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.
1. The kinetics of the heterologous recombination reaction of alkylated H chains of a myeloma protein (Jo) with alkylated L chains of another myeloma protein (Ita) were studied by following changes with time in the circular dichroism at 235 nm and the results were compared with those for the autologous recombination of Jo-H chains with Jo-H chains reported previously (T. Azuma et al.(1975) J. Biochem. 77, 473-479 and the preceding paper). The heterologous reaction also followed second-order kinetics. The second-order rate constant (kapp) for heterologous recombination was about seven times smaller than that for autologous recombination at pH 5.5, while they were similar between pH 4.2 and 4.7. 2. The apparent association constants (Kapp) for the reaction, H2+L2=H2L2, were determined by measuring the ellipticities at 235 nm of mixtures of H and L chains in various ratios. The values of Kapp for the autologous and heterologous recombinations were both pH-dependent and changed from 10(6) M-1 at pH 3.9 to 108 M-1 at pH 4.3. Using these values of kapp and Kapp, the half-time for the dissociation of autologous H2L2 to H2 and L2 at pH 4.3 was estimated to be 80 hr.  相似文献   

2.
The noncovalent interaction of light (L) chain with heavy (H) chain or Fd isolated from a human myeloma protein Jo (IgG1, kappa) was studied by following circular dichroic (CD) change at 235 nm. The dimerization constants of Jo-L chain determined by measuring the CD change at 293 nm with protein concentration showed that the Jo-L chain exists as the monomeric form under the experimental conditions used for recombination with H chain. The second-order rate constants for the interaction between H and L chains were in good agreement with those for the interaction between Fd and L chain at various pH values. The binding behavior of L chain to Fd could be described by a single association constant. In the interpretation of the binding of L chain to H chain, however, it was necessary to assume that the binding of L chain to one of the two sites on H chain dimer (H2) decreases the affinity of the other site for L chain. The binding constant of the first L chain to H2 was the same as that of L chain to Fd. Renaturation processes of L chain, Fd, Fab(SS) fragment (with intact interchain disulfide bond), and Fab(RA) fragment (in which the interchain disulfide bond had been reduced and alkylated) from the denatured states in 0.5 or 1 M acetic acid on neutralization were studied. The renaturation of Fd occurred very rapidly, while that of L chain consisted of a very rapid process and a slow process which followed first-order kinetics. The renaturation process of Fab(SS) consisted of rapid and slow phases, of which the latter followed first-order kinetics. The renaturation process of Fab(RA) also consisted of rapid and slow phases, but the latter process followed second-order kinetics. The overall rate constant of renaturation of Fab(RA) was the same as that of the reformation of Fab(RA) from isolated Fd and L chain. On the basis of these facts, the kinetic mechanism by which Fd and L chain recombine to yield Fab(RA) can be described in terms of the scheme Fd + L in equilibrium Fd ... L leads to Fab(RA), where Fd ... L is an intermediate, and CD change is only observed in the second unimolecular process and not in the first bimolecular process.  相似文献   

3.
S P Rao  W G Miller 《Biopolymers》1973,12(4):835-843
The circular dichroism (CD) spectrum of poly-L-lysine and poly-L -glutamic acid has been investigated in the presence of a small percent of side-chain blocking groups. The blocking groups benzyl, methyl, and carbobenzoxy show qualitatively similar effects. Less than five mole percent of aromatic blocking groups alters the CD spectrum. Consequently, unsuspected blocking groups may account for the variation observed in CD spectra of these polymers. A weak CD band at 235–240 nm was observed for the disordered unblocked polymer even in the absence of electrolyte. Viscosity data indicate that in salt-free solutions these chains at neutral pH still behave as random coils though with reduced conformational freedom, in contrast to some polyelectrolytes which behave as rigid rods in the absence of electrolyte. The viscosity data bring into question the relevance of isolated molecule conformational calculations to experimental CD spectra.  相似文献   

4.
CO recombination to the cloned cytochrome c peroxidase [CCP(MI)] and mutants of CCP(MI) prepared by site-directed mutagenesis was examined as a function of pH by flash photolysis. The mutants examined included distal Arg 48----Leu, Lys; proximal Asp 235----Asn; and His 181----Gly. At alkaline pH, ferrous CCP(MI) was converted to a hexacoordinate form by a cooperative two-proton ionization, apparent pK(a) = 8.0. This change was observed in all of the mutants, although in the His 181----Gly mutant, the conversion to the hexacoordinate form was the result of a single-proton ionization, implicating His 181 as one of the two residues deprotonated in this isomerization. The pH-dependent conversion of CO ferrous CCP(MI) from acidic to alkaline forms was also observed and was similar to that reported for cytochrome c peroxidase from bakers' yeast [Iizuka, T., Makino, R., Ishimura, Y., & Yonetani, T. (1985) J. Biol. Chem. 260, 1407-1412]. Photolysis of the acidic form of the CO complex of CCP(MI) produces a kinetic form of the ferrous enzyme (form A) which exhibits the slow rate of CO recombination (l1' approximately 10(3) M-1 s-1) characteristic of peroxidases, while photolysis of the alkaline form of the CO complex produces a second kinetic form (form B), which exhibits a much faster rate of recombination (l2' approximately 10(5) M-1 s-1). Kinetic forms analogous to forms A and B were observed in all of the mutants examined. A third kinetic form (form B*) with a bimolecular rate constant l3' approximately 10(6) M-1 s-1 was also observed in the mutants at alkaline pH. Although the pH dependence for the conversion of form A to form B with increasing pH was altered by changes in the local heme environment, the rate of CO recombination by the respective forms was not dramatically altered in the mutants. Transient spectra of the reaction of CO with ferrous CCP(MI) after photolysis show that equilibrium between penta- and hexacoordinate ferrous enzyme is rapid relative to CO recombination. The presence of the internal sixth ligand has no discernible effect on the observed rate of recombination, however. The results presented indicate that in CCP(MI) the rate of ligand binding is determined primarily by isomerization of the protein from a closed conformation at acidic pH to an open conformation at alkaline pH and that polar effects of proximal Asp 235 and distal Arg 48 are of minor significance in the rate of CO recombination in both conformations.  相似文献   

5.
Acid, guanidinium-Cl and urea denaturations of recombinant human macrophage migration inhibitory factor (MIF) were measured using CD and fluorimetry. The acid-induced denaturation was followed by CD at 200, 222, and 278 nm and by tryptophan fluorescence. All four probes revealed an acid-denatured state below pH 3 which resembled a typical molten globule. The pH transition is not two-state as the CD data at 222 nm deviated from all other probes. Urea and guanidinium-Cl denaturations (pH 7, 25 degrees C) both gave an apparent DeltaGU app H2O of 31 +/- 3 kJ.mol-1 when extrapolated to zero denaturant concentration. However, denaturation transitions recorded by fluorescence (at the same protein concentration) occurred at lower urea or guanidinium-Cl concentrations, consistent with an intermediate in the course of MIF denaturation. CD at 222 nm was not very sensitive to protein concentration (in 10-fold range) even though size-exclusion chromatogryphy (SEC) revealed a dimer-monomer dissociation prior to MIF unfolding. Refolding experiments were performed starting from acid, guanidinium-Cl and urea-denatured states. The kinetics were multiphasic with at least two folding intermediates. The intrinsic rate constant of the main folding phase was 5.0 +/- 0.5 s-1 (36.6 degrees C, pH 7) and its energy of activation 155 +/- 12 kJ.mol-1.  相似文献   

6.
According to its circular dichroism (CD) spectrum, modeccin, a toxic lectin from the roots of the South African plantModecca digitata, is structurally similar to the ricins and abrins. In nearly neutral and weakly alkaline solutions (pH 7.6–9.0) the CD spectra of modeccin displayed a positive CD band at 190–195 nm and a negative band at 210–220 nm, indicating the presence of some α-helix and β-sheet structures. In the near-ultraviolet zone, we observed positive CD bands at 232 and 245 nm and weak negative bands at 285 and 293 nm. In more strongly alkaline solutions of pH 9.5–10.2 the CD bands in the farultraviolet zone were not affected, but the CD band at 232 nm diminished and the CD band at 245 nm was enhanced. These transitions were reversible. At pH 11.2–11.5 the CD band at 232 nm disappeared completely, and the CD bands in the far-ultraviolet diminished. The CD bands at 285 and 293 nm were affected very little by the alkali, and these bands were assigned to buried tryptophan side chains. Sodium dodecyl sulfate and 2,2,2-trifluoroethanol disorganized the tertiary structure of modeccin and reconstructed the secondary structure into a new form with a higher helix content than in the native protein.  相似文献   

7.
Botulinum neurotoxin (BoNT) is an extremely toxic protein to animals and humans. In its mode of action, one of its subunits mediates its translocation by integrating itself into the membrane bilayer. We have examined the membrane channel activity of type A BoNT (BoNT/A) and its heavy (H) chain in planar lipid membrane under various pH conditions to understand the possible role of the channel activity in the translocation of the BoNT/A light (L) chain under physiological conditions. Only BoNT/A H chain, and not the BoNT/A, exhibited membrane channel activity for translocation of ions. The H chain-induced increase in conductance did not require a pH gradient across the lipid membrane, although it was enhanced by a pH gradient. To understand the molecular basis of the membrane channel activity and the translocation of the L chain, the secondary structure of BoNT/A and its H and L chains were analyzed using circular dichroism (CD) and Fourier-transform infrared (FT-IR) spectroscopy at different pH values. BoNT/A showed no structural alternation upon acidifying the buffer pH. However, an increase in beta-sheet content of BoNT/A H chain at low pH was noted when examined by FT-IR. The L chain structure significantly changed with decrease in pH, and the change was mostly reversible. In addition, the neurotoxin and its subunit chains induced a partially reversible aggregation of liposomes at low pH, which indicated their integration into the lipid bilayer. Temperature-induced denaturation studies of BoNT/A H chain indicated major structural reorganization upon its interaction with membrane, especially at low pH.  相似文献   

8.
Salt and pH-dependent effects on the local conformation of the two tyrosine residues of the histone H1 from the fruit fly Ceratitis capitata have been studied by difference spectroscopy, circular dichroism (CD), fluorescence emission and quenching of tyrosine emission by KI. Four different solvent conditions were used: pH 1.0 and pH 6.0 in the presence or absence of 1 M NaCl. The results clearly indicate a partial folding of the fruit fly histone H1 in the absence of NaCl upon pH increasing from 1.0 to 6.0. This folding is achieved only through ionization of acidic side chains which leads to an environment of the tyrosine residues similar to that found in salt-folded histone H1. Furthermore, the present results give new insights into the states of the tyrosine residue responsible for the tyrosinate-like fluorescence emission at 340 nm described previously (Jordano, J., Barbero, J.L., Montero, F. & Franco, L. (1983) J. Biol Chem. 258, 315-320).  相似文献   

9.
We recently reported results of an investigation of the reoxidation of a human, monoclonal immunoglobulin G, following selective reduction of its interchain disulfides by dithiothreitol (Sears, D.W., et al. (1975), Proc. Natl. Acad. Sci. U.S.A. 72, 353). In that work, we described the reoxidative behavior of the molecule under nondissociating conditions. In the present paper, results are presented of the reoxidation of heavy (H) and light (L) chains of this protein alone, or mixed in varying proportions after separation, or mixed with the L chains modified prior to recombination and reoxidation. The overall reoxidative asembly patterns in experiments with H and L separated prior to recombination are similar to those observed when the chains remain noncovalently associated throughout. With equimolar mixtures of H and L, the reoxidation rates also are similar to those of unseparated chains. However, when L chains are present in excess, the overall in vitro rates of covalent assembly are generally diminished, probably indicating transient nonproductive interactions. At the highest molar excesses of L (3:1), the assembly pathways may also be modified. In all experiments with excess L chains, covalent L2 dimers form at rates which are comparatively slow relative to the H2L2 assembly rates. Two kinds of reoxidation experiments with modified L chains are described here for the first time. In the first, the free half-cystine of L is irreversibly blocked by reaction with iodoacetamide, and the alkylated L chains are recombined with reduced H chains. This experiment isolates the reactions in which H2 disulfides are formed without the accompanying formation of HL bonds. Although the alkylated L chains do not play a direct role in the reoxidation, their presence is required to inhibit aggregation and precipitation of high-molecular-weight products which otherwise ensue; this suggests a possible biological role for excess L in vivo. In the second kind of experiment, covalent L2 dimers are mixed with reduced H chains. L2 rapidly disappears with the concurrent appearance of HL, H2L, and fully assembled H2L. H2 dimers are also reactive in this process. Special procedures were developed for analyzing the data from these experiments. A complete format is given for the quantitative determination of the concentration of each of the molecular components directly from spectroscopic scans of the gels. The computational methods solve the general analytical problem posed when staining is not proportional to mass and are applicable to a wide variety of systems utilizing gel electrophoresis to study subunit interactions. A theoretical analysis of pathway and kinetic cooperatively in this system is presented in the following paper (Sears, D.W., and Beychok, S. (1977), Biochemistry 16 (following paper in this issue)).  相似文献   

10.
J Zikán  O Kotynek 《Biopolymers》1968,6(5):681-690
The interaction of hapten (ε-DNP lys) with native and S-sulfonated antibodies specific towards the 2,4-dinitrophenyl group, as well as the interaction with isolated chains and a complex obtained by mixing light, (L) and heavy (H) chains of these antibodies, were followed both by polarography and by equilibrium dialysis. With the S-sulfonated antibodies and with the mixture of H and L chains the binding heterogeneity observed in the original antibodies was much lowered or entirely removed. At the same time, the amount of active proteins in the sample decreased approximately by half. The association constants of modified antibodies were of the same order as the average association constants of the original antibodies. A slow increase of the amounts of hapten bound with proteins was observed on mixing the H and L chains and adding hapten. This slow reactivation was not obtained with the original or S-sulfonated antibodies and with isolated chains. It was shown that the reaction determining the kinetics of this reactivation (the slowest reaction) was not the association of H and L chains but the interaction of complexes of the H and L chains with hapten. It was reported previously that H chains were nonspecifically reactivated by binding L chains. The amount of hapten bound by the complex of H and L chains increased with increasing excess of L chains following a curve resembling the Langmuir isotherm. The limiting value of the amount of hapten bound when using antibody L chains was higher than in the case of nonspecific L chains.  相似文献   

11.
Disulphide bonds of rabbit γ-G-globulin and the antibody of the γ-G-globulin type against the 2,4-dinitrophenyl group were split both by the oxidative sulphitolysis at pH 8.6 and by the reduction with 2-mercaptoethanol followed by carboxymethylation. The fractionation was carried out in 0.05 m formic acid containing 6m urea, in 1m propionic acid or in 6m guanidine hydrochloride. Both heavy (H) and light) (L) chains are released from the I+J fraction preceding on an elution diagram H chains when rechromatographed in a stronger desaggregation medium. A small amount of the L chains is also released on rechromatography of the H chains (isolated from 1m propionic acid) in 6m guanidine hydrochloride. The separation of the degraded γ-G-globulin in 0.05m formic acid containing 6m urea or in 6m guanidine hydrochloride showed a separation of the L chains to two fractions differing by electrophoretic properties, peptide maps and N-terminal amino acids. However, these chains exhibit a similar molecular weight, immunoelectrophoretic behaviour and similar properties on reactivation of the antibody H chain.  相似文献   

12.
Oxygen electrodes and spectrophotometric analysis have been used to evaluate the contribution of H2O2, in addition to available chlorine, to the high redox potential of electrolyzed anode water (EAW) with potassium chloride as an electrolyte. H2O2 was added externally to EAW, and the reaction between H2O2 and the available chlorine in the water was examined. EAW has a low pH (2.5), a high concentration of dissolved oxygen, and extremely high redox potentials (19 mg/l and 1,319 mV) when the available chlorine is at the concentration of about 580 microM. The addition of H2O2 to EAW led to H2O2 decomposition, and the amount of oxygen produced was equivalent to the amount of available chlorine. Oxygen production was reduced by ascorbic acid, and completely inhibited by 600 microM ascorbate. The rate of oxygen production was much affected by pH, and was slowest at or near pH 5.0. Rates were particularly high in alkaline solution. Absorbance at 235 nm (pH 3.0 and 5.0) and 292 nm (pH 10.0) decreased when H2O2 was added to the EAW at these pHs, and the extent of decrease was similar pH dependency to that of the oxygen production rate. Oxygen was not produced after H2O2 was added to EAW at pH 2.6 when available chlorine was absent, but oxygen was produced after potassium hypochlorite was added to such EAW. The oxygen production rates in EAW without available chlorine at pH 5.0 and 2.0, pH adjustment with KOH and HCl, respectively, were faster than the rate at pH 2.6, and fastest at pH 2.0. These results suggest that H2O2 or hydroxyl radicals derived from Fenton's reaction did not contribute to the high redox potential of EAW prepared with chlorine compounds as an electrolyte, so that the decomposition of H2O2 occurred rapidly with the reactions of chlorine and hypochlorite ions in EAW.  相似文献   

13.
Trifluoroethanol (TFE) has been used to probe differences in the stability of the native state and in the folding pathways of the homologous cysteine protein inhibitors, human stefin A and B. After complete unfolding in 4.5 mol/L GuHCl, stefin A refolded in 11% (vol/vol) TFE, 0.75 mol/L GuHCl, at pH 6.0 and 20 degrees C, with almost identical first-order rate constants of 4.1 s-1 and 5.5 s-1 for acquisition of the CD signal at 230 and 280 nm, respectively, rates that were markedly greater than the value of 0.11 s-1 observed by the same two probes when TFE was absent. The acceleration of the rates of refolding, monitored by tyrosine fluorescence, was maximal at 10% (vol/vol) TFE. Similar rates of refolding (6.2s-1 and 7.2 s-1 for ellipticity at 230 and 280 nm, respectively) were observed for stefin A denatured in 66% (vol/vol) TFE, pH 3.3, when refolding to the same final conditions. After complete unfolding in 3.45 mol/L GuHCl, stefin B refolded in 7% (vol/vol) TFE, 0.57 mol/L GuHCl, at pH 6.0 and 20 degrees C, with a rate constant for the change in ellipticity at 280 nm of 32.8 s-1; this rate was only twice that observed when TFE was absent. As a major point of distinction from stefin A, the refolding of stefin B in the presence of TFE showed an overshoot in the ellipticity at 230 nm to a value 10% greater than that in the native protein; this signal relaxed slowly (0.01 s-1) to the final native value, with little concomitant change in the near-ultraviolet CD signal; the majority of this changes in two faster phases. After denaturation in 42% (vol/vol) TFE, pH 3.3, the kinetics of refolding to the same final conditions exhibited the same rate-limiting step (0.01 s-1) but were faster initially. The results show that similarly to stefin A, stefin B forms its hydrophobic core and predominant part of the tertiary structure faster in the presence of TFE. The results imply that the alpha-helical intermediate of stefin B is highly structured. Proteins 1999;36:205-216.  相似文献   

14.
R S Lord  D J Cox 《Biopolymers》1973,12(10):2359-2373
Oligopeptides containing glycine and one or two L -alanyl or L -glutamyl residues have been studied by circular dichroism (CD) and optical rotatory dispersion (ORD) in aqueous solution at pH 1.0, pH 6.0, and pH 10.0 and in aqueous ethanol. Two glycyl residues are required to remove effects of α-carboxyl or amino titration on the optical activity of the internal alanyl or glutamyl residues. The CD spectra of the alanyl and protonated glutamyl residues are similar, having two regions of negative ellipticity around 215 nm resulting in a spectrum reassembling that of poly-α-L -glutamic acid (PGA) at high pH. Another large positive band below 190 nm was observed for gly2-glu2-gly2 in water at pH 6 and 10 and for several peptides in aqueous ethanol. Residue ellipticities were approximately additive in every case except for peptides containing intrenal glutamyl residu at pH 6.0.  相似文献   

15.
We study by small-angle neutron scattering (SANS) the structure of hyaluronan -lysozyme complexes. Hyaluronan (HA) is a polysaccharide of 9 nm intrinsic persistence length that bears one negative charge per disaccharide monomer (M(mol) = 401.3 g/mol); two molecular weights, M(w) = 6000 and 500,000 Da were used. The pH was adjusted at 4.7 and 7.4 so that lysozyme has a global charge of +10 and +8, respectively. The lysozyme concentration was varied from 3 to 40 g/L at constant HA concentration (10 g/L). At low protein concentration, samples are monophasic, and SANS experiments reveal only fluctuations of concentration, although, at high protein concentration, clusters are observed by SANS in the dense phase of the diphasic samples. In between, close to the onset of the phase separation, a distinct original scattering is observed. It is characteristic of a rod-like shape, which could characterize "single" complexes involving one or a few polymer chains. For the large molecular weight (500,000), the rodlike rigid domains extend to much larger length scale than the persistence length of the HA chain alone in solution and the range of the SANS investigation. They can be described as a necklace of proteins attached along a backbone of diameter of one or a few HA chains. For the short chains (M(w) ≈ 6000), the rod length of the complexes is close to the chain contour length (~ 15 nm).  相似文献   

16.
CD15 mRNA和nm23H1 mRNA表达与结直肠癌转移和预后的关系   总被引:1,自引:0,他引:1  
为了探讨评估结直肠癌预后指标。应用催化信号放大-原位杂交技术。研究90例结直肠癌组织中CD15和nm23H1的mRNA表达,并结合随访资料分析,结果表明,在结直肠癌中CD15和nm23H1的mRNA阳性表达分别为84.4%和66.7%。CD15mRNA高表达及nm23H1mRNA低表达与结直肠癌浸润程度,淋巴结转移和肝脏转移状况密切相关,提示CD15和nm23H1的mRNA表达均可作为预测结直肠癌侵袭转移和客观评估患者预后的生物学指标。  相似文献   

17.
Parallel stranded DNA with AT base pairing   总被引:1,自引:0,他引:1  
The concentration and temperature dependences of the UV and CD spectra of the oligonucleotide 3'-d(ApTpApTpApTpApTpApTp)-O(CH2)6O-5'-d(pApTpApTpApTpApT pApT) (eicosamer) in aqueous solution at pH 7 in the presence of 0.5 M NaCl were studied. At less than 10(-6) M, the eicosamer was shown to form in solution a hairpin with parallel orientation of chains (parallel hairpin). From thermal denaturation profiles [A260(T)] the thermodynamic parameters, delta H degrees, delta S degrees and Tm for parallel hairpin formation were calculated to be -90 +/- 8 kJ/mol. -300 +/- 20 J.mol-1.K-1 and 40.5 degrees C, respectively. The CD spectra of the parallel double helix differed from those of B-form DNA and had characteristic features: decreasing magnitude of the positive maximum at 265 nm and a negative peak at 285 nm.  相似文献   

18.
Bispecific mAb contain two Ag-combining sites each composed of a different combination of H and L chains. The resulting ability to react with and cross-link two different Ag makes these molecules a novel tool for application in biology and medicine. Intact bispecific mAb can be made only by biologic means, e.g., by fusion of two established hybridomas. Appropriate assembly of bispecific mAb by these hybrid cells depends on H = L chain behavior: strong preferential homologous H-L pairing would benefit the yield of bispecific antibodies. We have analyzed the Ig species produced by eight hybrid hybridomas (quadromas). Quadroma-produced IgG was fractionated and characterized for H and L chain content. The Ag reactivities were verified by using ELISA and immunofluorescence. Preferential homologous pairing was seen only with a minority of H-L chain pairs; L chains associated on average in a random fashion with H chains. This indicates that in the B cells from which the parental hybridomas were obtained, no strong selection had occurred on H-L recombination. Our results extend recent biochemical competitive H-L reassociation experiments, where on average an at random association of L chains with H chains was found; evidently this random association occurs in our biologic system as well. For the biologic production of bispecific antibodies this means that only in a small number of cases the "ideal" producer will be met. From the viewpoint of generation of antibody diversity, our results favor a large freedom for combinatorial binding of H and L chains during B cell ontogeny.  相似文献   

19.
Both cytochrome c peroxidase (CcP) and a mutant cytochrome c peroxidase in which the distal histidine has been replaced by leucine, CcP(H52L), are converted to hydroxy-ligated derivatives at alkaline pH. In CcP, the hydroxy-ligated derivative is subsequently converted to a bis-imidazole species prior to protein denaturation while the initial hydroxy-ligated CcP(H52L) is converted to a second, spectroscopically distinct hydroxy-ligated species prior to denaturation. The spectra of the alkaline forms of CcP and CcP(H52L) have been determined between 310 and 700 nm. The pH dependence of the rate of reaction between CcP(H52L) and hydrogen peroxide has been extended to pH 10. The hydroxy-ligated form of CcP(H52L) reacts with hydrogen peroxide 4 times more rapidly than the pentacoordinate, high-spin form of CcP(H52L) that exists at neutral pH. The rate of the reaction between p-nitroperoxybenzoic acid and CcP(H52L) has been measured between pH 4 and pH 8. Neutral p-nitroperoxybenzoic acid reacts with CcP(H52L) 10(5) times more slowly than with CcP while the negatively charged p-nitroperoxybenzoate reacts with CcP(H52L) 10(3) times more slowly than with CcP. These data indicate that the role of the distal histidine during the initial formation of the peroxy anion/heme iron complex is not simply base catalysis.  相似文献   

20.
The gross protein structure, the microenvironment of the iron-sulfur cluster, and the effect of neutral salts on the molecular structure of spinach ferredoxin were studied by CD and absorption spectroscopy in the alkaline pH range. In the pH range of 7-11, the existence of reversible isomerization which consisted of at least two proton dissociation processes was indicated by the statical CD and absorption spectra. The CD changes in the visible and far-UV regions were dramatic upon elevation of the pH from neutral to alkaline, indicating a significant alteration of the microenvironment of the cluster and a decrease in the ordered secondary structures. The absorption change in the visible region due to pH elevation was small but clearly observed with a high signal-to-noise ratio. The numbers of protons involved in the respective processes and the apparent pK values obtained from the pH-dependence of the CD changes were in good agreement with those obtained from the pH-dependence of the absorption changes in the visible region. In addition, the rate constants obtained from the time courses of the CD and absorption changes agreed with one another. By the addition of 1 M NaCl, the CD and absorption spectra at alkaline pH were reversed almost to those at neutral pH without significant pH change. On the other hand, above pH 11, ferredoxin was found to be irreversibly denatured. Based on analyses of the statical CD and absorption spectra and of the time courses of the CD changes, the probable mechanism of the isomerization was considered to be as follows: (Formula: see text) where H stands for a proton, N-form for native ferredoxin at neutral pH, N*-form for alkaline ferredoxin below pH 11 which still has the iron-sulfur cluster but with disordered secondary structures of the polypeptide chain, and D-form for completely denatured ferredoxin above pH 11. These results lead to the conclusions that (1) the interaction between the protein moiety and the iron-sulfur cluster is essential for maintaining the native ferredoxin structure, and (2) neutral salts protect the polypeptide chain from unfolding through electrostatic interaction with the ionized side chains, resulting in the stabilization of ferredoxin.  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