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 31 毫秒
1.
Potentiometric titration curves of oxidized and reduced horse heart cytochrome c in 0.15M KCl at 20°C have been obtained by timed titration (0.125–0.500 μmol/sec) from the isoionic points (pH 10.2–10.4) to pH 3 and back to the isoionic point. Computer-assisted (PROPHET) data acquisition and blank corrections give curves with good precision with a maximum standard deviation of 0.3 groups for an average error of 1%. The potentiometric titration curve of reduced cytochrome c is reversible within the precision of the method and for the pH range studied. The potentiometric curves for oxidized cytochrome c titrated upscale (pH 3–10) and downscale (pH 10–3) are not reversible. However, they show the same ionization behavior after the initial downscale titration. This is probably the result of a conformational change. Comparison of the data herein reported with the titration curves of oxidized cytochrome c already published by others indicates good agreement on the basis of a normalization of the concentration of protein or on the basis of 25 titrable groups between the acid end point and the isoionic pH. Titration of the 2 μmol imidazole in the upscale or downscale direction gives the correct analytical concentration and pK′ after correction for the solvent titration. Titration of reduced cytochrome c in the presence and absence of an additional equivalent of imidazole gave a difference titration curve, which indicates that a group on the protein shifts from pK′ 5.8 to pK′ 5.3 in the presence of imidazole. The pK′ of imidazole, in the presence of the protein, remains at a nearly normal value of 7.34.  相似文献   

2.
Simultaneous curve fitting for the ionization parameters of oxidized and reduced horse heart cytochrome c in 0.15M KCl and 20°C yields values for the ionization constants (as pK′) and the heats of ionization (ΔHi) which can reconstruct either the potentiometric or thermal titration curves. Reduced cytochrome c requires 8 sets of groups, whereas oxidized cytochrome c requires 10 sets of groups. The additional groups in the oxidized preparation appear to involve the ferriheme (pK′, 9.25; ΔHi, 13.7 kcal/mol) and a tyrosine (pK′ ? 10.24) that is not present in the reduced form. The potentiometric and thermal difference curves (reduced – oxidized) involve the appearance of 17 kcal/mol centered at pH 9.7 and 5.8 kcal/mol centered at pH 4.9. The carboxyl groups in both species appear to be normal for the hydrogen-bonded form. Only one histidine has normal ionization properties (pK′, 6.7; ΔHi, 7.5 kcal/mol), as do 17 of the lysine residues (pK′, 10.8; ΔHi, 11.5 kcal/mol).  相似文献   

3.
Chymotrypsinogen, nitrated chymotrypsinogen (two of the four tyrosyls nitrated), acetylated chymotrypsinogen (all amino groups blocked), and nitrated-acetylated chymotrypsinogen were titrated as f(pH) in an isoperibolic calorimeter at 20°C. After appropriate correction and reduction of both the potentiometric and thermal titration data, the parameters N (ionizable groups per group-set), pK′, and ΔHi (heat of ionization) were evaluated using the iterative curve-fitting algorithm of the MLAB computer program. The pK′ parameters so obtained for the two normally ionizing tyrosyl groups in chymotrypsinogen and the two nitrated tyrosyl groups in the nitrated proteins essentially agreed with the results of spectral titration. Excellent fits to all data could be obtained using evaluated parameter sets of N and pK′ for the potentiometric titration data (groups vs pH plots) and N, pK′, and ΔHi sets for the calorimetric data (total heat vs pH plots). The invocation of electrostatic interaction effects was not required to explain the data satisfactorily, despite the differences in charge number and type among the four proteins. Rather, the data can be represented by series expressions of the mass-action law. Using all information, viz., the consequences of functional group modification, the downscale shift in tyrosyl group pK's on nitration, and the numerical values of the evaluated N, pK′, and ΔHi parameter sets for all proteins, the chemical identity of the various classes of group sets can be assigned with reasonable assurance.  相似文献   

4.
J E Ruark  J B Ifft 《Biopolymers》1975,14(6):1161-1171
The buoyant density and potentiometric hydrogen ion titration curves of human immuno-gamma globulin in 3M cesium chloride have been recorded. In addition, the amino acid analysis of the IgG employed has been completed. The hydration of the protein and the variation of the hydration with pH have been calculated from the buoyant density data. The potentiomtric hydrogen ion titration curve has been employed to estimate the intrinsic pK′s of the acidic and histidyl residues of the molecule, and to confirm the hypothesis that it does in fact conform to the oil drop model of protein conformation. Correlations have been drawn between the three sets of data in the following manner. The results of the potentiometric hydrogen ion titration have been checked against the amino acid analysis to determine whether the numbers of groups observed to titrate and the numbers of groups observed in the amino acid analysis do correspond. Second, previous hypotheses as to the direct correlation between potentiometric hydrogen ion titration behaviour and buoyant density titration behaviour have been investigated and substantially confirmed.  相似文献   

5.
The ionization constants of the tyrosyl groups of chymotrypsinogen and of nitrated-chymotrypsinogen (two tyrosyl residues nitrated) have been determined by difference spectrophotometry. In chymotrypsinogen, two of the four tyrosyl groups ionize without any time dependence. Above pH greater than ca. 12.5, time-dependent spectral changes are seen for 0.7 group equivalent. The data can be fitted to the values of pK1 9.75 ± 0.07, pK2 11.55 ± 0.05, pK3 13.30 ± 0.05. In nitrated-chymotrypsinogen, the two nitrated tyrosyl residues have pK1 6.44 and pK2 8.30. For both proteins, these pK′ values are in agreement with those evaluated from potentiometric titration and calorimetric data using computer-assisted curve-fitting analysis.  相似文献   

6.
V. Saudek 《Biopolymers》1981,20(8):1625-1633
Polypeptides of dicarboxylic amino acids having the monomer units linked in α- and ω-peptide bonds contain two kinds of carboxyls of different acidity. How well potentiometric titration can distinguish these two carboxyls and so characterize the nature of the peptide bonds is evaluated critically. An analysis of the equation describing the dependence of pH on the degree of neutralization based on neglecting the polymer effect and a discussion of the dissociation behavior of polyanions show that the method of evaluating experimental data found in the literature is incorrect. Nevertheless, if a conformational transition does not interfere, some useful and reliable information may be gained by this method; namely, an indication of the presence of two different peptide bonds, their mole ratio, and an approximate pK value for the carboxyl of the amino acid linked in the ω-peptide bond. The presence of two types of carboxyls complicates the evaluation of the titration curves in the conformation studies.  相似文献   

7.
The interaction between poly (I) and poly (C) in acid medium has been studied by potentiometric titration, mixing curves and thermal denaturation. Phase diagramms as a function of ionic strength, pH, and temperature have been established. From these data it is shown that the acid titration of the complex poly (I) · poly (C) passes through a triple-stranded intermediate poly (I) · poly (C) · poly (C+) to yield finally the protonated double-helical complex poly (I) · poly (C+). The mixing curves indicate the sole presence of the three-stranded complex in the intermediate zone. On the basis of the pK's the coexistence between the three-stranded complex with the neighboring double-stranded structure is demonstrated in a narrow rang of pH and ionic strength. The geometry of the base arrangements, their conformation and the sense of the strands are discussed in the light of the data presented. A Hoogsteen-type pairing between the bases for poly (I) · poly (C+) is favored, although the reverse Hoogsteen pair cannot be excluded.  相似文献   

8.
The equilibrium oxidation-reduction mipoint potential (Em) of isolated Rhodopseudomonas sphaeroides cytochrome c2 exhibits a pH-dependent behavior which can be ascribed to a pK on the oxidized form at pH 8.0 (Pettigrew et al. (1975) Biochim. Biophys. Acta 430, 197–208). However, as with mammalian cytochrome c (Brandt, K.G., Parks, P.C., Czerlinski, G.H. and Hess, G.P. (1966) J. Biol. Chem. 241, 4180–4185) this pK can more properly be attributed to the combination of a pK beyond pH 11, and a slow conformational change of the ferricytochrome. This has been demonstrated by resolving the Em of cytochrome c2 before and after the conformational change. The Em of the unaltered form is essentially pH independent between pH 7 and 11.5, and the lower equilibrium Em is due solely to the conformational change. In vivo the conformational change is prevented by the binding of the cytochrome c2 to the photochemical reaction center, and the cytochrome exhibits an essentially pH-independent Em from pH 5 to 11. The alkaline transition thus has little physiological significance, and it is unlikely that the redox reactions of cytochrome c2 in vivo involve protons.  相似文献   

9.
NMR‐monitored pH titration experiments are routinely used to measure site‐specific protein pKa values. Accurate experimental pKa values are essential in dissecting enzyme catalysis, in studying the pH‐dependence of protein stability and ligand binding, in benchmarking pKa prediction algorithms, and ultimately in understanding electrostatic effects in proteins. However, due to the complex ways in which pH‐dependent electrostatic and structural changes manifest themselves in NMR spectra, reported apparent pKa values are often dependent on the way that NMR pH‐titration curves are analyzed. It is therefore important to retain the raw NMR spectroscopic data to allow for documentation and possible re‐interpretation. We have constructed a database of primary NMR pH‐titration data, which is accessible via a web interface. Here, we report statistics of the database contents and analyze the data with a global perspective to provide guidelines on best practice for fitting NMR titration curves. Titration_DB is available at http://enzyme.ucd.ie/Titration_DB . Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

10.
M A Marini  C J Martin  L Forlani 《Biopolymers》1981,20(10):2243-2252
Curve fitting for the ionization constants found in the potentiometric titration curve of reduced horse heart cytochrome c in 0.15M KCl at 20°C yields values which can precisely reconstruct the experimental curve. The parameters were evaluated assuming that there were 13 group sets with individual ionization constants, that these groups were subject to electrostatic interaction (ω), and that both pK′ and (ω) were necessary to describe the curve. On the basis of an analysis of the residuals, both the simple sum of mass-law expressions and the electrostatic model could be invoked, although the evaluated ω was both positive and negative. When the standard errors of the parameter values are considered, only the model which assumed no electrostatic interaction is acceptable. The experimental curve is completely described by eight group sets and no electrostatic interaction.  相似文献   

11.
In order to clarify the effect of degree of substitution of ionic and hydrophobic group on the polyelectrolytic behavior of polysaccharides, potentiometric titration and activity measurement of counterions were made for carboxymethyldextran (Cm-dextran) having various degrees of substituted carboxyl group and for carboxymethylbenzyldextran (Cm-Bzl-dextran) containing various degrees of substituted benzyl group. From the shape of titration curve, no conformational change was observed for both Cm-dextran and Cm-Bzl-dextran. The pK0 value of Cm-dextran was independent of the degree of the degree of substituted carboxy group. However, the pK0 of Cm-Bzl-dextran increased with an increasing degree of substituted benzyl group. The suppression of dissociation of a carboxyl group, caused by the surrounding hydrophobic groups, was discussed mainly in terms of the change of water structure around such groups. From the results of activity measurement for counterions of these dextran derivatives, we proposed the possibility of ion selectivity based on the hydrophobicity.  相似文献   

12.
R W Shaw  C R Hartzell 《Biochemistry》1976,15(9):1909-1914
Continuous hydrogen ion titration curves of deionized solutions of horse heart ferricytochrome c have been obtained at 25 degrees C. at a constant ionic strength of 0.10 from pH 3.0 to 11.0. Titration of the oxidized protein in KCl required 28.4 equiv over that pH range, and a small hysteresis between the forward and reverse limbs was displayed. The Linderstrom-Lang approximation, which takes into account electrostatic interactions between charged groups on the protein surface, was used in a computer simulation program to analyze the forward and reverse limbs of the titration curve separately. The results indicated 1 alpha-, 12 beta- and gamma-, and 1 heme propionic carboxylic, 1 imidazole, 1 phenolic, and 18 epsilon-amino residues appear to titrate normally. Variations in the electrostatic interaction factor omega suggest conformational changes in the protein at the extremes of pH, although the relationship of the variations in omega to the magnitude of the conformational changes does not appear to be strictly quantitative for cytochrome c. These results show the acid-base behavior of cytochrome c to be complex in nature, and suggest that the Lindenstrom-Lang model may not be adequate for cytochrome c.  相似文献   

13.
The ability of apocytochrome c and the heme containing respiratory chain component, cytochrome c, to induce fusion of phosphatidylcholine (PC) small unilamellar vesicles containing 0–50 mol% negatively charged lipids was examined. Both molecules mediated fusion of phosphatidylserine (PS):PC 1:1 vesicles as measured by energy transfer changes between fluorescent lipid probes in a concentration- and pH-dependent manner, although cytochrome c was less potent and interacted over a more limited pH range than the apocytochrome c. Maximal fusion occurred at pH 3, far below the pKa of the 19 lysine groups contained in the protein (pl = 10.5). A similar pH dependence was observed for vesicles containing 50 mol% cardiolipin (CL), phosphatidylglycerol (PG), and phosphatidylinositol (PI) in PC but the apparent pKa values varied somewhat. In the absence of vesicles, the secondary structure of apocytochrome c was unchanged over this pH range, but in the presence of negatively charged vesicles, the polypeptide underwent a marked conformational change from random coil to α-helix. By comparing the pH dependencies of fusion induced by poly-L-lysine and apocytochrome c, we concluded that the pH dependence derived from changes in the net charge on both the vesicles and apocytochrome c. Aggregation could occur under conditions where fusion was imperceptible. Fusion increased with increasing mole ratio of PS. Apocytochrome c did induce some fusion of vesicles composed only of PC with a maximum effect at pH 4. Biosynthesis of cytochrome c involves translocation of apocytochrome c from the cytosol across the outer mitochondrial membrane to the outer mitochondrial space where the heme group is attached. The ability of apocytochrome c to induce fusion of both PS-containing and PC-only vesicles may reflect characteristics of protein/membrane interaction that pertain to its biological translocation.  相似文献   

14.
A polarographic method based on the Brdi?ka current (the polarographic catalytic hydrogen evolution current produced by proteins in the presence of cobalt salts) was applied to direct titration of subtilisin BPN′ (S.BPIST) with plasminostreptin (PS) at a concentration level of 10”8 m. The first and second dissociation constants of the S.BPN-PS complex were determined by fitting theoretical curves to the titration data, in which the multiple equilibria involving microscopically distinct forms of S.BPN-PS complex were taken into account. The intrinsic free energy change in the first binding of S.BPN′ to dimeric PS was larger than that in the second binding. The dependence of the microscopic dissociation constants of S.BPN′-PS complex on pH indicates the participation of ionizable groups of pKa 8.0 and 9.4 in the complex formation. The repulsive effect between negatively charged molecules of S.BPN′ and PS in the complex formation at elevated pH is suggested.  相似文献   

15.
The C2H resonance of the active site histidine residue designated AS-2, which has the lower pKa of the two active site histidines, has been correlated in both RNase A and RNase S by comparing the pH 3 to 5.5 regions of the chemical shift titration curves, the effect of the inhibitor CMP-3′ on the chemical shifts at pH 4.0, and the effect of Cu II on the line widths at pH 3.6. It has been demonstrated that resonance AS-2 is absent in the spectrum of RNase S′ reconstituted using S-peptide deuterated at the C2 of His 12, and in that of the RNase S′-CMP-3′ complex. We thus demonstrate that histidine AS-2 is in fact His 12 in both enzymes. This finding is in agreement with out previous assignment of the exchangeable NH proton in RNase A to His 12, but reverses the assignments of the active site histidine C2H resonances made earlier by other authors.  相似文献   

16.
1H-NMR spectroscopy has been used to study the modifications of certain characteristic resonances of the Hansenula anomala yeast cytochrome c on binding to its specific reductase (flavocytochrome b2) or to the isolated cytochrome domain obtained from the entire molecule. Normal titration curves are observed for the resonances at 37.8 ppm assigned to heme c methyl 8 and at 19.4 ppm, line of cytochrome b2 spectrum. In contrast, the shifts near 3.2 and 3.4 ppm for trimethyl-lysine resonances of this cytochrome c present abnormal titration curves, saturation being apparently reached at low molar (cytochrome b2)/(cytochrome c) ratio. An interpretation is proposed in terms of shifts due to local conformational transitions induced by reductase binding but not rapidly reversible upon dissociation.  相似文献   

17.
The potentiometric acid-base titration curve of fully protonated lysozyme at ionic strengths of 0.10 and 1.0 m has been performed. The stoichiometry and the pKa values of each titratable group have been determined through the linearization of titration curves. Two types of carboxylic groups with pKa values of 3.76 and 5.02, the imidazole group with pKa 7.37 and the amine group with pKa 9.63, have been identified at an ionic strength of 0.10 m at 25.0°C. The number of titratable groups found per mole of protein has been 5.12 and 5.60 for the two types of carboxylic groups, 1.13 for the imidazole group, and 3.19 for the amino groups. The endpoint of the titration of the protein obtained by this method accords quite well with the endpoint obtained by the use of Gran function applied to the excess of strong base.  相似文献   

18.
The cytochrome c maturation process is carried out in the bacterial periplasm, where some specialized thiol‐disulfide oxidoreductases work in close synergy for the correct reduction of oxidized apocytochrome before covalent heme attachment. We present a structural and functional characterization of the soluble periplasmic domain of CcmG from the opportunistic pathogen P. aeruginosa (Pa‐CcmG), a component of the protein machinery involved in cyt c maturation in gram‐negative bacteria. X‐ray crystallography reveals that Pa‐CcmG is a TRX‐like protein; high‐resolution crystal structures show that the oxidized and the reduced forms of the enzyme are identical except for the active‐site disulfide. The standard redox potential was calculated to be E0′ = ?0.213 V at pH 7.0; the pKa of the active site thiols were pKa = 6.13 ± 0.05 for the N‐terminal Cys74 and pKa = 10.5 ± 0.17 for the C‐terminal Cys77. Experiments were carried out to characterize and isolate the mixed disulfide complex between Pa‐CcmG and Pa‐CcmH (the other redox active component of System I in P. aeruginosa). Our data indicate that the target disulfide of this TRX‐like protein is not the intramolecular disulfide of oxidized Pa‐CcmH, but the intermolecular disulfide formed between Cys28 of Pa‐CcmH and DTNB used for the in vitro experiments. This observation suggests that, in vivo, the physiological substrate of Pa‐CcmG may be the mixed‐disulfide complex between Pa‐CcmH and apo‐cyt. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

19.
Peaks corresponding to the C(2)-protons of all four histidine residues of porcine β-trypsin were resolved in 250 MHz nuclear magnetic resonance spectra after deuteration of the slowly exchangeable N-H groups (whose resonances obscure the histidine peaks) by reversible unfolding of the protein in D2O. One of the four peaks was assigned to the charge-relay histidine in the active site of trypsin (His(57) in the bovine chymotrypsinogen numbering system). Whereas the three other histidine C(2)-peaks exhibited normal titration curves with single pK′ values of 7.20, 6.71 and 6.67, the peak assigned to His(57) had an abnormal titration curve showing two protonation steps in the pH range from 1 to 9. The first protonation with a pH′mid of 5.0 is rapid on the nuclear magnetic resonance time-scale; the second with a pH′mid of 4.5 is slow and apparently involves conformational transitions between two states having lifetimes of approximately 18 ms.In the complex between porcine β-trypsin and bovine pancreatic trypsin inhibitor (Kunitz) His(57) was found to be insensitive to pH over the range from 4 to 9 and its chemical shift resembles that of His(57) in the singly protonated charge relay of free trypsin. This result provides direct evidence that the trypsin charge relay acts as a proton acceptor in the initial catalytic step which leads to the formation of a tetrahedral complex. In the presence of equimolar bovine pancreatic trypsin inhibitor (Kunitz) the pH'mid of the conformational transition that affects the charge-relay histidine is lowered from 4.5 to approximately 3.5.  相似文献   

20.
Masaru Tanokura 《Biopolymers》1983,22(12):2563-2576
The 1H-nmr titration curves of chemical shifts versus pH were observed for the protons of D,L-histidyl-D,L-histidine as representative of cases with two or more ionizable groups with similar pKa values. The titration curves of L -histidyl-L -histidine and D -histidyl-L -histidine were individually analyzed according to two mathematical models: one of a macroscopic dissociation series and one of a microscopic dissociation series. Most-probable values and standard deviations were obtained for pKa values and intrinsic chemical shifts. An analysis including the microscopic dissociation series yielded an electrostatic interaction between twoimidazole rings of about 0.3 pH units for L -histidyl-L -histidine and about 0.7 pH units for D -histidyl-L -histidine. The difference of the magnitude of imidazole-imidazole interactions between L -histidyl-L -histidyne and D -histidyl-L -histidine was interpreted in terms of the spatial arrangement of two imidazole rings in each molecule based on the solution conformation estimated from Gd(III)-induced relaxation enhancements.  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