literature we have redetermined the acid-base properties of riboflavin (=RiFl; vitamin B2), i.e. 7,8-dimethyl-10-ribityl-isoalloxazine, and of flavin mononucleotide (FMN2−), also known as riboflavin 5′-phosphate, via potentiometric pH titrations (I = 0.1 M, NaNO3; 25 °C). In contrast to various claims, the isoalloxazine ring cannot be protonated at pH > 1, a result in agreement with an early study (pKa = −0.2; L. Michaelis, M.P. Schubert and C.V. Smythe, J. Biol. Chem., 116 (1936) 587–607); deprotonation of the ring system occurs in both compounds with pKa 10. The pKa value of 0.7 determined for the deprotonation of H2(FMN) must be attributed to the release of the first proton from the fully protonated phosphate group; its second proton is released with pKa = 6.18 in agreement with the acidity constants of various other monoprotonated monophosphate esters. The stability constants of the 1:1 complexes formed between Mg2+, Ca2+, Sr2+, Ba2+, Mn2+, Co2+, Ni2+, Cu2+, Zn2+ or Cd2+ (---M2+) and FMN2− were determined by potentiometric pH titrations in aqueous solution (I = 0.1 M, NaNO3; 25 °C). The log stability constants of all these M(FMN) complexes are about 0.2 log units higher than expected from the basicity of the phosphate group. This slight stability increase cannot be attributed to the formation of a seven-membered chelate involving the ribit-hydroxy group at C-4′ as the stability constants for the M2+ 1:1 complexes of glycerol 1-phosphate (G1P2−) demonstrate: G1P2− contains the same structural unit which would also allow in this case the formation of the mentioned seven-membered chelate; however, the stability of the M(G1P) complexes is solely determined by the basicity of the phosphate group. Hence, in agreement with earlier conclusions (J. Bidwell, J. Thomas and J. Stuehr, J. Am. Chem. Soc., 108 (1986) 820–825) regarding Ni(FMN) one must conclude that the slight stability increase of the M(FMN) complexes has to be attributed to the isoalloxazine ring. The equality of the stability increase of the complexes for all the mentioned ten metal ions precludes its attribution to an interaction with an N site and makes a specific interaction with an O site also somewhat unlikely. In addition, carbonyl oxygens appear as not very favorable for the formation of macrochelates by a further interaction with already phosphate-coordinated metal ions. Therefore, we propose that the slight but significant stability increase originates from M(FMN) species (with a formation degree of about 30%) in which the hydrophobic flavin residue is close to the metal ion, thereby lowering the ‘effective’ dielectric constant in the microenvironment of the metal ion and thus indirectly promoting the −PO32−/M2+ interaction.  相似文献   

3.
The one-electron reduction potential of 3-amino-l, 2, 4-benzotriazine 1, 4-dioxide, tirapazamine (SR 4233) in aqueous solution has been determined by pulse radiol-ysis. Reversible electron transfer was achieved between radiolytically-generated one-electron reduced radicals of tirapazamine (T), and quinones or benzyl viologen as redox standards. The reduction potential Em7(T/T±) was -0.45 ± 0.01 V vs. NHE at pH 7. From the pH dependence of the reduction potential, pKa = 5.6 ± 0.2 was estimated for the tirapazamine radical, a value similar to the pKa determined by other methods.  相似文献   

4.
The catalytic activity of papain in the synthesis of Z-Gly-Phe-NH2 in tert-butanol has been studied in the presence of solid-state acid–base buffers (acids and their sodium salts). All buffer pairs tested reduced the reaction rate compared with the control, particularly the most acidic and basic (assessed by either aqueous pKa or the response of an organic phase indicator). The highest rates, close to the control, were found with glutamic acid/glutamate-Na, PIPES/PIPES-Na and NaH2PO4/Na2HPO4. However, these pairs were unable to erase the pH memory phenomenon, or to overcome the effect of spiking with acetic acid. Hence, at least these buffers do not seem to be able to affect the protonation state and catalytic activity of papain. In the last aqueous solution before drying, the presence of activating agents (cysteine plus EDTA) was more important than buffer ions.  相似文献   

5.
The relationship between the pKa of 8-quinolinol derivatives {8-quinolinol (Hqn), 2-methyl- (H2-Meqn), 2,4-dimethyl- (H2,4-diMeqn), 5-chloro- (H5-Clqn) and 5,7-dichloro-8-quinolinols (H5,7-diClqn)} and a π-donor ability of the 8-quinolinolato oxygens has been investigated by the identification of the structures of the major products, [RuCl(QN)(QN′)NO] (HQN=8-quinolinol derivative; HQN′=different 8-quinolinol derivatives), obtained by the reaction of [RuCl3(QN or QN′)NO] with HQN′ or HQN. The results obtained clearly showed that the oxygen of the 8-quinolinol derivative that has a higher pKa predominantly coordinates in the trans position to the NO ligand and is a better π-electron donor. The order of the π-electron donor ability for the oxygen of the 8-quinolinol derivatives is as follows: H2-Meqn≥H2,4-diMeqn>Hqn≥H5-Clqn>H5,7-diClqn, almost agreeing with the magnitude of the pKa values of the corresponding 8-quinolinols. The structures of cis-1 [RuCl(5,7-diClqn)2NO] and cis-1 [RuCl(5,7-diClqn)(2-Meqn)NO] were determined by X-ray diffraction.  相似文献   

6.
Ab initio (B3LYP) calculations show that PD·H---ReH4(PH3)3 (PD = Proton donor) interactions follow the order PD = pyrrole > NH3 > HCCH > C2H4 > CH3---H 0 and decrease with the pKa of the PD. For equivalent pKa's, NH interacts more strongly than CH. However, intermolecular hydrogen-bonding of the M---H·H---C type is too weak to be detected experimentally in FTIR or UV-vis studies between ReH5(PPh3)3 and PhCCH, C6F5H or PhCHCl2.  相似文献   

7.
Electrochemical analysis of lignin peroxidase (LiP) was performed using a pyrolytic graphite electrode coated with peroxidase-embedded tributylmethyl phosphonium chloride membrane. The formal redox potential of ferric/ferrous couples of LiP was −126 mV (versus SHE), which was comparable with that of manganese peroxidase (MnP) and horseradish peroxidase (HRP). Yet, only LiP is capable of oxidizing non-phenolic substrates with a high redox potential. Since with decreasing pH, the redox potential increased, an incredibly low pH optimum of LiP as peroxidase at 3.0 or lower was proposed as the clue to explain LiP mechanisms. A low pH might be the key for LiP to possess a high redox potential. The pKa values for the distal His in peroxidases were calculated using redox data and the Nernst equation, to be 5.8 for LiP, 4.7 for MnP, and 3.8 for HRP. A high pKa value of the distal His might be crucial for LiP compound II to uptake a proton from the solvent. As a result, LiP is able to complete its catalytic cycle during the oxidation of non-proton-donating substrates. In compensation, LiP has diminished its reactivity toward hydrogen peroxide.  相似文献   

8.
Guar gum/poly(acrylic acid) semi-interpenetrating polymer network (IPN) hydrogels have been prepared via free radical polymerization in the presence of a crosslinker of N,N′-methylene bisacrylamide (MBA). The kinetics of swelling and the water transport mechanism were studied as a function of the composition of the hydrogels and the pH of the swelling medium. Hydrogels showed enormous swelling in aqueous medium and displayed swelling characteristics, which were highly dependent on the chemical composition of the hydrogels and pH of the medium in which hydrogels were immersed (ionic strength I = 0.15 mol/L). The semi-INP hydrogels were characterized by evaluating various network parameters such as average molecular weight between crosslinks (Mc) crosslink density (ρ) and mesh size ξ.  相似文献   

9.
The crystal structure and magnetic properties of a dinuclear copper(II) complex of the ligand [2,8-dimethyl-5,11-di-(dimethylethyleneamine) 1,4,5,6,7,10,11,12-octahydroimidazo [4,5-h] imidazo [4,5-c] [1,6]diazecine] dimeim have been investigated. Also, its catecholase activity has been explored in different solvent mixtures: MeCN/H2O and OH/H2O, each at several pH values. In CH3OH/H2O, where the activity was superior, the optimal pH value for the catalytic activity was found to be lower than in CH3CN/H2O. The study of the complex’s electrochemical behavior (cyclic voltammetry) which was also investigated in these various media, revealed that although an increase in pH in both solvent mixtures results in an increase both in Me oxidizing power (E1/2) and reversibility (ipa/ipc) the change of solvent system seems to be a more influencing factor. The superior catalytic activity found in MeOH/H2OpH=8.0, is associated with a significantly more reversible behavior displayed in this medium. Potentiometric determination of the overall formation constant and three successive pKas for the complex, suggest the formation of stable hydroxo complexes which could be the catalytically active species.  相似文献   

10.
The oxidation of TEMPO (2,2,6,6-tetramethyl-piperidine-1-oxyl radical) has been studied in the presence of recombinant laccases (benzenediol:oxygen oxidoreductase, EC 1.10.3.2) from Polyporus pinsitus (rPpL), Myceliophthora thermophila (rMtL), Coprinus cinereus (rCcL) and Rhizoctonia solani (rRsL) in buffer solution pH 4.5–7.3 and at 25 °C. At pH 5.5 the oxidation constant calculated from the initial rate of TEMPO oxidation was 1.7 × 104, 1.4 × 103, 7.8 × 102 and 5.2 × 102 M−1 s−1 for rPpL, rRsL, rCcL and rMtL, respectively. The maximal activity of rPpL-catalysed TEMPO oxidation was at pH 5.0. The pKa obtained in neutral pH range was 6.2. The reactivity of laccases is in a good agreement with laccases copper type I redox potential.

TEMPO oxidation rate increased 541 times in the presence of 10-(3-propylsulfonate) phenoxazine (PSPX). The model of synergistic TEMPO and PSPX oxidation was proposed. Experimentally obtained rate constants for rPpL-catalysed PSPX oxidation were in a good agreement with those calculated from the synergistic model, therefore confirming the feasibility of the model. The acceleration of TEMPO oxidation with high reactive laccase substrates opens new possibilities for TEMPO application as a mediator.  相似文献   


11.
A series of dihydroxamic acid ligands of the formula [RN(OH)C(O)]2(CH2)n, (n = 2, 4, 6, 7, 8; R = CH3, H) has been studied in 2.0 M aqueous sodium perchlorate at 25.0 °C. These ligands may be considered as synthetic analogs to the siderophore rhodotorulic acid. Acid dissociation constants (pKa) have been determined for the ligands and for N-methylacetohydroxamic acid (NMHA). The pKa1 and pKa2 values are: n = 2, R = CH3 (8.72, 9.37); N = 4, R = CH3 (8.79, 9.37); N = 6, R = CH3; N = 7, R = CH3 (8.95, 9.47); N = 8, R = CH3 (8.93, 9.45); N = 8, R = H (9.05, 9.58). Equilibrium constants for the hydrolysis of coordinated water (log K) have been estimated for the 1:1 feeric complexes of the ligands n = 2, 4, 8; R = CH3. The N = 8 ligand forms a monomeric complex with Fe(III) while the n = 2 and 4 ligands form dimeric complexes. For hydrolysis of the n = 8 monomeric complex, log K1 = −6.36 and log K2 = −9.84. Analysis of the spectrophotometric data for the dimeric complexes indicates deprotonation of all four coordinated waters. The successive hydrolysis constants, log K1–4, for the dimeric complexes are as follows: n = 2 (−6.37, −5.77, −10.73, −11.8); n = 4 (−5.54, −5.07, −11.57, −10.17). The log K2 values for the dimers are unexpectedly high, higher in fact than log K1, inconsistent with the formation of simple ternary hydroxo complexes. A scheme is proposed for the hydrolysis of the ferric dihydroxamate dimers, which includes the possible formation of μ-hydroxo and μ-oxo bridges.  相似文献   

12.
Acid dissociation constants of 2,3-diphytanyl-sn-glycero-1-phosphoryl-sn-3′-glycero-1′-methylphosphate (PGP-Me), the major phospholipid in extreme halophiles (Halobacteriaceae), and of the demethylated 2,3-diphytanyl-sn-glycero-1-phosphoryl-sn-3′-glycero-1′-phosphate (PGP) and its deoxy analog 2,3-diphytanyl-sn-glycero-1-phosphoryl-1′-(1′,3′-propanediol-3′-phosphate) (dPGP), were calculated by an original mathematical procedure from potentiometric titration data in methanol/water (1:1, v/v) and found to be as follows: for PGP-Me (and presumably PGP), pK1=3.00 and pK2=3.61; for PGP, pK3=11.12; and for dPGP, pK1=2.72, pK2=4.09, and pK3=8.43. High-resolution 31P NMR spectra of intact PGP-Me in benzene/methanol or in aqueous dispersion showed two resonances corresponding to the two P-OH groups, each of which exhibited a chemical shift change in the pH range 2.0–4.5, corresponding to acid dissociation constants pK1=pK2=3.2; no further ionization (pK3) was detected at higher pH values in the range 5–12. The present results show that PGP-Me titrates as a dibasic acid in the pH range 2–8, but above pH 8, it titrates as a tribasic acid, presumably PGP, formed by hydrolysis of the methyl group during the titration with KOH. Calculation of the concentrations of the ionic molecular species of PGP-Me, PGP and dPGP as a function of pH showed that the dianionic species predominate in the pH range 5–9, covering the optimal pH for growth of Halobacteriaceae. The results are consistent with the concept that the hydroxyl group of the central glycerol moiety in PGP-Me and PGP is involved in the formation of an intramolecular hydrogen-bonded cyclic structure of the polar headgroup, which imparts greater stability to the dianionic form of PGP-Me and PGP in the pH range 5–9 and facilitates lateral proton conduction by a process of diffusion along the membrane surface of halobacterial cells.  相似文献   

13.
A series of novel tetrahydroimidazo[2,1-a]isoquinolines was prepared based on a hetero Diels–Alder reaction between an enamine and 1,2,4-triazine as key step. A structure–activity relationship was established focussing on the influence of the substitution pattern in position 3 and 6 of the heterocycle on antisecretory activity, lipophilicity, and pKa value. Potent inhibitors of the gastric acid pump were identified.  相似文献   

14.
Isothermal titration calorimetry (ITC) was used to investigate thermodynamic parameters of the cyclosporin A (CsA)-cyclophilin 18 (hCyp18) association reaction. We have calculated the thermodynamic parameters (enthalpy, entropy, heat capacity, and free energy of binding) of the CsA/hCyp18 complexation. All but two methods described in the literature underestimate the affinity to hCyp18 of CsA. We found that the association constant (1.1·108 M−1 at 10 °C) of CsA to hCyp18 is in close agreement with the reciprocal of the reported inhibitory constant of the peptidylprolyl cis/trans isomerase activity of hCyp18. Interpretation of the thermodynamic parameters in buffered solution of water, 30% glycerol and D2O leads to the conclusion that the highly specific binding of CsA to hCyp18 is mainly mediated through hydrogen bonding and to a lesser degree through hydrophobic interaction. Furthermore, the pH dependence of the association constant was determined and analyzed according to a single proton linkage model, resulting in a pKa value of 5.7 in free hCyp18 and below 4.5 in the CsA complexed form. Titration experiments using different single component buffers possessing different heats of ionization allowed us to estimate that statistically half a proton is transferred upon CsA binding from the binding interface of hCyp18 to the buffer at pH 5.5. No proton transfer was detected at pH 7.5. The thermodynamic results are discussed in relation to the published X-ray and NMR structure of the free and CsA complexed hCyp18.  相似文献   

15.
The stability constants of the 1:1 complexes formed between Mg2+, Ca2+, Sr2+, Ba2+, Mn2+, Co2+, Ni2+, Cu2+, Zn2+ or Cd2+ (M2+) and the simple, sterically unhindered imidazole-type ligands, imidazole, 1-methylimidazole, 5-chloro-1-methylimidazole, N-(2,3,5,6-tetrafluorophenyl)imidazole or 4′-(imidazol-1-yl)acetophenone (L), were determined by potentiometric pH titrations in aqueous solution (25°C; I = 0.5 M, NaNO3). The construction of log KMLM versus pKHLH plots results in straight lines; the equations for the least-squares lines are calculated and listed. These data allow calculation of the expected stability constant for a complex of any imidazole-type ligand, provided its pKHLH value (in the pKa range 4–8) is known. For the stabilities of Fe2+ complexes with imidazole-type ligands an estimation procedure is provided. It is shown further that the complex formation between 1-methylbenzimidazole (MBI) and Mn2+, Ni2+, Cu2+ or Zn2+ is s sterically hindered, i.e. the data points for these M(MBI)2+ complexes do not fall on the straight lines defined by the imidazole-type ligands.  相似文献   

16.
Considering the electrostatic potential of active site, four mutants of thermolysin (EC 3.4.24.4) are designed in an attempt to change the optimum pH of the hydrolytic activity toward acidic regions. On the basis of the numerical calculation of the electrostatic potential in the thermolysin molecule, Asp213 is targeted to be replaced by a basic residue, His, Lys, Arg or a neutral one, Asn. The mutant enzymes are produced inBacillus subtilis as a host using the method of site-directed mutagenesis and their optimum pH values for hydrolyzing a synthetic substrate furylacryloyl-Gly-l-Leu-NH2 are found to be lowered by 0.2–0.4 pH units with reference to the wild type enzyme. The pl shifts of the mutants are evaluated. Neither optimum pH nor pl shift can be explained by the contribution of the pK change only at the mutation site. We find a clear negative correlation between the activities at pH 7.0 and the pI values among the four mutants and wild-type enzyme. It suggests that the contribution of pK shift of other residues must be taken into account in order to explain the activity change. Little change of thermal stability is observed among the mutants and wild type enzymes.  相似文献   

17.
The stability constants of the 1:1 complexes formed between Cu(Arm)2+, where Arm = 2,2′-bipyridyl or 1,10-phenanthroline, and methyl phosphate, CH3OPO32−, or hydrogen phosphate, HOPO32−, were determined by potentiometric pH titration in aqueous solution (25°C; l = 0.1 M, NaNO3). On the basis of previously established log K versus pKa straight-line plots (D. Chen et al., J. Chem. Soc., Dalton Trans. (1993) 1537–1546) for the complexes of simple phosphate monoesters and phosphonate derivatives, R-PO32−, where R is a non-coordinating residue, it is shown that the stabilities of the Cu(Arm) (CH3OPO3) complexes are solely determined by the basicity of the -PO32− residue. In contrast, the Cu(Arm) (HOPO3) complexes are slightly more stable (on average by 0.15 log unit) than expected on the basicity of HPO42−; this is possibly due to a more effective solvation including hydrogen bonding, an interaction not possible with coordinated CH3OPO32− species. Regarding biological systems the observation that HOPO32− is somewhat favored over R-PO32− species in metal ion interactions is meaningful.  相似文献   

18.
A series of 3,6-substituted 2,5-bis(1-aziridinyl)-1,4-benzoquinone derivatives was shown to alkylate calf thymus DNA and to form DNA interstrand cross-links. Alkylation and cross-link formation were enhanced after electrochemical reduction of the compounds and increased with lower pH in the pH range from 4.5 to 8.0. Reduction especially shifts the pH at which cross-linking and alkylation occurs to higher values, which are more physiologically relevant. This shift is probably caused by the increase in pKa value of the aziridine ring after reduction of the quinone moiety. The inactivation of single-stranded bacteriophage M13mp19 DNA to form phages in an E. coli host, by the 3,6-unsubstituted parent compound 2,5-bis(1-aziridinyl)-1,4-benzoquinone (TW13) was dependent upon reduction and pH in a similar way as was alkylation. The compound in our series with the least bulky, 3,6-substitutents, TW13, caused a high amount of cross-link formation. Compounds with methyl-substituted aziridine rings showed low cross-linking ability. Our results support the concept that the protonated reduced compound is the reactive species that alkylates DNA, and that steric factors play an important role in the reactivity towards DNA. A correlation is observed between the ability to induce DNA interstrand cross-links and inactivation of M13mp19 bacteriophage DNA. Cross-link formation was also demonstrated in E. coli K12 cells, where the compounds are reduced endogenously by bacterial reductases.  相似文献   

19.
繁殖是植物生命活动的重要环节, 了解植物的繁殖特征是解释植物生态适应性和制定有效管理措施的重要依据。该研究以荒漠草原猪毛蒿(Artemisia scoparia)种群为研究对象, 通过测定不同土壤类型的理化性质和猪毛蒿的繁殖特征, 以期探讨影响其繁殖特征的主要土壤驱动因子。结果表明: 灰钙土、风沙土和基岩风化残积土的水溶性碳含量、全氮含量、全磷含量、全盐含量、土壤水分含量、土壤硬度存在显著差异。猪毛蒿平均个体大小、单株头状花序的平均质量和数量均以灰钙土生境下最大, 基岩风化残积土最小。繁殖分配在不同土壤类型下无显著差异, 但与单个头状花序质量、单株头状花序数量和质量间呈极显著正相关关系。单株头状花序数量与单个头状花序质量间呈负相关关系。在风沙土生境下, 单株头状花序质量主要受到土壤水溶性碳含量土壤水分含量以及pH值的共同影响; 单株头状花序数量受全盐含量的影响最大; 繁殖分配和单个头状花序质量主要受全碳含量的影响。灰钙土生境下, 单株头状花序质量与土壤水溶性碳含量土壤水分含量和有机碳含量呈正相关关系; 速效氮含量显著影响着单株头状花序数量。而基岩风化残积土生境下, 繁殖特征的变异主要受到土壤水溶性碳含量土壤硬度土壤水分含量全磷和速效磷含量的影响。综合分析发现, 土壤因子对猪毛蒿繁殖特征的影响程度不同, 其中单株头状花序数量和质量极显著地受到土壤水溶性碳含量和土壤水分含量的影响, 繁殖分配和单个头状花序质量与土壤水溶性碳含量、土壤水分含量呈负相关关系。因此, 土壤水溶性碳含量和土壤水分是荒漠草原地区影响猪毛蒿种群繁殖特征的主要土壤因子。  相似文献   

20.
Pseudomonas aeruginosa PD100 capable of producing an extracellular protease was isolated from the soil collected from local area (garbage site) from Shivage market in Pune, India. The purified protease showed a single band on native and SDS-PAGE with a molecular weight of 36 kDa on SDS-PAGE. The optimum pH value and temperature range were found to be 8 and 55–60 °C, respectively. The enzyme exhibited broad range of substrate specificity with higher activity for collagen. The enzyme was inhibited with low concentration of Ag2+, Ni2+, and Cu2+. β-Mercaptoethanol was able to inactivate the enzyme at 2.5 mM, suggesting that disulfide bond(s) play a critical role in the enzyme activity. Studies with inhibitors showed that different classes of protease inhibitors, known to inhibit specific proteases, could not inhibit the activity of this protease. Amino acid modification studies data and pKa values showed that Cys, His and Trp were involved in the protease activity. P. aeruginosa PD100 produces one form of protease with some different properties as compared to other reported proteases from P. aeruginosa strains. With respect to properties of the purified protease such as pH optimum, temperature stability with capability to degrade different proteins, high stability in the presences of detergents and chemicals, and metal ions independency, suggesting that it has great potential for different applications.  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