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 46 毫秒
1.
The PsaD subunit of photosystem I is one of the central polypeptides for the interaction with ferredoxin, its acidic electron acceptor. In the cyanobacterium Synechocystis 6803, this role is partly performed by a sequence extending approximately from histidine 97 to arginine 119, close to the C-terminus. In the present work, acidic amino acids D100, E105, and E109 are shown to moderate the affinity of Photosystem I for ferredoxin. Most single replacements of these residues by neutral amino acids increased the affinity for ferredoxin, resulting in a dissociation constant as low as 0.015 microM for the E105Q mutant (wild-type K(D) = 0.4 microM). This is the first report on the limitation of photosystem I affinity for ferredoxin due to acidic amino acids from PsaD subunit. It highlights the occurrence of a negative control on the binding during the formation of transient complexes between electron carriers.  相似文献   

2.
The electron transfer cascade from photosystem I to NADP+ was studied at physiological pH by flash-absorption spectroscopy in a Synechocystis PCC6803 reconstituted system comprised of purified photosystem I, ferredoxin, and ferredoxin-NADP+ reductase. Experiments were conducted with a 34-kDa ferredoxin-NADP+ reductase homologous to the chloroplast enzyme and a 38-kDa N-terminal extended form. Small differences in kinetic and catalytic properties were found for these two forms, although the largest one has a 3-fold decreased affinity for ferredoxin. The dissociation rate of reduced ferredoxin from photosystem I (800 s(-1)) and the redox potential of the first reduction of ferredoxin-NADP+ reductase (-380 mV) were determined. In the absence of NADP+, differential absorption spectra support the existence of a high affinity complex between oxidized ferredoxin and semireduced ferredoxin-NADP+ reductase. An effective rate of 140-170 s(-1) was also measured for the second reduction of ferredoxin-NADP+ reductase, this process having a rate constant similar to that of the first reduction. In the presence of NADP+, the second-order rate constant for the first reduction of ferredoxin-NADP+ reductase was 20% slower than in its absence, in line with the existence of ternary complexes (ferredoxin-NADP+ reductase)-NADP+-ferredoxin. A single catalytic turnover was monitored, with 50% NADP+ being reduced in 8-10 ms using 1.6 microM photosystem I. In conditions of multiple turnover, we determined initial rates of 360-410 electrons per s and per ferredox-in-NADP+ reductase for the reoxidation of 3.5 microM photoreduced ferredoxin. Identical rates were found with photosystem I lacking the PsaE subunit and wild type photosystem I. This suggests that, in contrast with previous proposals, the PsaE subunit is not involved in NADP+ photoreduction.  相似文献   

3.
Ferredoxin reduction by photosystem I has been studied by flash-absorption spectroscopy. Aspartate residues 20, 57, and 60 of ferredoxin were changed to alanine, cysteine, arginine, or lysine. On the one hand, electron transfer from photosystem I to all mutated ferredoxins still occurs on a microsecond time scale, with half-times of ferredoxin reduction mostly conserved compared to wild-type ferredoxin. On the other hand, the total amplitude of the fast first-order reduction varies largely when residues 57 or 60 are modified, in apparent relation to the charge modification (neutralized or inverted). Substituting these two residues for lysine or arginine induce strong effects on ferredoxin binding (up to sixfold increase in K(D)), whereas the same substitution on aspartate 20, a spatially related residue, results in moderate effects (maximum twofold increase in K(D)). In addition, double mutations to arginine or lysine were performed on both aspartates 57 and 60. The mutated proteins have a 15- to 20-fold increased K(D) and show strong modifications in the amplitudes of the fast reduction kinetics. These results indicate that the acidic area of ferredoxin including aspartates 57 and 60, located opposite to the C-terminus, is crucial for high affinity interactions with photosystem I.  相似文献   

4.
The arginine at position 148 is highly conserved in the inward rectifier K+ channel family. Increases of external pH decrease the single-channel conductance in mutant R148H of the Kir2.1 channel (arginine is mutated into histidine) but not in the wild type channel. Moreover, in 100 mM external K+, varying external pH induced biphasic changes of open channel noise, which peaks at around pH 7.4 in the R148H mutant but not in the wild type channel. The maximum single-channel conductances are higher in the wild type channel and R148H mutant at pH 6.0 than those in the R148H mutant at pH 7.4. However, the maximal conductance is achieved with much lower external [K+] for the latter. Interestingly, the single-channel conductances and open channel noise of the wild type channel at pH 6. 0 and the R148H mutant at pH 6.0 and 7.4 become the same in [K+] = 10 mM. These results indicate that the residue at position 148 is accessible to the external H+ and probably is involved in the formation of two K+ binding sites in the external pore mouth. Effective repulsion between permeating K+ ions in this area requires a positive charge at position 148, and such K+-K+ interaction is the essential mechanism underlying high K+ conduction rate through the Kir2.1 channel pore.  相似文献   

5.
A mutant of Synechocystis PCC 6803, deficient in psaE, assembles photosystem I reaction centers without the PsaE subunit. Under conditions of acceptor-side rate-limited photoreduction assays in vitro (with 15 microM plastocyanin included), using 100 nM ferredoxin:NADP(+) reductase (FNR) and either Synechocystis flavodoxin or spinach ferredoxin, lower rates of NADP(+) photoreduction were measured when PsaE-deficient membranes were used, as compared to the wild type. This effect of the psaE mutation proved to be due to a decrease of the apparent affinity of the photoreduction assay system for the reductase. In the psaE mutant, the relative petH (encoding FNR) expression level was found to be significantly increased, providing a possible explanation for the lack of a phenotype (i.e., a decrease in growth rate) that was expected from the lower rate of linear electron transport in the mutant. A kinetic model was constructed in order to simulate the electron transfer from reduced plastocyanin to NADP(+), and test for possible causes for the observed change in affinity for FNR. The numerical simulations predict that the altered reduction kinetics of ferredoxin, determined for the psaE mutant [Barth, P., et al., (1998) Biochemistry 37, 16233-16241], do not significantly influence the rate of linear electron transport to NADP(+). Rather, a change in the dissociation constant of ferredoxin for FNR does affect the saturation profile for FNR. We therefore propose that the PsaE-dependent transient ternary complex PSI/ferredoxin/FNR is formed during linear electron transport. Using the yeast two-hybrid system, however, no direct interaction could be demonstrated in vivo between FNR and PsaE fusion proteins.  相似文献   

6.
Five point mutations (R88H, R88P, T118I, 959delT, R468Q) previously identified in the iduronate-2-sulfatase (IDS) gene of Italian Hunter patients were expressed in COS cells to evaluate their functional consequence on enzyme activity, processing and intracellular localization. The 88 arginine residue belongs to the CXPSR pentapeptide conserved in all human sulfatases, where cysteine modification to formylglycine is required for enzyme activity. Substitution of arginine with histidine residue resulted in 13.7% residual enzyme activity, with an apparent K(m) value (133 microM) lower than that found for the normal enzyme (327 microM), indicating a higher affinity for the substrate; substitution of arginine with proline resulted in total absence of residual activity, in agreement with the phenotypes observed in patients carrying R88H and R88P mutations. For the four missense mutations, pulse-chase labelling experiments showed an apparently normal maturation; however, subcellular fractionation demonstrated poor transport to lysosomes. Therefore, residues 88, 118 and 468 appear to be not essential for processing but important for IDS conformation.  相似文献   

7.
PsaD subunit of Synechocystis sp PCC 6803 photosystem I (PSI) plays a critical role in the stability of the complex and is part of the docking site for ferredoxin (Fd). In the present study we describe major physiological and biochemical effects resulting from mutations in the accessible C-terminal end of the protein. Four basic residues were mutated: R111, K117, K131, and K135, and a large 36-amino acid deletion was generated at the C terminus. PSI from R111C mutant has a 5-fold decreased affinity for Fd, comparable with the effect of the C terminus deletion, and NADP+ is photoreduced with a 2-fold decreased rate, without consequence on cell growth. The K117A mutation has no effect on the affinity for Fd, but decreases the stability of PsaE subunit, a loss of stability also observed in R111C and the deletion mutants. The double mutation K131A/K135A does not change Fd binding and reduction, but decreases the overall stability of PSI and impairs the cell growth at temperatures above 30 degrees C. Three mutants, R111C, K117A, and the C-terminal deleted exhibit a higher content of the trimeric form of PSI, in apparent relation to the removal of solvent accessible positive charges. Various regions in the C terminus of cyanobacterial PsaD thus are involved in Fd strong binding, PSI stability, and accumulation of trimeric PSI.  相似文献   

8.
Gerharz T  Reinelt S  Kaspar S  Scapozza L  Bott M 《Biochemistry》2003,42(19):5917-5924
The sensor kinase CitA and the response regulator CitB of Klebsiella pneumoniae form the paradigm of a subfamily of bacterial two-component regulatory systems that are capable of sensing tri- or dicarboxylates in the environment and then induce transporters for the uptake of these compounds. We recently showed that the separated periplasmic domain of CitA, termed CitAP (encompasses residues 45-176 supplemented with an N-terminal methionine residue and a C-terminal hexahistidine tag), is a highly specific citrate receptor with a K(d) of 5.5 microM at pH 7. To identify positively charged residues involved in binding the citrate anion, each of the arginine, lysine, and histidine residues in CitAP was exchanged for alanine, and the resulting 17 muteins were analyzed by isothermal titration calorimetry (ITC). In 12 cases, the K(d) for citrate was identical to that of wild-type CitAP or slightly changed (3.9-17.2 microM). In one case (R98A), the K(d) was 6-fold decreased (0.8 microM), whereas in four cases (R66A, H69A, R107A, and K109A) the K(d) was 38- to >300-fold increased (0.2 to >1 mM). The secondary structure of the latter five proteins in their apo-form as deduced from far-UV circular dichroism (CD) spectra did not differ from the apo-form of wild-type CitAP; however, all of them showed an increased thermostability. Citrate increased the melting point (T(m)) of wild-type CitAP and mutein R98A by 6.2 and 9.5 degrees C, respectively, but had no effect on the T(m) of the four proteins with disturbed binding. Three of the residues important for citrate binding (R66, H69, and R107) are highly conserved in the CitA subfamily of sensor kinases, indicating that they might be involved in ligand binding by many of these sensor kinases.  相似文献   

9.
Guinamard R  Akabas MH 《Biochemistry》1999,38(17):5528-5537
The cystic fibrosis transmembrane conductance regulator forms an anion-selective channel. We previously showed that charge selectivity, the ability to discriminate between anions and cations, occurs near the cytoplasmic end of the channel. The molecular determinants of charge selectivity, however, are unknown. We investigated the role of Arg352, a residue flanking the predicted cytoplasmic end of the M6 segment, in the mechanism of charge selectivity. We determined the Cl- to Na+ permeability ratio (PCl/PNa) from the reversal potential measured in a 10-fold NaCl gradient. For the wild type, PCl/PNa was 36 (range of 28-51). For the R352H mutant, PCl/PNa was dependent on cytoplasmic pH. At pH 5.4, the PCl/PNa was 33 (range of 27-41), similar to that of the wild type, but at pH 7.2, where the histidine should be largely uncharged, PCl/PNa was 3 (range of 2.9-3.1). For the R352C and R352Q mutants, PCl/PNa was 7 (range of 6-8) and 4 (range of 3.5-4.4), respectively. Furthermore, Na+ which does not carry a significant fraction of the current through the wild type is measurably conducted through R352Q. Thus, the charge of the side chain at position 352 is a strong determinant of charge selectivity. In the wild type, the positive charge on Arg352 contributes to an electrostatic potential in the channel that forms a barrier to cation permeation. Mutation of Arg352 did not alter the halide selectivity sequence. Selectivity among halides must involve other residues.  相似文献   

10.
Affinity labeling studies of NADP(+)-glutamate dehydrogenase from Salmonella typhimurium have shown that the peptide Leu-282-Lys-286 is located near the coenzyme site [Haeffner-Gormley et al. (1991) J. Biol. Chem. 266, 5388-5394]. The present study was undertaken to evaluate the role of lysine-286. The mutant enzymes K286R, K286Q, and K286E were prepared by site-directed mutagenesis, expressed in Escherichia coli, and purified. The Vmax values (micromoles of NADPH per minute per milligram of protein) were similar for WT (270), K286R (529), K296Q (409), and K286E (382) enzymes. As measured at pH 7.9, the Km value for NADPH was much greater for K286E (280 microM) than for WT (9.8 microM), K286R (30 microM), or K286Q (66 microM) enzymes. The efficiencies (kcat/Km) of the WT and K286R mutant were similar (1.2 x 10(3) min-1 microM-1 and 1.0 x 10(3) min-1 microM-1, respectively) while those of K286Q (0.30 x 10(3) min-1 microM-1) and K286E (0.07 x 10(3) min-1 microM-1) were greatly reduced. The decreased efficiency of the K286E mutant results from the increase in Km-NADPH, consistent with a role for a basic residue at position 286 which enhances the binding of NADPH. Plots of Vmax vs pH showed the pH optima to be 8.1-8.3 for all enzymes at saturating NADPH concentrations. A 40-fold increase in Km-NADPH for K286E was observed as the pH increased from 5.98 to 8.08, from which a unique pKe of 6.5 was calculated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
1. Juvenile hormone esterase (JHE) is a serine hydrolase selective for hydrolysis of the conjugated methyl esters of insect juvenile hormones. 2. We have investigated the mechanism of catalytic action of this enzyme by site-directed mutagenesis of the cloned enzyme and expression of the mutants in a baculovirus system. 3. A series of individual mutations of JHE were made to residues possibly involved in catalysis of juvenile hormones, and which are highly conserved in both esterases and lipases. 4. Mutation of the serine residue at position 201 to glycine (S201G), or aspartate 173 to asparagine (D173N), or histidine 446 to lysine (H446K), removed all detectable activity and these mutagenized enzymes were determined to be at least 10(6)-fold less active than wild type JHE. 5. Mutation of arginine 47 to histidine (R47H) decreased but did not abolish activity, with Km essentially unchanged at 66 nM for R47H compared to 34 nM for wild type JHE. 6. The kcat for R47H was decreased from 103 min-1 for wild type JHE to 1.9 min-1. 7. In addition, glutamate residue 332 was altered to glutamine (E332Q) and expressed in an Escherichia coli system. 8. This mutation was also found to remove all detectable activity. 9. From the results presented in this study and by comparison of JHE to other serine esterases and lipases, we predict that JHE possesses a Ser201-His446-Glu332 catalytic triad. 10. In addition, aspartate 173 and arginine 47 are essential for the efficient functioning of JHE.  相似文献   

12.
PsaE is a small basic subunit located on the stromal (cytoplasmic) side of photosystem I. In cyanobacteria, this subunit participates in cyclic electron transport and modulates the interactions of the complex with soluble ferredoxin. The PsaE protein isolated from the cyanobacterium Synechococcus sp. strain PCC 7002 adopts the beta topology of an SH3 domain, with five beta strands (betaA through betaE) and a turn of 3(10) helix between strands betaD and betaE [Falzone, C. J., Kao, Y.-H., Zhao, J., Bryant, D. A., and Lecomte, J. T. J. (1994) Biochemistry 33, 6052-6062]. The primary structure of the PsaE protein is strongly conserved across all oxygen-evolving photosynthetic organisms. However, variability in loop lengths, as well as N- or C-terminal extensions, suggests that the structure of a second representative PsaE subunit would be useful to characterize the interactions among photosystem I polypeptides. In this work, the solution structure of PsaE from the cyanobacterium Nostoc sp. strain PCC 8009 was determined by NMR methods. Compared to PsaE from Synechococcus sp. strain PCC 7002, this PsaE has a seven-residue deletion in the loop connecting strands betaC and betaD, and an eight-residue C-terminal extension. Angular and distance restraints derived from homonuclear and heteronuclear NMR experiments were used to calculate structures by a distance-geometry/simulated-annealing protocol. A family of 20 structures (rmsd of 0.24 A in the regular secondary structure) is presented. Differences between the two cyanobacterial proteins are mostly confined to the CD loop region; the C-terminal extension is disordered. The thermodynamic stability of Nostoc sp. strain PCC 8009 PsaE toward urea denaturation was measured by circular dichroism and fluorescence spectroscopy, and thermal denaturation was monitored by UV absorption spectroscopy. Chemical and thermal denaturation curves are modeled satisfactorily with two-state processes. The DeltaG degrees of unfolding at room temperature is 12.4 +/- 0.3 kJ mol(-1) (pH 5), and the thermal transition midpoint is 59 +/- 1 degrees C (pH 7). Interactions with other proteins in the photosystem I complex may aid in maintaining PsaE in its native state under physiological conditions.  相似文献   

13.
A covalent stoichiometric complex between photosystem I (PSI) and ferredoxin from the cyanobacterium Synechocystis sp. PCC 6803 was generated by chemical cross-linking. The photoreduction of ferredoxin, studied by laser flash absorption spectroscopy between 460 and 600 nm, is a fast process in 60% of the covalent complexes, which exhibit spectral and kinetic properties very similar to those observed with the free partners. Two major phases with t(1/2) <1 micros and approximately 10-14 micros are observed at two different pH values (5.8 and 8.0). The remaining complexes do not undergo fast ferredoxin reduction and 20-25% of the complexes are still able to reduce free ferredoxin or flavodoxin efficiently, thus indicating that ferredoxin is not bound properly in this proportion of covalent complexes. The docking site of ferredoxin on PSI was determined by electron microscopy in combination with image analysis. Ferredoxin binds to the cytoplasmic side of PSI, with its mass center 77 angstroms distant from the center of the trimer and in close contact with a ridge formed by the subunits PsaC, PsaD and PsaE. This docking site corresponds to a close proximity between the [2Fe- 2S] center of ferredoxin and the terminal [4Fe-4S] acceptor FII of PSI and is very similar in position to the docking site of flavodoxin, an alternative electron acceptor of PSI.  相似文献   

14.
In the heme-based oxygen sensor protein FixL, conformational changes induced by oxygen binding to the heme sensor domain regulate the activity of a neighboring histidine kinase, eventually restricting expression of specific genes to hypoxic conditions. The conserved arginine 220 residue is suggested to play a key role in the signal transduction mechanism. To obtain detailed insights into the role of this residue, we replaced Arg(220) by histidine (R220H), glutamine (R220Q), glutamate (R220E), and isoleucine (R220I) in the heme domain FixLH from Bradyrhizobium japonicum. These mutations resulted in dramatic changes in the O(2) affinity with K(d) values in the order R220I < R220Q < wild type < R220H. For the R220H and R220Q mutants, residue 220 interacts with the bound O(2) or CO ligands, as seen by resonance Raman spectroscopy. For the oxy-adducts, this H-bond modifies the pi acidity of the O(2) ligand, and its strength is correlated with the back-bonding-sensitive nu(4) frequency, the k(off) value for O(2) dissociation, and heme core-size conformational changes. This effect is especially strong for the wild-type protein where Arg(220) is, in addition, positively charged. These observations strongly suggest that neither strong ligand fixation nor the displacement of residue 220 into the heme distal pocket are solely responsible for the reported heme conformational changes associated with kinase activity regulation, but that a significant decrease of the heme pi(*) electron density because of strong back-bonding toward the oxygen ligand also plays a key role.  相似文献   

15.
Yeast has at least three partially characterized aldehyde dehydrogenases. Previous studies by gene disrupted in our laboratory revealed that the Saccharomyces cerevisiae cytosol ALDH1 played an important role in ethanol metabolism as did the class 2 mitochondrial enzyme. To date, few mutagenesis studies have been performed with the yeast enzymes. An important human variant of ALDH is one found in Asian People. In it, the glutamate at position 487 is replaced by a lysine. This glutamate interacts with an arginine (475) that is located in the subunit that makes up the dimer pair in the tetrameric enzyme. Sequence alignment shows that these two residues are located at positions 492 and 480, respectively, in the yeast class 1 enzyme which shares just 45% sequence identity with the human enzymes. Mutating glutamate 492 to lysine produced an enzyme with altered kinetic properties when compared to the wild-type glutamate-enzyme. The K(m) for NADP of E492K increased to nearly 3600 microM compare to 40 microM for wild-type enzyme. The specific activity decreased more than 10-fold with respect to the recombinant wild-type yeast enzyme. Moreover, substituting a glutamine for a glutamate was not detrimental in that the E492Q had wild-type-like K(m) for NADP and V(max). These properties were similar to the changes found with the human class 2 E487K mutant form. Further, mutating arginine 480 to glutamine produced an enzyme that exhibited positive cooperativity in NADP binding. The K(m) for NADP increased 11-fold with a Hill coefficient of 1.6. The NADP-dependent activity of R480Q mutant was 60% of wild-type enzyme. Again, these results are very similar to what we recently showed to occur with the human enzyme [Biochemistry 39 (2000) 5295-5302]. These findings show that the even though the glutamate and arginine residues are not conserved, similar changes occur in both the human and the yeast enzyme when either is mutated.  相似文献   

16.
Three arginine residues (Arg-11, Arg-39, Arg-61) are found at the active site of 4-oxalocrotonate tautomerase in the X-ray structure of the affinity-labeled enzyme [Taylor, A. B., Czerwinski, R. M., Johnson, R. M., Jr., Whitman, C. P., and Hackert, M. L. (1998) Biochemistry 37, 14692-14700]. The catalytic roles of these arginines were examined by mutagenesis, kinetic, and heteronuclear NMR studies. With a 1,6-dicarboxylate substrate (2-hydroxymuconate), the R61A mutation showed no kinetic effects, while the R11A mutation decreased k(cat) 88-fold and increased K(m) 8.6-fold, suggesting both binding and catalytic roles for Arg-11. With a 1-monocarboxylate substrate (2-hydroxy-2,4-pentadienoate), no kinetic effects of the R11A mutation were found, indicating that Arg-11 interacts with the 6-carboxylate of the substrate. The stereoselectivity of the R11A-catalyzed protonation at C-5 of the dicarboxylate substrate decreased, while the stereoselectivity of protonation at C-3 of the monocarboxylate substrate increased in comparison with wild-type 4-OT, indicating the importance of Arg-11 in properly orienting the dicarboxylate substrate by interacting with the charged 6-carboxylate group. With 2-hydroxymuconate, the R39A and R39Q mutations decreased k(cat) by 125- and 389-fold and increased K(m) by 1.5- and 2.6-fold, respectively, suggesting a largely catalytic role for Arg-39. The activity of the R11A/R39A double mutant was at least 10(4)-fold lower than that of the wild-type enzyme, indicating approximate additivity of the effects of the two arginine mutants on k(cat). For both R11A and R39Q, 2D (1)H-(15)N HSQC and 3D (1)H-(15)N NOESY-HSQC spectra showed chemical shift changes mainly near the mutated residues, indicating otherwise intact protein structures. The changes in the R39Q mutant were mainly in the beta-hairpin from residues 50 to 57 which covers the active site. HSQC titration of R11A with the substrate analogue cis, cis-muconate yielded a K(d) of 22 mM, 37-fold greater than the K(d) found with wild-type 4-OT (0.6 mM). With the R39Q mutant, cis, cis-muconate showed negative cooperativity in active site binding with two K(d) values, 3.5 and 29 mM. This observation together with the low K(m) of 2-hydroxymuconate (0.47 mM) suggests that only the tight binding sites function catalytically in the R39Q mutant. The (15)Nepsilon resonances of all six Arg residues of 4-OT were assigned, and the assignments of Arg-11, -39, and -61 were confirmed by mutagenesis. The binding of cis,cis-muconate to wild-type 4-OT upshifts Arg-11 Nepsilon (by 0.05 ppm) and downshifts Arg-39 Nepsilon (by 1.19 ppm), indicating differing electronic delocalizations in the guanidinium groups. A mechanism is proposed in which Arg-11 interacts with the 6-carboxylate of the substrate to facilitate both substrate binding and catalysis and Arg-39 interacts with the 1-carboxylate and the 2-keto group of the substrate to promote carbonyl polarization and catalysis, while Pro-1 transfers protons from C-3 to C-5. This mechanism, together with the effects of mutations of catalytic residues on k(cat), provides a quantitative explanation of the 10(7)-fold catalytic power of 4-OT. Despite its presence in the active site in the crystal structure of the affinity-labeled enzyme, Arg-61 does not play a significant role in either substrate binding or catalysis.  相似文献   

17.
The reaction center of photosystem I (PSI) reduces soluble ferredoxin on the stromal side of the photosynthetic membranes of cyanobacteria and chloroplasts. The X-ray structure of PSI from the cyanobacterium Synechococcus elongatus has been recently established at a 2.5 A resolution [Nature 411 (2001) 909]. The kinetics of ferredoxin photoreduction has been studied in recent years in many mutants of the stromal subunits PsaC, PsaD and PsaE of PSI. We discuss the ferredoxin docking site of PSI using the X-ray structure and the effects brought by the PSI mutations to the ferredoxin affinity.  相似文献   

18.
In this study, we present the location of the ferredoxin-binding site in photosystem I from spinach. Image analysis of negatively stained two-dimensional crystals indicates that the addition of ferredoxin and chemical cross-linkers do not significantly alter the unit cell parameters (for untreated photosystem I, a = 26.4 nm, b = 27.6 nm, and gamma = 90 degrees, space group p22(1)2(1) and for ferredoxin cross-linked photosystem I, a = 26.2 nm, b = 27.2 nm, and gamma = 90 degrees, space group p22(1)2(1)). Fourier difference analysis reveals that ferredoxin is bound on top of the stromal ridge principally interacting with the extrinsic subunits PsaC and PsaE. This location would be accessible to the stroma, thereby promoting efficient electron transfer away from photosystem I. This observation is significantly different from that of the ferredoxin binding site proposed for cyanobacteria. A model for the binding of ferredoxin in vascular plants is proposed and is discussed relative to observations in cyanobacteria.  相似文献   

19.
Bicarbonate (HCO3-) causes a significant and reversible stimulation of anion-inhibited electron flow in photosystem II of higher plants and cyanobacteria. To test if selected arginine (Arg) residues are involved in the binding of HCO3-, we utilized oligonucleotide-directed mutagenesis to construct Synechocystis sp. PCC 6803 mutants carrying mutations in Arg residues in the D2 protein. Measurements of oxygen evolution showed that the D2 mutants R233Q (arginine-233----glutamine) and R251S (arginine-251----serine) were 10-fold more sensitive to formate than the wild type. The formate concentration giving half-maximal inhibition of the steady-state oxygen evolution rate was 48 mM, 4.5 mM and 4 mM for the wild type, R233Q and R251S, respectively. Measurements of oxygen evolution in single-turnover flashes confirm that the mutants are more sensitive to formate than the wild type. Measurements of chlorophyll a fluorescence decay kinetics after the second saturating actinic flash indicated that, after formate treatment, the halftime of QA- oxidation was decreased by approximately a factor of 2, 4 and 6 in the wild type, R251S and R233Q, respectively. The recombination rate between QA- and S2 was approx. 2-fold slower in R251S and R233Q than in the wild type. In the presence of 100 mM sodium formate, reactivation of the Hill reaction by bicarbonate showed that the wild type had an apparent Km for bicarbonate of 0.5 mM, while the Km values for R233Q and R251S were 1.4 and 1.5 mM, respectively. We suggest that Arg-233 and Arg-251 in the D2 polypeptide contribute to stabilization of HCO3- binding in Photosystem II.  相似文献   

20.
Legler PM  Massiah MA  Mildvan AS 《Biochemistry》2002,41(35):10834-10848
GDP-mannose mannosyl hydrolase (GDPMH) is an unusual Nudix family member, which catalyzes the hydrolysis of GDP-alpha-D-mannose to GDP and the beta-sugar by nucleophilic substitution at carbon rather than at phosphorus (Legler, P. M., Massiah, M. A., Bessman, M. J., and Mildvan, A. S. (2000) Biochemistry 39, 8603-8608). Using the structure and mechanism of MutT, the prototypical Nudix enzyme as a guide, we detected six catalytic residues of GDPMH, three of which were unique to GDPMH, by the kinetic and structural effects of site-specific mutations. Glu-70 (corresponding to Glu-57 in MutT) provides a ligand to the essential divalent cation on the basis of the effects of the E70Q mutation which decreased kcat 10(2.2)-fold, increased the dissociation constant of Mn2+ from the ternary E-Mn2+-GDP complex 3-fold, increased the K(m)Mg2+ 20-fold, and decreased the paramagnetic effect of Mn2+ on 1/T1 of water protons, indicating a change in the coordination sphere of Mn2+. In the E70Q mutant, Gln-70 was shown to be very near the active site metal ion by large paramagnetic effects of Mn2+ on its side chain -NH2 group. With wild-type GDPMH, the effect of pH on log(kcat/K(m)GDPmann) at 37 degrees C showed an ascending limb of unit slope, followed by a plateau yielding a pK(a) of 6.4, which increased to 6.7 +/- 0.1 in the pH dependence of log(kcat). The general base catalyst was identified as a neutral His residue by the DeltaH(ionization) = 7.0 +/- 0.7 kcal/mol, by the increase in pK(a) with ionic strength, and by mutation of each of the four histidine residues of GDPMH to Gln. Only the H124Q mutant showed the loss of the ascending limb in the pH versus log(kcat) rate profile, which was replaced by a weak dependence of rate on hydroxide concentration, as well as an overall 10(3.4)-fold decrease in kcat, indicating His-124 to be the general base, unlike MutT, which uses Glu-53 in this role. The H88Q mutant showed a 10(2.3)-fold decrease in kcat, a 4.4-fold increase in K(m)GDPmann, and no change in the pH versus log(kcat) rate profile, indicating an important but unidentified role of His-88 in catalysis. One and two-dimensional NMR studies permitted the sequence specific assignments of the imidazole HdeltaC, H(epsilon)C, N(delta), and N(epsilon) resonances of the four histidines and defined their protonation states. The pK(a) of His-124 (6.94 +/- 0.04) in the presence of saturating Mg2+ was comparable to the kinetically determined pK(a) at the same temperature (6.40 +/- 0.20). The other three histidines were neutral N(epsilon)H tautomers with pK(a) values below 5.5. Arg-52 and Arg-65 were identified as catalytic residues which interact electrostatically with the GDP leaving group by mutating these residues to Gln and Lys. The R52Q mutant decreased kcat 309-fold and increased K(m)GDPmann 40.6-fold, while the R52K mutant decreased kcat by only 12-fold and increased K(m)GDPmann 81-fold. The partial rescue of kcat, but not of K(m)GDPmann in the R52K mutant, suggests that Arg-52 is a bifunctional hydrogen bond donor to the GDP leaving group in the ground state and a monofunctional hydrogen bond donor in the transition state. Opposite behavior was found with the Arg-65 mutants, suggesting this residue to be a monofunctional hydrogen bond donor to the GDP leaving group in the ground state and a bifunctional hydrogen bond donor in the transition state. From these observations, a mechanism for GDPMH is proposed involving general base catalysis and electrostatic stabilization of the leaving group.  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