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.
The Met16Phe mutant of the type 1 copper protein pseudoazurin (PACu), in which a phenyl ring is introduced close to the imidazole moiety of the His81 ligand, has been characterized. NMR studies indicate that the introduced phenyl ring is parallel to the imidazole group of His81. The mutation has a subtle effect on the position of the two S(Cys)-->Cu(II) ligand-to-metal charge transfer bands in the visible spectrum of PACu(II) and a more significant influence on their intensities resulting in a A(459)/A(598) ratio of 0.31 for Met16Phe as compared to a A(453)/A(594) ratio of 0.43 for wild-type PACu(II) at pH 8. The electron paramagnetic resonance spectrum of the Met16Phe variant is more axial than that of the wild-type protein, and the resonance Raman spectrum of the mutant exhibits subtle differences. A C(gamma)H proton of Met86 exhibits a much smaller hyperfine shift in the paramagnetic (1)H NMR spectrum of Met16Phe PACu(II) as compared to its position in the wild-type protein, which indicates a weaker axial Cu-S(Met86) interaction in the mutant. The Met16Phe mutation results in an approximately 60 mV increase in the reduction potential of PACu. The pK(a) value of the ligand His81 decreases from 4.9 in wild-type PACu(I) to 4.5 in Met16Phe PACu(I) indicating that the pi-pi contact with Phe16 stabilizes the Cu-N(His81) interaction. The Met16Phe variant of PACu has a self-exchange rate constant at pH 7.6 (25 degrees C) of 9.8 x 10(3) M(-)(1) s(-)(1) as compared to the considerably smaller value of 3.7 x 10(3) M(-)(1) s(-)(1) for the wild-type protein under identical conditions. The enhanced electron transfer reactivity of Met16Phe PACu is a consequence of a lower reorganization energy due to additional active site rigidity caused by the pi-pi interaction between His81 and the introduced phenyl ring.  相似文献   

2.
The effect of pH on the (1)H NMR spectrum, reduction potential, and self-exchange rate constant of the novel plastocyanin (PCu) from the fern plant Dryopteris crassirhizoma has been studied. The results are compared with those for the higher-plant PCu from parsley. In the (1)H NMR spectrum of D. crassirhizoma PCu(I), there is no sign that either of the His ligands is protonated at pH* down to 5.4. The reduction potentials of D. crassirhizoma and parsley PCu are 382 and 379 mV, respectively, at pH 7.4. When the pH value is decreased, the reduction potential of parsley PCu is seen to increase quite dramatically, consistent with protonation at His87 in PCu(I). A pK(a) of 5.8 is obtained from the electrochemistry data, consistent with a value of 5.6 determined by NMR. The reduction potential of D. crassirhizoma PCu exhibits a much less pronounced dependence on pH. The self-exchange rate constant of D. crassirhizomaPCu(I) is 3.4 x 10(3) M(-1) s(-1) at pH* 7.9. This is the smallest self-exchange rate constant reported to date for a PCu and can be rationalized by considering the altered distribution of charged residues on the surface of the D. crassirhizoma protein compared to the charge distributions of other higher-plant PCus. The self-exchange rate constant increases to 9 x 10(3) M(-1) s(-1) at pH* 5.4, consistent with enhanced protein-protein association at lower pH*, and the absence of His87 protonation in D. crassirhizoma PCu(I) in the accessible pH range.  相似文献   

3.
Ultrafast, laser-induced pH jump with time-resolved photoacoustic detection has been used to investigate the early protonation steps leading to the formation of the compact acid intermediate (I) of apomyoglobin (ApoMb). When ApoMb is in its native state (N) at pH 7.0, rapid acidification induced by a laser pulse leads to two parallel protonation processes. One reaction can be attributed to the binding of protons to the imidazole rings of His24 and His119. Reaction with imidazole leads to an unusually large contraction of -82 +/- 3 ml/mol, an enthalpy change of 8 +/- 1 kcal/mol, and an apparent bimolecular rate constant of (0.77 +/- 0.03) x 10(10) M(-1) s(-1). Our experiments evidence a rate-limiting step for this process at high ApoMb concentrations, characterized by a value of (0. 60 +/- 0.07) x 10(6) s(-1). The second protonation reaction at pH 7. 0 can be attributed to neutralization of carboxylate groups and is accompanied by an apparent expansion of 3.4 +/- 0.2 ml/mol, occurring with an apparent bimolecular rate constant of (1.25 +/- 0.02) x 10(11) M(-1) s(-1), and a reaction enthalpy of about 2 kcal/mol. The activation energy for the processes associated with the protonation of His24 and His119 is 16.2 +/- 0.9 kcal/mol, whereas that for the neutralization of carboxylates is 9.2 +/- 0.9 kcal/mol. At pH 4.5 ApoMb is in a partially unfolded state (I) and rapid acidification experiments evidence only the process assigned to carboxylate protonation. The unusually large contraction and the high energetic barrier observed at pH 7.0 for the protonation of the His residues suggests that the formation of the compact acid intermediate involves a rate-limiting step after protonation.  相似文献   

4.
The proton nuclear magnetic resonance spectrum of azurin from Alcaligenes denitrificans at pH 6.0 and 309 K is reported. Proton signals from all methionine and histidine residues (among them the copper ligands) have been assigned. The data have been used to study the pH behaviour of His35 and to establish the electron self-exchange rate of the protein. His35 appears to be protonated at pH less than 4.5, possibly after rupture of a salt bridge. No effects of this protonation on the tertiary structure around the copper site are observed, however, contrary to the case of Pseudomonas aeruginosa azurin. The electron self-exchange rate amounts to 4 x 10(5) M-1 S-1 at pH 6.7 and 297 K. The data support the conclusion that the electron self-exchange takes place by way of the hydrophobic surface patch around His117, and that His35 is not involved in this reaction. Oxidation of azurin increases the acidity of the freely titrating His32 and His83 by 0.07 and 0.25 pKa units, respectively. The data can be used to test the theory of electrostatic interactions in proteins. The optical extinction coefficient at 625 nm was experimentally determined and amounts to 4.8(+/- 0.1) x 10(3) M-1 cm-1.  相似文献   

5.
Hass MA  Christensen HE  Zhang J  Led JJ 《Biochemistry》2007,46(50):14619-14628
Exchange on the microsecond time scale between the protonated and deprotonated forms of His92 in the copper site of reduced plastocyanin from the cyanobacteria Anabaena variabilis was monitored using 15N NMR relaxation measurements. On the basis of the dependence of the kinetics on pH and phosphate buffer concentration, we propose a two-step model for the protonation of the copper site in agreement with previous crystallographic studies. It is shown that the proton transfer is the rate-limiting step in the reaction at low buffer concentrations, whereas at high buffer concentrations, another step becomes rate-limiting. We suggest that the latter step is a concerted dissociation of His92 from the Cu(I) ion and a 180 degrees rotation of the imidazole ring, which precede the protonation. The first-order rate constant for the dissociation of His92 from the Cu(I) ion is estimated to be 2.4 x 10(4) s(-1). Also, a cooperative effect of the protonation of the remote His61 on the protonation of His92 and the redox properties of the protein was investigated by substituting His61 with asparagine. The mutation causes a modest change in both the pKa value of His92 and the redox potential of the protein.  相似文献   

6.
The mutant replacing the Met86 ligand of Achromobacter cycloclastes pseudoazurin (Ac-pAz) with Gln has been prepared and spectroscopically and electrochemically characterized. Ac-pAz has four ligands (2His, Cys, and Met) and donates one electron to its cognate Cu-containing nitrite reductase (Ac-NIR). The mutant ([Met86Gln]pAz) shows the electronic absorption and CD spectra considerably similar to those of zucchini mavicyanin (Mv) and lacquer and cucumber stellacyanins (St) having 2His, Cys, and Gln. The EPR signal of the mutant has an axial character, although those of Mv and St show rhombic signals. The findings indicate that the Cu site having Gln might be a distorted trigonal geometry. The half-wave potentials (E(1/2)) of [Met86Gln]pAz and the intermolecular electron-transfer rate constant (kET) from the mutant to Ac-NIR were determined by cyclic voltammetry at pH 7.0 and 25 degrees C. The E(1/2) is +134 mV (versus NHE) and the coordination of Gln instead of Met negatively shifts the E(1/2) of Ac-pAz (+260 mV (versus NHE)). The kET of [Met86Gln]pAz (1.2x10(6) M(-1) s(-1)) is larger than that of the recombinant Ac-pAz (7.5x10(5) M(-1) s(-1)).  相似文献   

7.
Overexpression of hepatocyte growth factor (HGF) and its receptor Met often occurs in carcinoma cells, leading to establishment of an HGF/Met autocrine loop. Therefore, disruption of the HGF/Met autocrine loop may lead to down-regulation of tumorigenesis. To study the HGF/Met interaction, we have developed a cell-free system to detect HGF binding to a Met fusion protein, Met-IgG, using a modified enzyme-linked immunosorbent assay methodology. Since we previously showed that HGF can be purified by copper(II) affinity chromatography, we further explored the effect of copper(II) on the HGF/Met interaction. The divalent metal cations copper(II) and zinc(II) significantly inhibited HGF binding to immobilized Met-IgG with IC(50) values of 230-270 microM, respectively, whereas manganese(II) and magnesium(II) were less inhibitory with 20-60-fold higher IC(50) values. Incubation of 1 mM copper(II) with HGF resulted in nondenaturing and denaturing gel-mobility shifts, indicating that copper(II) binds directly to HGF. This interaction occurs at the N terminus of HGF, as incubation of 1 mM copper(II) with both HGF and the HGF derivative NK1 yielded similar results on SDS-PAGE. HGF-induced activation of Met and cell scattering were inhibited upon addition of HGF in the presence of 1 mM and 500 microM copper(II), respectively. Chemical protonation with diethyl pyrocarbonate of HGF histidine residues impeded the ability of 500 microM copper(II) to inhibit the binding of HGF to immobilized Met-IgG. Based on the NK1 domain structure, we propose that copper(II) may interact with HGF via the histidine residues in either N-terminal or kringle domains. The inhibition of HGF/Met interaction and subsequent downstream cellular functions may be through direct interference by copper(II), such as a change in charge or an induced local conformational change. This putative copper(II) binding domain may be the basis for developing potential inhibitors of HGF/Met binding and downstream functions and could lead to novel strategies for anti-cancer treatment.  相似文献   

8.
The mutant (M150Q-NIR) replacing the Met150 ligand of the type 1 Cu center in Achromobacter cycloclastes nitrite reductase (AcNIR) with Gln has been physicochemically and functionally characterized. The electronic absorption and CD spectra of M150Q-NIR are similar to those of mavicyanin and stellacyanin having the 2His, Cys, and Gln ligands, but the EPR signal has an axial character, although their blue copper proteins show rhombic EPR signals. The mutant has about 80% catalytic activity of AcNIR. Moreover, the midpoint potential (E(1/2)) of M150Q-NIR is +113 mV vs. NHE at pH 7.0, being negatively shifted compared to that of AcNIR (+240 mV). Although the intermolecular electron-transfer process from Achromobacter cycloclastes pseudoazurin (pAz) to M150Q-NIR was not detected, the pAz mutant (M86Q-pAz) replacing the Met86 ligand with Gln transfers one electron to the NIR mutant with an intermolecular electron-transfer rate constant (k(ET)) of 2.3 x 10(5)M(-1)s(-1).  相似文献   

9.
The catalytic properties of organophosphate hydrolase (OPH) containing a hexahistidine tag His6 (His6-OPH) and purified to 98% homogeneity were investigated. The pH optimum of enzymatic activity and isoelectric point of His6-OPH, which were shown to be 10.5 and 8.5, respectively, are shifted to the alkaline range as compared to the same parameters of the native OPH. The recombinant enzyme possessed improved catalytic activity towards S-containing substrates: the catalytic efficiency of methylparathion hydrolysis by His6-OPH is 4.2 x 10(6) M(-1) x sec(-1), whereas by native OPH it is 3.5 x 10(5) M(-1) x sec(-1).  相似文献   

10.
Ma QF  Hu J  Wu WH  Liu HD  Du JT  Fu Y  Wu YW  Lei P  Zhao YF  Li YM 《Biopolymers》2006,83(1):20-31
Amyloid-beta peptide (Abeta) is the principal constituent of plaques associated with Alzheimer's disease (AD) and is thought to be responsible for the neurotoxicity associated with the disease. Copper binding to Abeta has been hypothesized to play an important role in the neruotoxicity of Abeta and free radical damage, and Cu2+ chelators represent a possible therapy for AD. However, many properties of copper binding to Abeta have not been elucidated clearly, and the location of copper binding sites on Abeta is also in controversy. Here we have used a range of spectroscopic techniques to characterize the coordination of Cu2+ to Abeta(1-16) in solution. Electrospray ionization mass spectrometry shows that copper binds to Abeta(1-16) at pH 6.0 and 7.0. The mode of copper binding is highly pH dependent. Circular dichroism results indicate that copper chelation causes a structural transition of Abeta(1-16). UV-visible absorption spectra suggest that three nitrogen donor ligands and one oxygen donor ligand (3N1O) in Abeta(1-16) may form a type II square-planar coordination geometry with Cu2+. By means of fluorescence spectroscopy, competition studies with glycine and L-histidine show that copper binds to Abeta(1-16) with an affinity of Ka approximately 10(7) M(-1) at pH 7.8. Besides His6, His13, and His14, Tyr10 is also involved in the coordination of Abeta(1-16) with Cu2+, which is supported by 1H NMR and UV-visible absorption spectra. Evidence for the link between Cu2+ and AD is growing, and this work has made a significant contribution to understanding the mode of copper binding to Abeta(1-16) in solution.  相似文献   

11.
The oxidation of serotonin (5-hydroxytryptamine) by the myeloperoxidase intermediates compounds I and II was investigated by using transient-state spectral and kinetic measurements at 25.0 +/- 0.1 degrees C. Rapid scan spectra demonstrated that both compound I and compound II oxidize serotonin via one-electron processes. Rate constants for these reactions were determined using both sequential-mixing and single-mixing stopped-flow techniques. The second order rate constant obtained for the one-electron reduction of compound I to compound II by serotonin is (1.7 +/- 0.1) x 10(7) M(-1) x s(-1), and that for compound II reduction to native enzyme is (1.4 +/- 0.1) x 10(6) M(-1) x s(-1) at pH 7.0. The maximum pH of the compound I reaction with serotonin occurs in the pH range 7.0-7.5. At neutral pH, the rate constant for myeloperoxidase compound I reacting with serotonin is an order of magnitude larger than for its reaction with chloride, (2.2 +/- 0.2) x 10(6) M(-1) x s(-1). A direct competition of serotonin with chloride for myeloperoxidase compound I oxidation was observed. Our results suggest that serotonin may have a role to protect lipoproteins from oxidation and to prevent enzymes from inactivation caused by the potent oxidants HOCl and active oxygen species.  相似文献   

12.
Effect of a hydrophobic peptide on folding of oxidized cytochrome c (cyt c) is studied with trityrosine. Folding of cyt c was initiated by pH jump from 2.3 (acid-unfolded) to 4.2 (folded). The Soret band of the 2-ms transient absorption spectrum during folding decreased its intensity and red-shifted from 397 to 400 nm by interaction with trityrosine, whereas tyrosinol caused no significant effect. The change in the transient absorption spectrum by interaction with trityrosine was similar to that obtained with 100 mM imidazole, which showed that the population of the intermediate His/His coordinated species increased during folding of cyt c by interaction with trityrosine. The absorption change was biphasic, the fast phase (82+/-9s(-1)) corresponding to the transition from the His/H(2)O to the His/Met coordinated species, whereas the slow phase (24+/-3s(-1)) from His/His to His/Met. By addition of trityrosine, the relative ratio of the slow phase increased, due to increase of the His/His species at the initial stage of folding. According to the resonance Raman spectra of cyt c, the high-spin 6-coordinate and low-spin 6-coordinate species were dominated at pH 2.3 and 4.2, respectively, and these species were not affected by addition of trityrosine. These results demonstrated that the His/His species increased by interaction with trityrosine at the initial stage of cyt c folding, whereas the heme coordination structure was not affected by trityrosine when the protein was completely unfolded or folded. Hydrophobic peptides thus may be useful to study the effects of hydrophobic interactions on protein folding.  相似文献   

13.
Dennison C  Lawler AT 《Biochemistry》2001,40(10):3158-3166
The effect of pH on Cu(I) and Cu(II) umecyanin (UCu), a phytocyanin obtained from horseradish roots, has been studied by electronic and NMR spectroscopy and using direct electrochemical measurements. A pK(a) value of approximately 9.5-9.8 is observed for the alkaline transition in UCu(II), and this leads to a slightly altered active site structure, as indicated by the changes in the paramagnetic 1H NMR spectrum. Electrochemical studies show that the pK(a) value for this transition in UCu(I) is 9.9. The alkaline transition is caused by the deprotonation of a surface lysine residue, with Lys96 being the most likely candidate. The isotropically shifted resonances in the (1)H NMR spectrum of UCu(II) also shift upon lowering the pH (pK(a) 5.8), and this can be assigned to the protonation of the surface (noncoordinating) His65 residue. This histidine titrates in UCu(I) with a pK(a) of 6.3. The reduction potential of the protein in this range is also dependent on pH, and pK(a) values matching those from NMR, for the two oxidation states of the protein, are obtained. There is no evidence for either of the active site histidines (His44 and His90) titrating in UCu(I) in the pH range studied (down to pH 3.7). Also highlighted in these studies are the remarkable active site similarities between umecyanin and the other phytocyanins which possess an axial Gln ligand.  相似文献   

14.
Conventional opioids including opioid peptides require an "opioid" nitrogen which exists in protonated state while interacting with the receptor. In the present paper we demonstrate that the Tyr-Pro-Gly-Phe-Leu-Thr hexapeptide sequence accepts N-terminal substituents such as N-t-Boc, N-phenylacetyl and N-diphenylacetyl where the N cannot become protonated, as well as "traditional" substitutions such as N,N-diallyl, where protonation is likely under physiological conditions. The opioid peptides bearing these substituents are pure antagonists of medium affinity (Ke values in the mouse vas deferens bioassay against [Met5]-enkephalin are in the 3 x 10(-7)-4 x 10(-6) M range) with a high delta receptor preference (50-350-fold delta over mu selectivity ratios).  相似文献   

15.
The kinetics and spectra of the reactions of clozapine with compounds I and II of myeloperoxidase were investigated using both single- and sequential-mixing stopped-flow techniques, steady-state kinetics, and spectrophotometric measurements. The results show conclusively that both compounds I and II are reduced in one-electron reactions with clozapine. At pH 7.0 the rate constant for compound I reacting with clozapine is (1.5 +/- 0.1) x 10(6) M(-1) s(-1) and for compound II (4.8 +/- 0.1) x 10(4) M(-1) s(-1). The physiological pH of 7.4 was found to be optimal for the oxidation of clozapine by compound I. The rate constant for compound I reacting with ascorbate is (1.1 +/- 0.1) x 10(6) M(-1) s(-1) and for compound II (1.1 +/- 0.2) x 10(4) M(-1) s(-1), both obtained at pH 7.0. Experiments with both clozapine and ascorbate present showed that ascorbate acts both as a competitive inhibitor and free radical scavenger.  相似文献   

16.
Herold S  Exner M  Nauser T 《Biochemistry》2001,40(11):3385-3395
The second-order rate constants for the reactions between nitrogen monoxide and oxymyoglobin or oxyhemoglobin, determined by stopped-flow spectroscopy, increase with increasing pH. At pH 7.0 the rates are (43.6 +/- 0.5) x 10(6) M(-1) x s(-1) for oxymyoglobin and (89 +/- 3) x 10(6) M(-1) x s(-1) for oxyhemoglobin (per heme), whereas at pH 9.5 they are (97 +/- 3) x 10(6) M(-1) x s(-1) and (144 +/- 3) x 10(6) M(-1) x s(-1), respectively. The rate constants for the reaction between oxyhemoglobin and NO* depend neither on the association grade of the protein (dimer/tetramer) nor on the concentration of the phosphate buffer (100-1 mM). The nitrogen monoxide-mediated oxidations of oxymyoglobin and oxyhemoglobin proceed via intermediate peroxynitrito complexes which were characterized by rapid scan UV/vis spectroscopy. The two complexes MbFe(III)OONO and HbFe(III)OONO display very similar spectra with absorption maxima around 500 and 635 nm. These species can be observed at alkaline pH but rapidly decay to the met-form of the proteins under neutral or acidic conditions. The rate of decay of MbFe(III)OONO increases with decreasing pH and is significantly larger than those of the analogous complexes of the two subunits of hemoglobin. No free peroxynitrite is formed during these reactions, and nitrate is formed quantitatively, at both pH 7.0 and 9.0. This result indicates that, as confirmed from protein analysis after reacting the proteins with NO* for 10 times, when peroxynitrite is coordinated to the heme of myoglobin or hemoglobin it rapidly isomerizes to nitrate without nitrating the globins in physiologically significant amounts.  相似文献   

17.
In human myeloperoxidase the heme is covalently attached to the protein via two ester linkages between the carboxyl groups of Glu242 and Asp94 and modified methyl groups on pyrrole rings A and C of the heme as well as a sulfonium ion linkage between the sulfur atom of Met243 and the beta-carbon of the vinyl group on pyrrole ring A. In the present study, wild-type recombinant myeloperoxidase (recMPO) and the variant Glu242Gln were produced in Chinese hamster ovary cells and investigated in a comparative sequential-mixing stopped-flow study in order to elucidate the role of the Glu242-heme ester linkage in the individual reaction steps of both the halogenation and peroxidase cycle. Disruption of the ester bond increased heme flexibility, blue shifted the UV-vis spectrum, and, compared with recMPO, decelerated cyanide binding (1.25 x 10(4) versus 1.6 x 10(6) M(-)(1) s(-)(1) at pH 7 and 25 degrees C) as well as compound I formation mediated by either hydrogen peroxide (7.8 x 10(5) versus 1.9 x 10(7) M(-)(1) s(-)(1)) or hypochlorous acid (7.5 x 10(5) versus 2.3 x 10(7) M(-)(1) s(-)(1)). The overall chlorination and bromination activity of Glu242Gln was 2.0% and 24% of recMPO. The apparent bimolecular rate constants of compound I reduction by chloride (65 M(-)(1) s(-)(1)), bromide (5.4 x 10(4) M(-)(1) s(-)(1)), iodide (6.4 x 10(5) M(-)(1) s(-)(1)), and thiocyanate (2.2 x10(5) M(-)(1) s(-)(1)) were 500, 25, 21, and 63 times decreased compared with recMPO. By contrast, Glu242Gln compound I reduction by tyrosine was only 5.4 times decreased, whereas tyrosine-mediated compound II reduction was 60 times slower compared with recMPO. The effects of exchange of Glu242 on electron transfer reactions are discussed.  相似文献   

18.
Stoichiometry, stability constants and solution structures of the copper(II) complexes of the (1-16H), (1-28H), (1-16M), (1-28M), (Ac-1-16H) and (Ac-1-16M) fragments of human (H) and mouse (M) beta-amyloid peptide were determined in aqueous solution in the pH range 2.5-10.5. The potentiometric and spectroscopic data (UV-Vis, CD, EPR) show that acetylation of the amino terminal group induces significant changes in the coordination properties of the (Ac-1-16H) and (Ac-1-16M) peptides compared to the (1-16H) and (1-16M) fragments, respectively. The (Ac-1-16H) peptide forms the 3N [N(Im)(6), N(Im)(13), N(Im)(14)] complex in a wide pH range (5-8), while for the (Ac-1-16M) fragment the 2N [N(Im)(6), N(Im)(14)] complex in the pH range 5-7 is suggested. At higher pH values sequential amide nitrogens are deprotonated and coordinated to copper(II) ions. The N-terminal amino group of the (1-16) and (1-28) fragments of human and mouse beta-amyloid peptide takes part in the coordination of the metal ion, although, at pH above 9 the complexes with the 4N [N(Im), 3N(-)] coordination mode are formed. The phenolate -OH group of the Tyr(10) residue of the human fragments does not coordinate to the metal ion.  相似文献   

19.
We have studied the binding of inositol pentaphosphate (IPP) to the hemoglobins from two species of goose living at low and high altitudes, using the proton absorption method. Measurements were done at 25 and 37 degrees C in a pH range between 6.0 and 8.8. The bird hemoglobins show a high affinity and a binding stoichiometry of 1 IPP molecule/hemoglobin tetramer both in the ligated and unligated state, indicating the same binding site for IPP in oxy- and deoxyhemoglobin. The results indicate that the interaction of IPP with both geese hemoglobins is very similar. For the deoxyhemoglobins of both species the IPP-binding constant shows a strong pH dependence extending over a wide pH range (i.e. +/- 2 x 10(6) M at pH 8.8 and +/- 6 x 10(10) M at pH 6.0). The binding constant of IPP for the oxyhemoglobins shows a much weaker pH dependence (i.e. +/- 4 x 10(4) M at pH 8.8 and +/- 3 x 10(6) M at pH 6.0), indicating that the interaction of IPP with the goose hemoglobin is strongly dependent on the state of ligation of the protein. The IPP binding constants for the oxy- and deoxyhemoglobins are found to be in good agreement with the IPP-induced change in oxygen affinity of both hemoglobins as estimated from oxygen binding curves.  相似文献   

20.
Nyarko A  Cochrun L  Norwood S  Pursifull N  Voth A  Barbar E 《Biochemistry》2005,44(43):14248-14255
LC8 is a highly conserved light-chain subunit of cytoplasmic dynein that interacts with a wide variety of cellular proteins and is presumed to play a fundamental role in dynein assembly and cargo recruitment and in the assembly of protein complexes unrelated to dynein. LC8 is a dimer at physiological pH but dissociates to a folded monomer at pH < 4.8. We have suggested that acid-induced dimer dissociation is due to protonation of His 55, which is stacked against His 55' and completely buried in the dimer interface. In this work, we show that the pH-induced dissociation is reversible and indeed governed by the ionization state of His 55. Mutagenesis of His 55 to Lys results in a monomer in the pH range of 3-8, while the mutation to Ala results in a dimer in the same pH range. Mutations that disrupt intermolecular hydrogen bonds between Tyr 65 and Lys 44' and His 55 and Thr 67' do not change the association state of the dimer. Titration curves for His 55 and the two other histidines, His 72 and 68, were determined by (13)C-(1)H NMR for H55K and for WT-LC8 in the monomeric and dimeric states. The pK(a) values of His 72 and His 68 are 6 in the WT dimer and 6.2-6.5 in monomeric H55K, while the pK(a) of His 55 is about 4.5 in the WT dimer. These results indicate that deprotonation of His 55 is linked to dimer formation and that mutation of His 55 to a small neutral residue or to a positively charged residue uncouples the protonation and dissociation processes.  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