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.
Transhydrogenase couples the redox reaction between NADH and NADP+ to proton translocation across a membrane. The enzyme comprises three components; dI binds NAD(H), dIII binds NADP(H), and dII spans the membrane. The 1,4,5,6-tetrahydro analogue of NADH (designated H2NADH) bound to isolated dI from Rhodospirillum rubrum transhydrogenase with similar affinity to the physiological nucleotide. Binding of either NADH or H2NADH led to closure of the dI mobile loop. The 1,4,5,6-tetrahydro analogue of NADPH (H2NADPH) bound very tightly to isolated R. rubrum dIII, but the rate constant for dissociation was greater than that for NADPH. The replacement of NADP+ on dIII either with H2NADPH or with NADPH caused a similar set of chemical shift alterations, signifying an equivalent conformational change. Despite similar binding properties to the natural nucleotides, neither H2NADH nor H2NADPH could serve as a hydride donor in transhydrogenation reactions. Mixtures of dI and dIII form dI2dIII1 complexes. The nucleotide charge distribution of complexes loaded either with H2NADH and NADP+ or with NAD+ and H2NADPH should more closely mimic the ground states for forward and reverse hydride transfer, respectively, than previously studied dead-end species. Crystal structures of such complexes at 2.6 and 2.3 A resolution are described. A transition state for hydride transfer between dihydronicotinamide and nicotinamide derivatives determined in ab initio quantum mechanical calculations resembles the organization of nucleotides in the transhydrogenase active site in the crystal structure. Molecular dynamics simulations of the enzyme indicate that the (dihydro)nicotinamide rings remain close to a ground state for hydride transfer throughout a 1.4 ns trajectory.  相似文献   

2.
The reaction of pentaerythritol tetranitrate reductase with reducing and oxidizing substrates has been studied by stopped-flow spectrophotometry, redox potentiometry, and X-ray crystallography. We show in the reductive half-reaction of pentaerythritol tetranitrate (PETN) reductase that NADPH binds to form an enzyme-NADPH charge transfer intermediate prior to hydride transfer from the nicotinamide coenzyme to FMN. In the oxidative half-reaction, the two-electron-reduced enzyme reacts with several substrates including nitroester explosives (glycerol trinitrate and PETN), nitroaromatic explosives (trinitrotoluene (TNT) and picric acid), and alpha,beta-unsaturated carbonyl compounds (2-cyclohexenone). Oxidation of the flavin by the nitroaromatic substrate TNT is kinetically indistinguishable from formation of its hydride-Meisenheimer complex, consistent with a mechanism involving direct nucleophilic attack by hydride from the flavin N5 atom at the electron-deficient aromatic nucleus of the substrate. The crystal structures of complexes of the oxidized enzyme bound to picric acid and TNT are consistent with direct hydride transfer from the reduced flavin to nitroaromatic substrates. The mode of binding the inhibitor 2,4-dinitrophenol (2,4-DNP) is similar to that observed with picric acid and TNT. In this position, however, the aromatic nucleus is not activated for hydride transfer from the flavin N5 atom, thus accounting for the lack of reactivity with 2,4-DNP. Our work with PETN reductase establishes further a close relationship to the Old Yellow Enzyme family of proteins but at the same time highlights important differences compared with the reactivity of Old Yellow Enzyme. Our studies provide a structural and mechanistic rationale for the ability of PETN reductase to react with the nitroaromatic explosive compounds TNT and picric acid and for the inhibition of enzyme activity with 2,4-DNP.  相似文献   

3.
Rubach JK  Plapp BV 《Biochemistry》2003,42(10):2907-2915
Amino acid residues Thr-178, Val-203, and Val-292, which interact with the nicotinamide ring of the coenzyme bound to alcohol dehydrogenase (ADH), may facilitate hydride transfer and hydrogen tunneling by orientation and dynamic effects. The T178S, T178V, V203A, V292A, V292S, and V292T substitutions significantly alter the steady state and transient kinetics of the enzyme. The V292A, V292S, and V292T enzymes have decreased affinity for coenzyme (NAD+ by 30-50-fold and NADH by 35-75-fold) as compared to the wild-type enzyme. The substitutions in the nicotinamide binding site decrease the rate constant of hydride transfer for benzyl alcohol oxidation by 3-fold (for V292T ADH) to 16-fold (for V203A ADH). The modest effects suggest that catalysis does not depend critically on individual residues and that several residues in the nicotinamide binding site contribute to catalysis. The structures of the V292T ADH-NAD+-pyrazole and wild-type ADH-NAD+-4-iodopyrazole ternary complexes are very similar. Only subtle changes in the V292T enzyme cause the large changes in coenzyme binding and the small change in hydride transfer. In these complexes, one pyrazole nitrogen binds to the catalytic zinc, and the other nitrogen forms a partial covalent bond with C4 of the nicotinamide ring, which adopts a boat conformation that is postulated to be relevant for hydride transfer. The results provide an experimental basis for evaluating the contributions of dynamics to hydride transfer.  相似文献   

4.
The flavoenzyme ferredoxin-NADP+ reductase (FNR) catalyzes the production of NADPH during photosynthesis. Whereas the structures of FNRs from spinach leaf and a cyanobacterium as well as many of their homologs have been solved, none of these studies has yielded a productive geometry of the flavin-nicotinamide interaction. Here, we show that this failure occurs because nicotinamide binding to wild type FNR involves the energetically unfavorable displacement of the C-terminal Tyr side chain. We used mutants of this residue (Tyr 308) of pea FNR to obtain the structures of productive NADP+ and NADPH complexes. These structures reveal a unique NADP+ binding mode in which the nicotinamide ring is not parallel to the flavin isoalloxazine ring, but lies against it at an angle of approximately 30 degrees, with the C4 atom 3 A from the flavin N5 atom.  相似文献   

5.
G Blankenhorn 《Biochemistry》1975,14(14):3172-3176
The reaction of N-methyldihydronicotinamide (NMNH) with flavine analogs saturates at high dihydronicotinamide concentrations. Complex formation between the reactants depends mainly on steric but not on electronic factors. Thus flavine analogs that differ up to 243 mV in their oxidation-reduction potential vary only between 0.09 and 0.17 M in Kd. When the flavine plane becomes blocked by bulky substituents, however, complex stability decreases by more than an order of magnitude. NMNH-flavine complexes show long wave optical absorption. The energy of the long wave transition decreases with increasing oxidation-reduction potential of the flavine as expected for charge transfer complexes. The first-order rate constants of flavine-dependent dihydronicotinamide dehydrogenation increase with increasing oxidation-reduction potential of the flavine but they are almost independent of Kd. The reaction is not subject to general acid-base catalysis. Thus flavine-dependent dihydronicotinamide dehydrogenation may be interpreted to proceed via a charge transfer complex between oxidized flavine and reduced nicotinamide. In the rate-limiting conversion of the charge transfer complex into products hydrogen is transferred directly, the rate being governed by the difference in oxidation-reduction potential between flavine and dihydronicotinamide. An alternative mechanism where the observed charge transfer complex is not on the reaction pathway appears to be improbable but cannot be eliminated.  相似文献   

6.
Transhydrogenase couples proton translocation across a membrane to hydride transfer between NADH and NADP+. Previous x-ray structures of complexes of the nucleotide-binding components of transhydrogenase ("dI2dIII1" complexes) indicate that the dihydronicotinamide ring of NADH can move from a distal position relative to the nicotinamide ring of NADP+ to a proximal position. The movement might be responsible for gating hydride transfer during proton translocation. We have mutated three invariant amino acids, Arg-127, Asp-135, and Ser-138, in the NAD(H)-binding site of Rhodospirillum rubrum transhydrogenase. In each mutant, turnover by the intact enzyme is strongly inhibited. Stopped-flow experiments using dI2dIII1 complexes show that inhibition results from a block in the steps associated with hydride transfer. Mutation of Asp-135 and Ser-138 had no effect on the binding affinity of either NAD+ or NADH, but mutation of Arg-127 led to much weaker binding of NADH and slightly weaker binding of NAD+. X-ray structures of dI2dIII1 complexes carrying the mutations showed that their effects were restricted to the locality of the bound NAD(H). The results are consistent with the suggestion that in wild-type protein movement of the Arg-127 side chain, and its hydrogen bonding to Asp-135 and Ser-138, stabilizes the dihydronicotinamide of NADH in the proximal position for hydride transfer.  相似文献   

7.
Ferredoxin-NADP(H) reductases catalyse the reversible hydride/electron exchange between NADP(H) and ferredoxin/flavodoxin, comprising a structurally defined family of flavoenzymes with two distinct subclasses. Those present in Gram-negative bacteria (FPRs) display turnover numbers of 1-5 s(-1) while the homologues of cyanobacteria and plants (FNRs) developed a 100-fold activity increase. We investigated nucleotide interactions and hydride transfer in Rhodobacter capsulatus FPR comparing them to those reported for FNRs. NADP(H) binding proceeds as in FNRs with stacking of the nicotinamide on the flavin, which resulted in formation of charge-transfer complexes prior to hydride exchange. The affinity of FPR for both NADP(H) and 2'-P-AMP was 100-fold lower than that of FNRs. The crystal structure of FPR in complex with 2'-P-AMP and NADP(+) allowed modelling of the adenosine ring system bound to the protein, whereas the nicotinamide portion was either not visible or protruding toward solvent in different obtained crystals. Stabilising contacts with the active site residues are different in the two reductase classes. We conclude that evolution to higher activities in FNRs was partially favoured by modification of NADP(H) binding in the initial complexes through changes in the active site residues involved in stabilisation of the adenosine portion of the nucleotide and in the mobile C-terminus of FPR.  相似文献   

8.
Ferredoxin-NADP+ reductase (FNR) catalyzes the reduction of NADP+ to NADPH in an overall reversible reaction, showing some differences in the mechanisms between cyanobacterial and higher plant FNRs. During hydride transfer it is proposed that the FNR C-terminal Tyr is displaced by the nicotinamide. Thus, this C-terminal Tyr might be involved not only in modulating the flavin redox properties, as already shown, but also in nicotinamide binding and hydride transfer. FNR variants from the cyanobacterium Anabaena in which the C-terminal Tyr has been replaced by Trp, Phe, or Ser have been produced. All FNR variants show enhanced NADP+ and NAD+ binding, especially Tyr303Ser, which correlates with a noticeable improvement of NADH-dependent reactions. Nevertheless, the Tyr303Ser variant shows a decrease in the steady-state kcat value with NADPH. Fast kinetic analysis of the hydride transfer shows that the low efficiency observed for this mutant FNR under steady-state conditions is not due to a lack of catalytic ability but rather to the strong enzyme-coenzyme interaction. Three-dimensional structures for Tyr303Ser and Tyr303Trp variants and its complexes with NADP+ show significant differences between plant and cyanobacterial FNRs. Our results suggest that modulation of coenzyme affinity is highly influenced by the strength of the C-terminus-FAD interaction and that subtle changes between plant and cyanobacterial structures are able to modify the energy of that interaction. Additionally, it is shown that the C-terminal Tyr of FNR lowers the affinity for NADP+/H to levels compatible with steady-state turnover during the catalytic cycle, but it is not involved in the hydride transfer itself.  相似文献   

9.
José Ramón Peregrina 《BBA》2010,1797(9):1638-1264
Two transient charge-transfer complexes (CTC) form prior and upon hydride transfer (HT) in the reversible reaction of the FAD-dependent ferredoxin-NADP+ reductase (FNR) with NADP+/H, FNRox-NADPH (CTC-1), and FNRrd-NADP+ (CTC-2). Spectral properties of both CTCs, as well as the corresponding interconversion HT rates, are here reported for several Anabaena FNR site-directed mutants. The need for an adequate initial interaction between the 2′P-AMP portion of NADP+/H and FNR that provides subsequent conformational changes leading to CTC formation is further confirmed. Stronger interactions between the isoalloxazine and nicotinamide rings might relate with faster HT processes, but exceptions are found upon distortion of the active centre. Thus, within the analyzed FNR variants, there is no strict correlation between the stability of the transient CTCs formation and the rate of the subsequent HT. Kinetic isotope effects suggest that, while in the WT, vibrational enhanced modulation of the active site contributes to the tunnel probability of HT; complexes of some of the active site mutants with the coenzyme hardly allow the relative movement of isoalloxazine and nicotinamide rings along the HT reaction. The architecture of the WT FNR active site precisely contributes to reduce the stacking probability between the isoalloxazine and nicotinamide rings in the catalytically competent complex, modulating the angle and distance between the N5 of the FAD isoalloxazine and the C4 of the coenzyme nicotinamide to values that ensure efficient HT processes.  相似文献   

10.
The frequency of nonstereospecific hydride transfers to and from the C-4 of NAD+ by porcine heart lactate dehydrogenase has been determined to be less than 1 out of every 10(7) hydride transfers, using a method employing a dynamic equilibrium which allows for the detection of extremely rare (less than 1 in 10(8] nonstereospecific transfers. A detailed protocol is presented. The inability to detect unequivocally a nonstereospecific transfer either to or from the si face of the nicotinamide ring (pro-4S hydrogen to NADH) indicates that the transition state for the more favorable transfer of the pro-4R hydrogen is at least 10 kcal/mol lower in energy.  相似文献   

11.
Accumulating experimental evidence suggests that the occurrence of hydrogen tunneling is likely to be widespread in enzyme-catalyzed reactions. The realization that hydrogen can transfer via tunneling mechanisms has far-reaching implications for our understanding of enzyme catalysis involving proton, hydride or hydrogen atom transfer reactions. The current status of the field is highlighted by three enzyme systems that have been under intensive study in recent years, including soybean lipoxygenase-1, thermophilic alcohol dehydrogenase and dihydrofolate reductase. Particular attention has been devoted to the issues of whether protein dynamics modulate hydrogen tunneling probability and whether the tunneling process contributes to the catalytic power of enzymes.  相似文献   

12.
D S Sem  C B Kasper 《Biochemistry》1992,31(13):3391-3398
The stereospecificity of hydride abstraction from NADPH and the conformation of the nicotinamide ring around the glycosidic bond have been determined for the flavoprotein NADPH-cytochrome P-450 oxidoreductase (P-450R). The A-side (pro-R) hydrogen is abstracted from NADPH, and the nicotinamide ring is in the anti conformation. These results are consistent with the apparently strong correlation between A-side stereospecificity and anti conformation and between B-side stereospecificity and syn conformation [You, K. (1985) CRC Crit. Rev. Biochem. 17, 313]. This correlation reveals how the flavin and nicotinamide rings are oriented relative to each other. In P-450R, the flavin is then "on top of" (on the exo side of) the nicotinamide ring. In another flavoprotein dehydrogenase, glutathione reductase, which is a B-side/anti enzyme [Pai, E. F., & Schulz, G. E. (1983) J. Biol. Chem. 258, 1752], the flavin is "underneath" (on the endo side of) the nicotinamide ring. We argue that all enzymes that are evolutionarily related to these two flavoproteins should have their respective overall configurations. The overall configuration is defined by the following five properties: (1) relative orientation of the isoalloxazine and nicotinamide rings, (2) stereospecificity of hydride transfer to/from the nicotinamide ring, (3) conformation of the nicotinamide ring around the glycosidic bond, (4) stereospecificity of hydride transfer to/from the flavin, and (5) conformation of the flavin around its N5-N10 axis. There are only eight possible overall configurations, and a knowledge of only three of the five properties is needed to determine which one is present (as long as the combination of properties is not 1, 2, 3 or 1, 4, 5).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The thiamin and biotin biosynthetic pathways utilize elaborate strategies for the transfer of sulfur from cysteine to cofactor precursors. For thiamin, the sulfur atom of cysteine is transferred to a 66-amino-acid peptide (ThiS) to form a carboxy-terminal thiocarboxylate group. This sulfur transfer requires three enzymes and proceeds via a ThiS-acyladenylate intermediate. The biotin synthase Fe-S cluster functions as the immediate sulfur donor for biotin formation. C-S bond formation proceeds via radical intermediates that are generated by hydrogen atom transfer from dethiobiotin to the adenosyl radical. This radical is formed by the reductive cleavage of S-adenosylmethionine by the reduced Fe-S cluster of biotin synthase.  相似文献   

14.
Directed mutagenesis has been used to study the nicotinamide subsite of the glycolytic NAD-dependent glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Residue Asn313 is involved together with the carboxyamide moiety of the nicotinamide ring in a complex network of hydrogen bonding interactions which fix the position of the pyridinium ring of NAD to which hydride transfer occurs at the C-4 position in the catalytic reaction. The asparagine side-chain has been replaced by that of the Thr and Ala residues and results in mutants with very similar properties. Both mutants show much weaker binding of NAD and lower catalytic efficiency. The mutant Asn313----Thr still exhibits strict B-stereospecificity in hydride transfer and retains the property of negative co-operativity in NAD binding. These experiments strongly suggest that the mutant enzyme undergoes the apo----holo sub-unit structural transition associated with coenzyme binding but that the nicotinamide ring is no longer as rigidly held in its pocket as in the wild type enzyme. The results shed light on the details of the molecular interactions which are responsible for negative co-operativity in this enzyme.  相似文献   

15.
The nature of beryllium bonds formed between BeX2 (X is H, F and Cl) and some Lewis bases have been investigated. The distribution of the Laplacian of electron density shows that there is a region of charge depletion around the Be atom, which, according to Laplacian complementary principal, can interact with a region of charge concentration of an atom in the base and form a beryllium bond. The molecular graphs of the investigated complexes indicate that beryllium in BeH2 and BeF2 can form “beryllium bonds” with O, N and P atoms but not with halogens. In addition, eight criteria based on QTAIM properties, including the values of electron density and its Laplacian at the BCP, penetration of beryllium and acceptor atom, charge, energy, volume and first atomic moment of beryllium atom, have been considered and compared with the corresponding ones in conventional hydrogen bonds. These bonds share many common features with very strong hydrogen bonds, however,some differences have also been observed.  相似文献   

16.
The dicarboxylate radical -OOC--CH--CH(OH)COO- was generated in an N2O-saturated fumarate solution by high energy ionizing radiation. When NADH was present in the solution, product analysis indicated a stoichiometry of 2 molecules of the radical reacted with 1 NADH molecule to form 2 malate and 1 enzymatically active NAD+ molecules. In a similar experiment using tritium label on position A of NADH, due to an isotope effect, only 10% of the label was transferred to malate; most of the remaining tritium was found in the NAD+ formed. When lactate dehydrogenase was added, however, no la bel was detectable in NAD+, and over 80% of the tritium lost from NADH was found in malate. The stereospecific transfer of the hydrogen atom from lactate dehydrogenase-bound NADH to the dicarboxylate radical suggested that the free radical reaction must have taken place at the active site. The hydrogen atom transfer was inhibited by oxamate. Results from flow experiments in which an irradiated fumarate solution was mixed with a solutionof lactate dehydrogenase and NADH are in support of a mechanism in which the hydrogen atom transfer occurs in the first oxidation step.  相似文献   

17.
Transhydrogenase (TH) is a dimeric integral membrane enzyme in mitochondria and prokaryotes that couples proton translocation across a membrane with hydride transfer between NAD(H) and NADP(H) in soluble domains. Crystal structures of the NAD(H) binding alpha1 subunit (domain I) of Rhodospirillum rubrum TH have been determined at 1.8 A resolution in the absence of dinucleotide and at 1.9 A resolution with NADH bound. Each structure contains two domain I dimers in the asymmetric unit (AB and CD); the dimers are intimately associated and related by noncrystallographic 2-fold axes. NADH binds to subunits A and D, consistent with the half-of-the-sites reactivity of the enzyme. The conformation of NADH in subunits A and D is very similar; the nicotinamide is in the anti conformation, the A-face is exposed to solvent, and both N7 and O7 participate in hydrogen bonds. Comparison of subunits A and D to six independent copies of the subunit without bound NADH reveals multiple conformations for residues and loops surrounding the NADH site, indicating flexibility for binding and release of the substrate (product). The NADH-bound structure is also compared to the structures of R. rubrum domain I with NAD bound (PDB code 1F8G) and with NAD bound in complex with domain III of TH (PDB code 1HZZ). The NADH- vs NAD-bound domain I structures reveal conformational differences in conserved residues in the NAD(H) binding site and in dinucleotide conformation that are correlated with the net charge, i.e., oxidation state, of the nicotinamides. The comparisons illustrate how nicotinamide oxidation state can affect the domain I conformation, which is relevant to the hydride transfer step of the overall reaction.  相似文献   

18.
Fan F  Gadda G 《Biochemistry》2007,46(21):6402-6408
The hydride transfer reaction catalyzed by choline oxidase under irreversible regime, i.e., at saturating oxygen, was shown in a recent study to occur quantum mechanically within a highly preorganized active site, with the reactive configuration for hydride tunneling being minimally affected by environmental vibrations of the reaction coordinate other than those affecting the distance between the alpha-carbon of the choline alkoxide substrate and the N(5) atom of the enzyme-bound flavin cofactor [Fan, F., and Gadda, G. (2005) J. Am. Chem. Soc. 127, 17954-17961]. In this study, we have determined the effects of pH and temperature on the substrate kinetic isotope effects with 1,2-[2H4]choline as substrate for choline oxidase at 0.2 mM oxygen to gain insights on the mechanism of hydride transfer under reversible catalytic regime. The data presented indicated that the kinetic complexity arising from the net flux through the reverse of the hydride transfer step changed with temperature, with the hydride transfer reaction becoming more reversible with increasing temperatures. After this kinetic complexity was accounted for, analyses of the kcat/Km and D(kcat/Km) values determined at 0.2 mM according to the Eyring and Arrhenius formalisms suggested that the quantum mechanical nature of the hydride transfer reaction is, not surprisingly, maintained during enzymatic catalysis under reversible regime. A comparison of the thermodynamic and kinetic parameters of the hydride transfer reaction under reversible and irreversible catalytic regimes showed that the enthalpies of activation (DeltaH++) were significantly larger in the reversible catalytic regime. This reflects the presence of an enthalpically unfavorable internal equilibrium of the enzyme-substrate Michaelis complex occurring prior to, and independently from, CH bond cleavage. Such an internal equilibrium is required to preorganize the enzyme-substrate complex for efficient quantum mechanical tunneling of the hydride ion from the substrate alpha-carbon to the flavin N(5) atom.  相似文献   

19.
The hydrogen bonds formed by the interaction of nitriles with water, hydrogen fluoride, ammonia and hydrogen sulphide have been studied using B3LYP and second-order Møller–Plesset perturbation (MP2) methods and 6-311+ + G(d,p) basis set. The energies and structures of 80 hydrogen-bonded complexes between nitriles and small molecules were examined systematically using B3LYP and MP2 procedure. Categorisation of the hydrogen bonds involved in the various complexes led to an ordering of hydrogen bond donor and acceptor abilities for some functional groups. The interaction energies have been corrected for the basis set superposition error using Boy's counterpoise correction method. The Morokuma energy decomposition analysis reveals that the strong interactions are due to the attractive contributions from the electrostatic (ES), polarisation (PL) and charge transfer (CT) components. The topological parameters, electron density and Laplacian of electron density show excellent correlation with the hydrogen bond length. Natural bond orbital (NBO) analysis has also been performed to study the CT from proton acceptor to the antibonding orbital of the H–Y bond in the proton donor part of complexes. The frequency analysis of C–H…Y bond in the complexes indicates the blue-shifting nature largely in case of sp2 hybridised carbon atom.  相似文献   

20.
In chlorophyll biosynthesis protochlorophyllide reductase (POR) catalyzes the light-driven reduction of protochlorophyllide (Pchlide) to chlorophyllide, providing a rare opportunity to trap and characterize catalytic intermediates at low temperatures. Moreover, the presence of a chlorophyll-like molecule allows the use of EPR, electron nuclear double resonance, and Stark spectroscopies, previously used for the analysis of photosynthetic systems, to follow catalytic events in the active site of POR. Different models involving the formation of either radical species or charge transfer complexes have been proposed for the initial photochemical step, which forms a nonfluorescent intermediate absorbing at 696 nm (A696). Our EPR data show that the concentration of the radical species formed in the initial photochemical step is not stoichiometric with conversion of substrate. Instead, a large Stark effect, indicative of charge transfer character, is associated with A696. Two components were required to fit the Stark data, providing clear evidence that charge transfer complexes are formed during the initial photochemistry. The temperature dependences of both A696 formation and NADPH oxidation are identical, and we propose that formation of the A696 state involves hydride transfer from NADPH to form a charge transfer complex. A catalytic mechanism of POR is suggested in which Pchlide absorbs a photon, creating a transient charge separation across the C-17-C-18 double bond, which promotes ultrafast hydride transfer from the pro-S face of NADPH to the C-17 of Pchlide. The resulting A696 charge transfer intermediate facilitates transfer of a proton to the C-18 of Pchlide during the subsequent first "dark" reaction.  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