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 93 毫秒
1.
The uptake of glutamate in nerve synapses is carried out by the excitatory amino acid transporters (EAATs), involving the cotransport of a proton and three Na+ ions and the countertransport of a K+ ion. In this study, we use an EAAT3 homology model to calculate the pKa of several titratable residues around the glutamate binding site to locate the proton carrier site involved in the translocation of the substrate. After identifying E374 as the main candidate for carrying the proton, we calculate the protonation state of this residue in different conformations of EAAT3 and with different ligands bound. We find that E374 is protonated in the fully bound state, but removing the Na2 ion and the substrate reduces the pKa of this residue and favors the release of the proton to solution. Removing the remaining Na+ ions again favors the protonation of E374 in both the outward- and inward-facing states, hence the proton is not released in the empty transporter. By calculating the pKa of E374 with a K+ ion bound in three possible sites, we show that binding of the K+ ion is necessary for the release of the proton in the inward-facing state. This suggests a mechanism in which a K+ ion replaces one of the ligands bound to the transporter, which may explain the faster transport rates of the EAATs compared to its archaeal homologs.  相似文献   

2.
The uptake of glutamate in nerve synapses is carried out by the excitatory amino acid transporters (EAATs), involving the cotransport of a proton and three Na+ ions and the countertransport of a K+ ion. In this study, we use an EAAT3 homology model to calculate the pKa of several titratable residues around the glutamate binding site to locate the proton carrier site involved in the translocation of the substrate. After identifying E374 as the main candidate for carrying the proton, we calculate the protonation state of this residue in different conformations of EAAT3 and with different ligands bound. We find that E374 is protonated in the fully bound state, but removing the Na2 ion and the substrate reduces the pKa of this residue and favors the release of the proton to solution. Removing the remaining Na+ ions again favors the protonation of E374 in both the outward- and inward-facing states, hence the proton is not released in the empty transporter. By calculating the pKa of E374 with a K+ ion bound in three possible sites, we show that binding of the K+ ion is necessary for the release of the proton in the inward-facing state. This suggests a mechanism in which a K+ ion replaces one of the ligands bound to the transporter, which may explain the faster transport rates of the EAATs compared to its archaeal homologs.  相似文献   

3.
The solvent kinetic isotope effects (SKIE) on the yeast α-glucosidase-catalyzed hydrolysis of p-nitrophenyl and methyl-d-glucopyranoside were measured at 25 °C. With p-nitrophenyl-d-glucopyranoside (pNPG), the dependence of kcat/Km on pH (pD) revealed an unusually large (for glycohydrolases) solvent isotope effect on the pL-independent second-order rate constant, DOD(kcat/Km), of 1.9 (±0.3). The two pKas characterizing the pH profile were increased in D2O. The shift in pKa2 of 0.6 units is typical of acids of comparable acidity (pKa=6.5), but the increase in pKa1 (=5.7) of 0.1 unit in going from H2O to D2O is unusually small. The initial velocities show substrate inhibition (Kis/Km~200) with a small solvent isotope effect on the inhibition constant [DODKis=1.1 (±0.2)]. The solvent equilibrium isotope effects on the Kis for the competitive inhibitors d-glucose and α-methyl d-glucoside are somewhat higher [DODKi=1.5 (±0.1)]. Methyl glucoside is much less reactive than pNPG, with kcat 230 times lower and kcat/Km 5×104 times lower. The solvent isotope effect on kcat for this substrate [=1.11 (±0. 02)] is lower than that for pNPG [=1.67 (±0.07)], consistent with more extensive proton transfer in the transition state for the deglucosylation step than for the glucosylation step.  相似文献   

4.
The reductive carboxylation of ribulose-5-phosphate (Ru5P) by 6-phosphogluconate dehydrogenase (6PGDH) from Candida utilis was investigated using kinetic isotope effects. The intrinsic isotope effect for proton abstraction from Ru5P was found at 4.9 from deuterium isotope effects on V and V/K and from tritium isotope effects on V/K. The presence of 6-phosphogluconate (6PG) in the assay mixture changes the magnitude of the observed isotope effects. In the absence of 6PG D(V/K) and D(V) are 1.68 and 2.46, respectively, whereas the presence of 6PG increases D(V/K) to 2.84 and decreases D(V) to 1.38. A similar increase of T(V/K) is observed as 6PG builds up in the reaction mixture. These data indicate that in the absence of 6PG, a slow step, which precedes the chemical process, is rate-limiting for the reaction, whereas in the presence of 6PG, the rate-limiting step follows the isotope-sensitive step. Kinetic analysis of reductive carboxylation shows that 6PG at low concentrations decreases the Km of Ru5P, whereas at higher concentrations, the usual competitive pattern is observed. These data indicate that full activity of 6PGDH is achieved when one subunit carries out the catalysis and the other subunit carries an unreacted 6PG. Thus, 6PG is like an allosteric activator of 6PGDH.  相似文献   

5.
A key feature of the modified Q-cycle of the cytochrome bc1 and related complexes is a bifurcation of QH2 oxidation involving electron transfer to two different acceptor chains, each coupled to proton release. We have studied the kinetics of proton release in chromatophore vesicles from Rhodobacter sphaeroides, using the pH-sensitive dye neutral red to follow pH changes inside on activation of the photosynthetic chain, focusing on the bifurcated reaction, in which 4H+are released on complete turnover of the Q-cycle (2H+/ubiquinol (QH2) oxidized). We identified different partial processes of the Qo-site reaction, isolated through use of specific inhibitors, and correlated proton release with electron transfer processes by spectrophotometric measurement of cytochromes or electrochromic response. In the presence of myxothiazol or azoxystrobin, the proton release observed reflected oxidation of the Rieske iron?sulfur protein. In the absence of Qo-site inhibitors, the pH change measured represented the convolution of this proton release with release of protons on turnover of the Qo-site, involving formation of the ES-complex and oxidation of the semiquinone intermediate. Turnover also regenerated the reduced iron-sulfur protein, available for further oxidation on a second turnover. Proton release was well-matched with the rate limiting step on oxidation of QH2 on both turnovers. However, a minor lag in proton release found at pH?7 but not at pH?8 might suggest that a process linked to rapid proton release on oxidation of the intermediate semiquinone involves a group with a pK in that range.  相似文献   

6.
Bacteriorhodopsin (BR) functions as a light-driven proton pump, whereas Anabaena sensory rhodopsin (ASR) is believed to function as a photosensor despite the high similarity in their protein sequences. In Fourier transform infrared (FTIR) spectroscopic studies, the lowest O-D stretch for D2O was observed at ∼2200 cm−1 in BR but was significantly higher in ASR (>2500 cm−1), which was previously attributed to a water molecule near the Schiff base (W402) that is H-bonded to Asp-85 in BR and Asp-75 in ASR. We investigated the factors that differentiate the lowest O-D stretches of W402 in BR and ASR. Quantum mechanical/molecular mechanical calculations reproduced the H-bond geometries of the crystal structures, and the calculated O-D stretching frequencies were corroborated by the FTIR band assignments. The potential energy profiles indicate that the smaller O-D stretching frequency in BR originates from the significantly higher pKa(Asp-85) in BR relative to the pKa(Asp-75) in ASR, which were calculated to be 1.5 and −5.1, respectively. The difference is mostly due to the influences of Ala-53, Arg-82, Glu-194–Glu-204, and Asp-212 on pKa(Asp-85) in BR and the corresponding residues Ser-47, Arg-72, Ser-188-Asp-198, and Pro-206 on pKa(Asp-75) in ASR. Because these residues participate in proton transfer pathways in BR but not in ASR, the presence of a strongly H-bonded water molecule near the Schiff base ultimately results from the proton-pumping activity in BR.  相似文献   

7.
The K+-dependent p-nitrophenylphosphatase activity catalyzed by purified (Na+ + K+)-ATPase from pig kidney shows substrate inhibition (Ki about 9.5 mM at 2.1 mM Mg2+). Potassium antagonizes and sodium favours this inhibition. In addition, K+ reduces the apparent affinity for substrate activation, whereas p-nitrophenyl phosphate reduces the apparent affinity for K+ activation. In the absence of Mg2+, p-nitrophenyl phosphate, as well as ATP, accelerates the release of Rb+ from the Rb+ occluded unphosphorylated enzyme. With no Mg2+ and with 0.5 mM KCl, trypsin inactivation of (Na+ + K+)-ATPase as a function of time follows a single exponential but is transformed into a double exponential when 1 mM ATP or 5 mM p-nitrophenyl phosphate are also present. In the presence of 3 mM MgCl2, 5 mM p-nitrophenyl phosphate and without KCl the trypsin inactivation pattern is that described for the E1 enzyme form; the addition of 10 mM KCl changes the pattern which, after about 6 min delay, follows a single exponential. These results suggest that (i) the shifting of the enzyme toward the E1 state is the basis for substrate inhibition of the p-nitrophenulphosphatase acitivy of (Na+ + K+)-ATPase, and (ii) the substrate site during phosphatase activity is distinct from the low-affinity ATP site.  相似文献   

8.
A role for redox-active tyrosines has been demonstrated in many important biological processes, including water oxidation carried out by photosystem II (PSII) of oxygenic photosynthesis. The rates of tyrosine oxidation and reduction and the Tyr/Tyr reduction potential are undoubtedly controlled by the immediate environment of the tyrosine, with the coupling of electron and proton transfer, a critical component of the kinetic and redox behavior. It has been demonstrated by Faller et al. that the rate of oxidation of tyrosine D (TyrD) at room temperature and the extent of TyrD oxidation at cryogenic temperatures, following flash excitation, dramatically increase as a function of pH with a pKa of ≈ 7.6 [Faller et al. 2001 Proc. Natl. Acad. Sci. USA 98, 14368-14373; Faller et al. 2001 Biochemistry 41, 12914-12920]. In this work, we investigated, using FTIR difference spectroscopy, the mechanistic reasons behind this large pH dependence. These studies were carried out on Mn-depleted PSII core complexes isolated from Synechocystis sp. PCC 6803, WT unlabeled and labeled with 13C6-, or 13C1(4)-labeled tyrosine, as well as on the D2-Gln164Glu mutant. The main conclusions of this work are that the pH-induced changes involve the reduced TyrD state and not the oxidized TyrD state and that TyrD does not exist in the tyrosinate form between pH 6 and 10. We can also exclude a change in the protonation state of D2-His189 as being responsible for the large pH dependence of TyrD oxidation. Indeed, our data are consistent with D2-His189 being neutral both in the TyrD and TyrD states in the whole pH6-10 range. We show that the interactions between reduced TyrD and D2-His189 are modulated by the pH. At pH greater than 7.5, the ν(CO) mode frequency of TyrD indicates that TyrD is involved in a strong hydrogen bond, as a hydrogen bond donor only, in a fraction of the PSII centers. At pH below 7.5, the hydrogen-bonding interaction formed by TyrD is weaker and TyrD could be also involved as a hydrogen bond acceptor, according to calculations performed by Takahashi and Noguchi [J. Phys. Chem. B 2007 111, 13833-13844]. The involvement of TyrD in this strong hydrogen-bonding interaction correlates with the ability to oxidize TyrD at cryogenic temperatures and rapidly at room temperature. A strong hydrogen-bonding interaction is also observed at pH 6 in the D2-Gln164Glu mutant, showing that the residue at position D2-164 regulates the properties of TyrD. The IR data point to the role of a protonatable group(s) (with a pKa of ≈ 7) other than D2-His189 and TyrD, in modifying the characteristics of the TyrD hydrogen-bonding interactions, and hence its oxidation properties. It remains to be determined whether the strong hydrogen-bonding interaction involves D2-His189 and if TyrD oxidation involves the same proton transfer route at low and at high pH.  相似文献   

9.
The catalytic mechanism of 6-phosphogluconate dehydrogenase requires the inversion of a Lys/Glu couple from its natural ionization state. The pKa of these residues in free and substrate bound enzymes has been determined measuring by ITC the proton release/uptake induced by substrate binding at different pH values. Wt 6-phosphogluconate dehydrogenase from Trypanosoma brucei and two active site enzyme mutants, K185H and E192Q were investigated. Substrate binding was accompanied by proton release and was dependent on the ionization of a group with pKa 7.07 which was absent in the E192Q mutant. Kinetic data highlighted two pKa, 7.17 and 9.64, in the enzyme–substrate complex, the latter being absent in the E192Q mutant, suggesting that the substrate binding shifts Glu192 pKa from 7.07 to 9.64. A comparison of wt and E192Q mutant appears to show that the substrate binding shifts Lys185 pKa from 9.9 to 7.17. By comparing differences in proton release and the binding enthalpy of wt and mutant enzymes, the enthalpic cost of the change in the protonation state of Lys185 and Glu192 was estimated at ≈ 6.1 kcal/mol. The change in protonation state of Lys185 and Glu192 has little effect on Gibbs free energy, 240–325 cal/mol. However proton balance evidences the dissociation of other group(s) that can be collectively described by a single pKa shift from 9.1 to 7.54. This further change in ionization state of the enzyme causes an increase of free energy with a total cost of 1.2–2.3 kcal/mol to set the enzyme into a catalytically competent form.  相似文献   

10.
The flavoprotein nitroalkane oxidase catalyzes the oxidative denitrification of a broad range of primary and secondary nitroalkanes to yield the respective aldehydes or ketones, hydrogen peroxide and nitrite. With nitroethane as substrate the D2O(kcat/KM) value is 0.6 and the D2Okcat value is 2.4. The kcat proton inventory is consistent with a single exchangeable proton in flight, while the kcat/KM is consistent with either a single proton in flight in the transition state or a medium effect. Increasing the solvent viscosity did not affect the kcat or kcat/KM value significantly, establishing that nitroethane binding is at equilibrium and that product release does not limit kcat.  相似文献   

11.
Alain Boussac  Anne Lise Etienne 《BBA》1984,766(3):576-581
In Tris-washed Photosystem-II particles we are able to induce an EPR signal in the dark by addition of an iridium salt (K2IrCl6). This signal is attributed to signal IIs (slow) (D+) and the redox titration gives an Em value of 760 mV for the couple D+D. On the basis of our previous studies on the equilibrium between D+Z and DZ+ (K = 104) (Boussac, A. and Etienne, A.L. (1982) Biochem. Biophys. Res. Commun. 109, 1200–1205), we therefore attribute a value of 1 V for the Em of the Z+Z couple. A second effect of K2IrCl6 is to modify the spectral characteristics of signal II. We conclude that K2IrCl6 is able to change the environment of the species from which signal IIs and signal IIf originate.  相似文献   

12.
Proton nmr spectroscopic evidence is presented for methylmercury(II) binding to the deprotonated amino groups in adenosine, 9-methyladenine, guanosine, 1-methylguanosine, and cytidine under basic conditions. Except for the guanosine case, 1H nmr spectra of the products from aqueous or ethanolic 1:1 mixtures of substrate and MeHgOH are consistent with methylmercuration of the deprotonated amino groups. Guanosine undergoes initial binding of MeHg to N1, and a second equivalent of MeHgOH is necessary to effect amino binding. The nmr spectra of the complexed adenine derivatives suggest that different geometrical isomers exist in (CD3)2SO solution, reflecting the partial double bond character of the C6N bond in these systems. Using a correlation relating the magnitude of the 199Hg-1H coupling constant (J) for MeHg-ligand complexes with the ligand pKa (J = ?3.88 pKa + 248.5, extending over 13 pK units, based on a variety of N and O donor ligands), estimates (± 0.3 pK unit) of the pKas of the amino groups of the above substrates have been made. In this way, pKa values of 15.5 (cytidine), 17.0 (adenosine and 9-methyladenine), 15.1 (guanosine), and 14.9 (1-methylguanosine) are obtained. In the cases where comparisons with literature pKa data can be made, good agreement is found.  相似文献   

13.
G. Renger  H.J. Eckert 《BBA》1981,638(1):161-171
The role of the protein matrix embedding the functionally active redox components of Photosystem II reaction centers has been studied by investigating the effects of procedures which modify the structure of proteins. In order to reduce the influence of the electron transport involving secondary donor and acceptor components, Triswashed chloroplasts were used which are completely deprived of their oxygen-evolving capacity. The functional activity was detected via absorption changes, reflecting at 334 and 690 or 834 nm the turnover of the primary plastoquinone acceptor, X320, and of the photochemically active chlorophyll a complex, Chl aII, respectively, and at 520 nm the transient formation of a transmembrane electric potential gradient. Under repetitive flash excitation of Tris-washed chloroplasts it was found that: (a) The relaxation kinetics at 690 nm become significantly accelerated in the presence of external electron donors. (b) Trypsin treatment blocks to a high degree the turnover of Chl aII and X320 unless exogenous acceptors are present, which directly oxidize X320?, such as K3Fe(CN)6. (c) In the presence of K3Fe(CN)6 the recovery kinetics of Chl aII and X320 are retarded markedly by trypsin, followed by a progressive decline in the extent thereof. (d) 2-(3-Chloro-4-trifluoromethyl)anilino-3,5-dinitrothiophene (ANT 2p), known to reduce the lifetime of S2 and S3 in normal chloroplasts, significantly accelerates the recovery of Chl aII. 10 μs kinetics are observed which correspond with the electron-transfer rate from D1 to Chl a+II. ANT 2p simultaneously retards the decay kinetics of X320? and of the electrochromic absorption changes. (e) The kinetic pattern of the electrochromic absorption changes is also affected by the salt content of the suspension. Under dark-adapted conditions, the 10 μs relaxation kinetics of the 834 nm absorption change due to the first flash are hardly affected by mild trypsinization of 5–10 min duration, whereas the amplitude decreases by approx. 30%. The data obtained in Tris-washed chloroplasts could consistently be interpreted as a modification of the back reaction between X320? and Chl a+II which is caused solely by a change in the reactivity of X320 due to trypsin-induced degradation of the native X320-B apoprotein. Furthermore, ADRY agents are inferred to stimulate cyclic electron flow, which leads to reduction of D+1 between the flashes. A simplified scheme is discussed which describes the functional organization of the reaction center complex.  相似文献   

14.
The tris(pyrazolyl)borate and related tripodal N-donor ligands originally developed by Trofimenko stabilize mononuclear compounds containing MoVIO2, MoVIO, MoVO, and MoIVO units and effectively inhibit their polynucleation in organic solvents. Dioxo-Mo(VI) complexes of the type LMoO2(SPh), where L = hydrotris(3,5-dimethylpyrazol-1-yl)borate (Tp), hydrotris(3-isopropylpyrazol-1-yl)borate (TpiPr), and hydrotris(3,5-dimethyl-1,2,4-triazol-1-yl)borate (Tz) and related derivatives are the only model systems that mimic the complete reaction sequence of sulfite oxidase, in which oxygen from water is ultimately incorporated into product. The quasi-reversible, one-electron reduction of TpMoO2(SPh) in acetonitrile exhibits a positive potential shift upon addition of a hydroxylic proton donor, and the magnitude of the shift correlates with the acidity of the proton donor. These reductions produce two Mo(V) species, [TpMoVO2(SPh)] and TpMoVO(OH)(SPh), that are related by protonation. Measurement of the relative amounts of these two Mo(V) species by EPR spectroscopy enabled the pKa of the MoV(OH) unit in acetonitrile to be determined and showed it to be several pKa units smaller than that for water in acetonitrile. Similar electrochemical-EPR experiments for TpiPrMoO2(SPh) indicated that the pKa for its MoV(OH) unit was ∼1.7 units smaller than that for TpMoVO(OH)(SPh). Density functional theory calculations also predict a smaller pKa for TpiPrMoVO(OH)(SPh) compared to TpMoVO(OH)(SPh). Analysis of these results indicates that coupled electron-proton transfer (CEPT) is thermodynamically favored over the indirect process of metal reduction followed by protonation. The crystal structure of TpiPrMoO2(SPh) is also presented.  相似文献   

15.
We studied interaction of the lectin from the bark of Golden Rain shrub (Laburnum anagyroides, LABA) with a number of basic fucose-containing carbohydrate antigens by changes in its tryptophan fluorescence. The strongest LABA binding was observed for the trisaccharide H of type 6 [α-L-Fucp-(1-2)-β-D-Galp-(1-4)-D-Glc, K a = 4.2 × 103 M?1]. The following antigens were bound with a weaker affinity: H-disaccharide α-L-Fucp-(1-2)-D-Gal, a glucoanalogue of tetrasaccharide Ley α-L-Fucp-(1-2)-β-D-Galp-(1-4)-[α-L-Fucp-(1-3)]-D-Glc, and 6-fucosyl-N-acetylglucosamine, a fragment of core of the N-glycans family (K a 1.1?1.7 × 103 M?1). The lowest binding was observed for L-fucose (K a = 2.7 × 102 M?1) and trisaccharide Lea, (β-Galp-(1-3)-[α-L-Fucp-(1-4)]-GlcNAc (K a = 6.4 × 102 M?1). The Led, Lea, and Lex pentasaccharides and Leb hexasaccharide were not bound to LABA.  相似文献   

16.
1. The absorption spectra of deutero- and proto-ferrihaem in aqueous solution at 25°C show marked changes with concentration and pH in the Soret band region. Quantitative studies of these phenomena imply that they are associated with ferrihaem dimerization and with protolytic equilibria involving monomeric (M) and dimeric (D) ferrihaem species according to the scheme: [Formula: see text] 2. For deuteroferrihaem we obtain K=1.9×10−2, pKa(M)=7.1, pKa(D)=7.4. Protoferrihaem has a much higher dimerization constant, K=4.5 and pKa(D)=7.5 (pKa(M) is not accessible). 3. Possible structural relationships between monomeric and dimeric ferrihaem species in solution are discussed in relation to recent work on the oxo-bridged nature of crystalline ferrihaem dimers.  相似文献   

17.
1. The properties of 3,5-di-tert-butyl-4-hydroxybenzylidenemalononitrile (SF 6847) were studied chemically and spectroscopically. Two molecular species of SF6847 were identified: the undissociated form (SFH; ?363, 10 mM?1) and the dissociated form (SF?; ?454, 35 mM?1). The pKa value of the molecule was determined to be 6.9.2. On the basis of these properties the interactions of SF6847 with liposomes and valinomycin · K+ were studied. The partition constants of SFH (Knp and SF? (K?p) to liposomes were determined separately; Knp was 56 mM?1 and was independent of the pH of the medium, whereas K?p dependend greatly on the pH, being 1.2 mM?1 at pH 7.0 and 2.9 mM?1 at pH 8.0. Using these values, the partition constant of total SF6847 (Kp) was calculated and found to be essentially the same as that calculated from the kinetics of proton uptake. It was concluded that the amount of SF? bound to liposomes is rate limiting for proton uptake.3. The effects of membrane potential on partition constants were studied. The K?p decreased greatly upon generation of a membrane potential negative inside the liposomes but increased upon generation of a membrane potential positive inside the liposomes.4. The interaction of SF6847 with valinomycin in aqueous solution and in liposomes was demonstrated only in the presence of potassium ion. Potassium ion could not be replaced by sodium ion. Evidence was obtained for the formation of the ternary complex valinomycin · K+ · SF? in liposomes and in hexane. It was concluded that SF? became more soluble in the liposomal membranes on formation of this ternary complex. All these results support our proposed mechanism for the proton uptake cycle (Yamaguchi, A. and Anraku, Y. (1978) Biochim. Biophys. Acta 501, 136–149).  相似文献   

18.
The binding of Mn2+, Ca2+, and rare earth ions to apoconcanavalin A has been studied by water proton relaxation enhancement, electron paramagnetic resonance spectroscopy, and fluorescence spectroscopy. An electron paramagnetic resonance and water proton relaxation rate study of the titration of apoconcanavalin A with Mn2+ gives evidence of two equivalent binding sites per monomer with KD = 50 μm ± 4 μm. When a similar Mn2+ titration of apoconcanavalin A is performed in the presence of Ca2+ ion, very little free Mn2+ is detected by electron paramagnetic resonance until the two Mn2+ binding sites per monomer are filled. The substitution of a rare earth ion for Ca2+ ion in the above experiment often resulted in a slight displacement of Mn2+ from the transition metal site as detected by electron paramagnetic resonance. A water proton relaxation rate study of the titration of apoconcanavalin A with Gd3+ reflects two binding sites with a KD = 40 μm ± 4 μm and two with a KD = 200 μm ± 50 μm. The fluorescence emission spectrum of concanavalin A (λem = 340 nm) is slightly quenched by the addition of Tb3+ while Tb3+ fluorescence is greatly enhanced. A fluorometric titration of apoconcanavalin A with Tb3+ also reflects two sites with a KD = 40 μm ± 15 μm and two with a KD = 270 μm ± 50 μm.  相似文献   

19.
Nine resonances in the 270 MHz proton magnetic resonance spectrum of human carbonic anhydrase B have been identified with imidazole C(2) protons of histidine residues, six of which are observed to titrate with pKa values in the range 4.7 to 7.4. The behaviour of the nine resonances has been studied in the presence of the inhibitors, iodide, cyanide, acetate, hexacyanochromate, and imidazole. Measurements have also been made of the enzyme in its apo, cobalt, and mono-alkylated forms. Used in conjunction with the crystal structure, these results have enabled the tentative assignment of all nine resonances to particular histidine residues in the amino-acid sequence. Three of the active-site histidines at positions 64, 67, and 200 have low pKa values and cannot be directly linked to the activity of the enzyme. However, the resonances assigned to the three metal-liganding histidines do exhibit changes on anion binding and with pH, which parallel changes in the esterase activity. These results are consistent with the model of an ionizable water molecule bound to the zinc ion.Linewidth measurements of the resonances of the histidine residues on the enzyme surface are used to estimate pseudo-first-order rate constants of the order of 4 × 103 s?1 for D+ exchange between imidazole N and solvent in the absence of buffer. These rates are observed to increase in the presence of small amounts of the buffers Tris and imidazole.  相似文献   

20.
D. Kleinfeld  M.Y. Okamura  G. Feher 《BBA》1984,766(1):126-140
The electron-transfer reactions and thermodynamic equilibria involving the quinone acceptor complex in bacterial reaction centers from R. sphaeroides were investigated. The reactions are described by the scheme: We found that the charge recombination pathway of D+QAQ?B proceeds via the intermediate state D+Q?AQB, the direct pathway contributing less than approx. 5% to the observed recombination rate. The method used to obtain this result was based on a comparison of the kinetics predicted for the indirect pathway (given by the product kAD-times the fraction of reaction centers in the Q?AQB state) with the observed recombination rate, kobsD+ →D. The kinetic measurements were used to obtain the pH dependence (6.1 ? pH ? 11.7) of the free energy difference between the states Q?AQB and QAQ?B. At low pH (less than 9) QAQ?B is stabilized relative to Q?AQB by 67 meV, whereas at high pH Q?AQB is energetically favored. Both Q?A and Q?B associate with a proton, with pK values of 9.8 and 11.3, respectively. The stronger interaction of the proton with Q?B provides the driving force for the forward electron transfer.  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