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.
In higher plants ferredoxin (Fd):NADP(+) oxidoreductase (FNR) and Fd are each distributed in photosynthetic and non-photosynthetic organs as distinct isoproteins. We have cloned cDNAs for leaf FNR (L-FNR I and L-FNR II) and root FNR (R-FNR) from maize (Zea mays L.), and produced recombinant L-FNR I and R-FNR to study their enzymatic functions through kinetic and Fd-binding analyses. The K(m) value obtained by assay for a diaphorase activity indicated that R-FNR had a 10-fold higher affinity for NADPH than L-FNR I. When we assayed for NADPH-cytochrome c reductase activity using maize photosynthetic Fd (Fd I) and non-photosynthetic Fd (Fd III), the R-FNR showed a marked difference in affinity between these two Fd isoproteins; the K(m) for Fd III was 3.0 microM and that for Fd I was 29 microM. Consistent with this, the dissociation constant for the R-FNR:Fd III complex was 10-fold smaller than that of the R-FNR:Fd I complex. This differential binding capacity was confirmed by an affinity chromatography of R-FNR on Fd-sepharose with stronger binding to Fd III. L-FNR I showed no such differential interaction with Fd I and Fd III. These data demonstrated that R-FNR has the ability to discriminate between these two types of Fds. We propose that the stronger interaction of R-FNR with Fd III is crucial for an efficient electron flux of NADPH-FNR-Fd cascade, thus supporting Fd-dependent metabolism in non-photosynthetic organs.  相似文献   

2.
3.
Ferredoxin (Fd) and Fd-NADP(+) reductase (FNR) are redox partners responsible for the conversion between NADP(+) and NADPH in the plastids of photosynthetic organisms. Introduction of specific disulfide bonds between Fd and FNR by engineering cysteines into the two proteins resulted in 13 different Fd-FNR cross-linked complexes displaying a broad range of activity to catalyze the NADPH-dependent cytochrome c reduction. This variability in activity was thought to be mainly due to different levels of intramolecular electron transfer activity between the FNR and Fd domains. Stopped-flow analysis revealed such differences in the rate of electron transfer from the FNR to Fd domains in some of the cross-linked complexes. A group of the cross-linked complexes with high cytochrome c reduction activity comparable to dissociable wild-type Fd/FNR was shown to assume a similar Fd-FNR interaction mode as in the native Fd:FNR complex by analyses of NMR chemical shift perturbation and absorption spectroscopy. However, the intermolecular electron transfer of these cross-linked complexes with two Fd-binding proteins, nitrite reductase and photosystem I, was largely inhibited, most probably due to steric hindrance by the FNR moiety linked near the redox center of the Fd domain. In contrast, another group of the cross-linked complexes with low cytochrome c reduction activity tends to mediate higher intermolecular electron transfer activity. Therefore, reciprocal relationship of intramolecular and intermolecular electron transfer abilities was conferred by the linkage of Fd and FNR, which may explain the physiological significance of the separate forms of Fd and FNR in chloroplasts.  相似文献   

4.
Plant-type ferredoxin (Fd), a [2Fe-2S] iron-sulfur protein, functions as an one-electron donor to Fd-NADP(+) reductase (FNR) or sulfite reductase (SiR), interacting electrostatically with them. In order to understand the protein-protein interaction between Fd and these two different enzymes, 10 acidic surface residues in maize Fd (isoform III), Asp-27, Glu-30, Asp-58, Asp-61, Asp-66/Asp-67, Glu-71/Glu-72, Asp-85, and Glu-93, were substituted with the corresponding amide residues by site-directed mutagenesis. The redox potentials of the mutated Fds were not markedly changed, except for E93Q, the redox potential of which was more positive by 67 mV than that of the wild type. Kinetic experiments showed that the mutations at Asp-66/Asp-67 and Glu-93 significantly affected electron transfer to the two enzymes. Interestingly, D66N/D67N was less efficient in the reaction with FNR than E93Q, whereas this relationship was reversed in the reaction with SiR. The static interaction of the mutant Fds with each the two enzymes was analyzed by gel filtration of a mixture of Fd and each enzyme, and by affinity chromatography on Fd-immobilized resins. The contributions of Asp-66/Asp-67 and Glu-93 were found to be most important for the binding to FNR and SiR, respectively, in accordance with the kinetic data. These results allowed us to map the acidic regions of Fd required for electron transfer and for binding to FNR and SiR and demonstrate that the interaction sites for the two enzymes are at least partly distinct.  相似文献   

5.
6.
7.
NADPH is an intermediate in the oxidation of organic compounds coupled to Fe(III) reduction in Geobacter species, but Fe(III) reduction with NADPH as the electron donor has not been studied in these organisms. Crude extracts of Geobacter sulfurreducens catalyzed the NADPH-dependent reduction of Fe(III)-nitrilotriacetic acid (NTA). The responsible enzyme, which was recovered in the soluble protein fraction, was purified to apparent homogeneity in a four-step procedure. Its specific activity for Fe(III) reduction was 65 micromol. min(-1). mg(-1). The soluble Fe(III) reductase was specific for NADPH and did not utilize NADH as an electron donor. Although the enzyme reduced several forms of Fe(III), Fe(III)-NTA was the preferred electron acceptor. The protein possessed methyl viologen:NADP(+) oxidoreductase activity and catalyzed the reduction of NADP(+) with reduced methyl viologen as electron donor at a rate of 385 U/mg. The enzyme consisted of two subunits with molecular masses of 87 and 78 kDa and had a native molecular mass of 320 kDa, as determined by gel filtration. The purified enzyme contained 28.9 mol of Fe, 17.4 mol of acid-labile sulfur, and 0.7 mol of flavin adenine dinucleotide per mol of protein. The genes encoding the two subunits were identified in the complete sequence of the G. sulfurreducens genome from the N-terminal amino acid sequences derived from the subunits of the purified protein. The sequences of the two subunits had about 30% amino acid identity to the respective subunits of the formate dehydrogenase from Moorella thermoacetica, but the soluble Fe(III) reductase did not possess formate dehydrogenase activity. This soluble Fe(III) reductase differs significantly from previously characterized dissimilatory and assimilatory Fe(III) reductases in its molecular composition and cofactor content.  相似文献   

8.
Maize (Zea mays L.) has two types of ferredoxin (Fd) differentially expressed in photosynthetic and nonphotosynthetic organs. A cDNA fragment encoding the mature polypeptide of Fd III, an Fd isoprotein of the nonphotosynthetic type, was expressed in Escherichia coli, and the Fd was synthesized as a holo-form assembled with the [2Fe-2S] cluster, which was completely identical with authentic Fd III prepared from maize roots. This expression system made it possible to prepare Fd present at fairly low levels in plants in amounts sufficient for functional and structural studies. Comparison of electron transfer activity of Fd III with that of Fd I, an Fd isoprotein of the photosynthetic type, showed that Fd III was superior as an electron acceptor from NADPH, and Fd I was superior as an electron donor for NADP+, in reactions catalyzed by Fd-NADP+ reductase from maize leaf. The circular dichronism spectra of the two Fds also indicated a subtle difference in the geometry of their iron-sulfur clusters. These results are consistent with the view that photosynthetic and nonphotosynthetic Fds may be functionally differentiated. An artificial chimeric Fd, Fd III/Fd I, whose amino-terminal and carboxylterminal halves are derived from the corresponding regions of Fd III and Fd I, respectively, showed an activity and CD spectrum significantly similar to those of Fd I. This suggests that 18 amino acid substitutions between Fd III and Fd III/Fd I alter the properties of Fd III so that they resemble those of Fd I.  相似文献   

9.
Wright DP  Huppe HC  Turpin DH 《Plant physiology》1997,114(4):1413-1419
Pyridine nucleotide pools were measured in intact plastids from roots of barley (Hordeum vulgare L.) during the onset of NO2- assimilation and compared with the in vitro effect of the NADPH/NADP ratio on the activity of plastidic glucose-6-phosphate dehydrogenase (G6PDH, EC 1.1.1.49) from N-sufficient or N-starved roots. The NADPH/NADP ratio increased from 0.9 to 2.0 when 10 mM glucose-6-phosphate was supplied to intact plastids. The subsequent addition of 1 mM NaNO2 caused a rapid decline in this ratio to 1.5. In vitro, a ratio of 1.5 inactivated barley root plastid G6PDH by approximately 50%, suggesting that G6PDH could remain active during NO2- assimilation even at the high NADPH/NADP ratios that would favor a reduction of ferredoxin, the electron donor of NO2- reductase. Root plastid G6PDH was sensitive to reductive inhibition by dithiothreitol (DTT), but even at 50 mM DTT the enzyme remained more than 35% active. In root plastids from barley starved of N for 3 d, G6PDH had a substantially reduced specific activity, had a lower Km for NADP, and was less inhibited by DTT than the enzyme from N-sufficient root plastids, indicating that there was some effect of N starvation on the G6PDH activity in barley root plastids.  相似文献   

10.
Yan X  Khan S  Hase T  Emes MJ  Bowsher CG 《FEBS letters》2006,580(27):6509-6512
The photosynthetic proteins RuBiSCO, ferredoxin I and ferredoxin NADP(+)-oxidoreductase (pFNR) were efficiently imported into isolated pea chloroplasts but not into pea root plastids. By contrast non-photosynthetic ferredoxin III and heterotrophic FNR (hFNR) were efficiently imported into both isolated chloroplasts and root plastids. Chimeric ferredoxin I/III (transit peptide of ferredoxin I attached to the mature region of ferredoxin III) only imported into chloroplasts. Ferredoxin III/I (transit peptide of ferredoxin III attached to the mature region of ferredoxin I) imported into both chloroplasts and root plastids. This suggests that import depends on specific interactions between the transit peptide and the translocon apparatus.  相似文献   

11.
Rapid reaction studies presented herein show that ferredoxin:NADP+ oxidoreductase (FNR, EC 1.18.1.2) catalyzes electron transfer from spinach ferredoxin (Fd) to NADP+ via a ternary complex, Fd X FNR X NADP+. In the absence of NADP+, reduction of ferredoxin:NADP+ reductase by Fd was much slower than the catalytic rate: 37-80 s-1 versus at least 445 e-s-1; dissociation of oxidized spinach ferredoxin (Fdox) from one-electron reduced ferredoxin:NADP+ reductase (FNRsq) limited the reduction of FNR. This confirms the steady-state kinetic analysis of Masaki et al. (Masaki, R., Yoshikaya, S., and Matsubara, H. (1982) Biochim. Biophys. Acta 700, 101-109). Occupation of the NADP+ binding site of FNR by NADP+ or by 2',5'-ADP (a nonreducible NADP+ analogue) greatly increased the rate of electron transfer from Fd to FNR, releiving inhibition by Fdox. NADP+ (and 2',5'-ADP) probably facilitate the dissociation of Fdox; equilibrium studies have shown that nucleotide binding decreases the association of Fd with FNR (Batie, C. J. (1983) Ph.D. dissertation, Duke University; Batie, C. J., and Kamin, H. (1982) in Flavins and Flavoproteins VII (Massey, V., and Williams, C. H., Jr., eds) pp. 679-683, Elsevier, New York; Batie, C.J., and Kamin, H. (1982) Fed. Proc. 41, 888; and Batie, C.J., and Kamin, H. (1984) J. Biol. Chem. 259, 8832-8839). Premixing Fd with FNR was found to inhibit the reaction of the flavoprotein with NADP+ and with NADPH; thus, substrate binding may be ordered, NADP+ first, then Fd. FNRred and NADP+ very rapidly formed an FNRred X NADP+ complex with flavin to nicotinamide charge transfer bands. The Fdred X NADP+ complex then relaxed to an equilibrium species; the spectrum indicated a predominance of FNRox X NADPH charge-transfer complex. However, charge-transfer species were not observed during turnover; thus, their participation in catalysis of electron transfer from Fd to NADP+ remains uncertain. The catalytic rate of Fd to NADP+ electron transfer, as well as the rates of electron transfer from Fd to FNR, and from FNR to NADP+ were decreased when the reactants were in D2O; diaphorase activity was unaffected by solvent. On the basis of the data presented, a scheme for the catalytic mechanism of catalysis by FNR is presented.  相似文献   

12.
Ferredoxin:NADP+ oxidoreductase (ferredoxin: NADP+ reductase, EC 1.18.1.2) was shown to form a ternary complex with its substrates ferredoxin (Fd) and NADP(H), but the ternary complex was less stable than the separate binary complexes. Kd for oxidized binary Fd-ferredoxin NADP+ reductase complex was less than 50 nM; Kd(Fd) increased with NADP+ concentration, approaching 0.5-0.6 microM when the flavoprotein was saturated with NADP+ K(NADP+) also increased from about 14 microM to about 310 microM, on addition of excess Fd. The changes in Kd were consistent with negative cooperativity between the associations of Fd and NADP+ and with our unpublished observations which suggest that product dissociation is rate-limiting in the reaction mechanism. Similar interference in binding was observed in more reduced states; NADPH released much ferredoxin:NADP+ reductase from Fd-Sepharose whether the proteins were initially oxidized or reduced. Complexation between Fd and ferredoxin: NADP+ reductase was found to shield each center from paramagnetic probes; charge specificity suggested that the active sites of Fd and ferredoxin:NADP+ reductase were, respectively, negatively and positively charged.  相似文献   

13.
《Biophysical journal》2022,121(10):1799-1812
Precursor molecules for biomass incorporation must be imported into cells and made available to the molecular machines that build the cell. Sulfur-containing macromolecules require that sulfur be in its S2? oxidation state before assimilation into amino acids, cofactors, and vitamins that are essential to organisms throughout the biosphere. In α-proteobacteria, NADPH-dependent assimilatory sulfite reductase (SiR) performs the final six-electron reduction of sulfur. SiR is a dodecameric oxidoreductase composed of an octameric flavoprotein reductase (SiRFP) and four hemoprotein metalloenzyme oxidases (SiRHPs). SiR performs the electron transfer reduction reaction to produce sulfide from sulfite through coordinated domain movements and subunit interactions without release of partially reduced intermediates. Efforts to understand the electron transfer mechanism responsible for SiR’s efficiency are confounded by structural heterogeneity arising from intrinsically disordered regions throughout its complex, including the flexible linker joining SiRFP’s flavin-binding domains. As a result, high-resolution structures of SiR dodecamer and its subcomplexes are unknown, leaving a gap in the fundamental understanding of how SiR performs this uniquely large-volume electron transfer reaction. Here, we use deuterium labeling, in vitro reconstitution, analytical ultracentrifugation (AUC), small-angle neutron scattering (SANS), and neutron contrast variation (NCV) to observe the relative subunit positions within SiR’s higher-order assembly. AUC and SANS reveal SiR to be a flexible dodecamer and confirm the mismatched SiRFP and SiRHP subunit stoichiometry. NCV shows that the complex is asymmetric, with SiRHP on the periphery of the complex and the centers of mass between SiRFP and SiRHP components over 100 Å apart. SiRFP undergoes compaction upon assembly into SiR’s dodecamer and SiRHP adopts multiple positions in the complex. The resulting map of SiR’s higher-order structure supports a cis/trans mechanism for electron transfer between domains of reductase subunits as well as between tightly bound or transiently interacting reductase and oxidase subunits.  相似文献   

14.
The thyroid plasma membrane contains a Ca2(+)-regulated NADPH-dependent H2O2 generating system which provides H2O2 for the thyroid peroxidase-catalyzed biosynthesis of thyroid hormones. The plasma membrane fraction contains a Ca2(+)-independent cytochrome c reductase activity which is not inhibited by superoxide dismutase. But it is not known whether H2O2 is produced directly from molecular oxygen (O2) or formed via dismutation of super-oxide anion (O2-). Indirect evidence from electron scavenger studies indicate that the H2O2 generating system does not liberate O2-, but studies using the modified peroxidase, diacetyldeuteroheme horseradish peroxidase, to detect O2- indicate that H2O2 is provided via the dismutation of O2-. The present results provide indirect evidence that the cytochrome c reductase activity is not a component of the NADPH-dependent H2O2 generator, since it was removed by washing the plasma membranes with 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonic acid without affecting H2O2 generation. Spectral studies with diacetyldeuteroheme-substituted horseradish peroxidase showed that the thyroid NADPH-dependent H2O2 generator does not catalyze superoxide anion formation. The O2- adduct compound (compound III) was formed but was completely inhibited by catalase, indicating that the initial product was H2O2. The rate of NADPH oxidation also increased in the presence of diacetylheme peroxidase. This increase was blocked by catalase and was greatly enhanced by superoxide dismutase. The O2- adduct compound (compound III) was produced in the presence of NADPH when glucose-glucose oxidase (which does not produce O2-) was used as the H2O2 generator. NADPH oxidation occurred simultaneously and was enhanced by superoxide dismutase. We conclude that O2- formation occurs in the presence of an H2O2 generator, diacetylheme peroxidase and NADPH, but that it is not the primary product of the H2O2 generator. We suggest that O2- formation results from oxidation of NADPH, catalyzed by the diacetylheme peroxidase compound I, producing NADP degree, which in turn reacts with O2 to give O2-.  相似文献   

15.
Addition of NADP(+) to thylakoid membranes or isolated photosystem I (PSI) submembrane fractions quenched chlorophyll fluorescence by up to 40% at low or room temperature. This quenching was reversed by NADPH. Similar quenching was also observed with the addition of heparin or thenoyltrifluoroacetone (TTFA), inhibitors that bind ferredoxin:NADP(+) reductase (FNR) and prevent reduction of NADP(+). The NADP(+)-induced quenching coincided with a reversible conformational change of the secondary protein structure in the PSI submembrane fractions where 20% of the alpha-helix conformations were transformed mainly into beta-sheet-like structures. Further, P700 photooxidation was retarded due to this conformational change, and about 25% of the centers could not be photooxidized, these changes being also reversible with addition of NADPH. The above modifications in the presence of NADP(+) also increased photodamage processes under strong illumination, and NADPH protected it. Conformational modification of FNR upon binding of NADP(+) or NADPH is proposed to trigger the macromolecular changes in a larger part of the protein complex of PSI. The conformational changes must increase the intermolecular distances and change the mutual orientation between the various cofactors in the PSI complex. This new control mechanism of energy dissipation and photochemical activity by NADP(+)/NADPH is proposed to increase the turnover rate of PSI under conditions when both linear and cyclic electron transport activities must be supported.  相似文献   

16.
Ferredoxin-NAD(P)(+) reductase [EC 1.18.1.3, 1.18.1.2] was isolated from the green sulfur bacterium Chlorobium tepidum and purified to homogeneity. The molecular mass of the subunit is 42 kDa, as deduced by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). The molecular mass of the native enzyme is approximately 90 kDa, estimated by gel-permeation chromatography, and is thus a homodimer. The enzyme contains one FAD per subunit and has absorption maxima at about 272, 385, and 466 nm. In the presence of ferredoxin (Fd) and reaction center (RC) complex from C. tepidum, it efficiently catalyzes photoreduction of both NADP(+) and NAD(+). When concentrations of NADP(+) exceeded 10 microM, NADP(+) photoreduction rates decreased with increased concentration. The inhibition by high concentrations of substrate was not observed with NAD(+). It also reduces 2,6-dichlorophenol-indophenol (DPIP) and molecular oxygen with either NADPH or NADH as efficient electron donors. It showed NADPH diaphorase activity about two times higher than NADH diaphorase activity in DPIP reduction assays at NAD(P)H concentrations less than 0.1 mM. At 0.5 mM NAD(P)H, the two activities were about the same, and at 1 mM, the former activity was slightly lower than the latter.  相似文献   

17.
Ferredoxin and the flavoprotein, ferredoxin: NADP reductase, have been covalently linked by incubation in the presence of a water soluble carbodiimide. The cross-linking reaction yields an adduct having a 1:1 stoichiometry. The adduct has depressed levels of diaphorase and NADPH oxidase activity and is inactive in reduction of cytochrome c using NADPH as an electron donor. Thus, although similar to an adduct described by Zanetti and coworkers [J Biol Chem 259: 6153–6157 (1984)] in its stoichiometry, the adduct described herein has significantly different enzymatic properties. It is suggested that this may be a reflection of differences in the interaction between the two proteins resulting from differences in experimental conditions in which the two adducts were prepared.Abbreviations Fd ferredoxin - Fp ferredoxin: NADP reductase - Fd Fp covalently linked Fd-Fp adduct - Fd:Fp noncovalently linked complex between Fd and Fp - EDC 1-ethyl-3-(dimethylaminopropyl) carbodiimide - Tris tris-hydroxymethylaminomethane - MOPS 3-(N-morpholino)propane sulfonic acid - DCIP 2,6-dichloropenolindophenol  相似文献   

18.
Previous studies have shown that the interaction of P450 reductase with bound NADP(H) is essential to ensure fast electron transfer through the two flavin cofactors. In this study we investigated in detail the interaction of the house fly flavoprotein with NADP(H) and a number of nucleotide analogues. 1,4,5,6-Tetrahydro-NADP, an analogue of NADPH, was used to characterize the interaction of P450 reductase with the reduced nucleotide. This analogue is inactive as electron donor, but its binding affinity and rate constant of release are very close to those for NADPH. The 2'-phosphate contributes about 5 kcal/mol of the binding energy of NADP(H). Oxidized nicotinamide does not interact with the oxidized flavoprotein, while reduced nicotinamide contributes 1.3 kcal/mol of the binding energy. Oxidized P450 reductase binds NADPH with a K(d) of 0.3 microM, while the affinity of the reduced enzyme is considerably lower, K(d) = 1.9 microM. P450 reductase catalyzes a transhydrogenase reaction between NADPH and oxidized nucleotides, such as thionicotinamide-NADP(+), acetylpyridine-NADP(+), or [(3)H]NADP(+). The reverse reaction, reduction of [(3)H]NADP(+) by the reduced analogues, is also catalyzed by P450 reductase. We define the mechanism of the transhydrogenase reaction as follows: NADPH binding, hydride ion transfer, and release of the NADP(+) formed. An NADP(+) or its analogue binds to the two-electron-reduced flavoprotein, and the electron-transfer steps reverse to transfer hydride ion to the oxidized nucleotide, which is released. Measurements of the flavin semiquinone content, rate constant for NADPH release, and transhydrogenase turnover rates allowed us to estimate the steady-state distribution of P450 reductase species during catalysis, and to calculate equilibrium constants for the interconversion of catalytic intermediates. Our results demonstrate that equilibrium redox potentials of the flavin cofactors are not the sole factor governing rapid electron transfer during catalysis, but conformational changes must be considered to understand P450 reductase catalysis.  相似文献   

19.
In the C(4) plant maize (Zea mays L.), two ferredoxin isoproteins, Fd I and Fd II, are expressed specifically in mesophyll and bundle-sheath cells, respectively. cDNAs for these ferredoxins were introduced separately into the cyanobacterium Plectonema boryanum with a disrupted endogenous ferredoxin gene, yielding TM202 and KM2-9 strains expressing Fd I and Fd II. The growth of TM202 was retarded under high light (130 micromol/m(2)/s), whereas KM2-9 grew at a normal rate but exhibited a nitrogen-deficient phenotype. Measurement of photosynthetic O(2) evolution revealed that the reducing power was not efficiently partitioned into nitrogen assimilation in KM2-9. After starvation of the cells in darkness, the P700 oxidation level under far-red illumination increased significantly in TM202. However, it remained low in KM2-9, indicating an active cyclic electron flow. In accordance with this, the cellular ratio of ATP/ADP increased and that of NADPH/NADP(+) decreased in KM2-9 as compared with TM202. These results demonstrated that the two cell type-specific ferredoxins differentially modulate electron flow around photosystem I.  相似文献   

20.
Zeng J  Wang M  Zhang X  Wang Y  Ai C  Liu J  Qiu G 《Biotechnology letters》2008,30(7):1239-1244
Sulfite reductase (SiR) is a large and soluble enzyme which catalyzes the transfer of six electrons from NADPH to sulfite to produce sulfide. The sulfite reductase flavoprotein (SiR-FP) contains both FAD and FMN, and the sulfite reductase hemoprotein (SiR-HP) contains an iron-sulfur cluster coupled to a siroheme. The enzyme is arranged so that the redox cofactors in the FAD-FMN-Fe(4)S(4)-Heme sequence make an electron pathway between NADPH and sulfite. Here we report the cloning, expression, and characterization of the SiR-HP of the sulfite reductase from Acidithiobacillus ferrooxidans. The purified SiR-HP contained a [Fe(4)S(4)] cluster. Site-directed mutagenesis results revealed that Cys427, Cys433, Cys472 and Cys476 were in ligating with the [Fe(4)S(4)] cluster of the protein.  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