首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The 2.2-A crystal structure of chicken liver dihydrofolate reductase (EC 1.5.1.3, DHFR) has been solved as a ternary complex with NADP+ and biopterin (a poor substrate). The space group and unit cell are isomorphous with the previously reported structure of chicken liver DHFR complexed with NADPH and phenyltriazine [Volz, K. W., Matthews, D. A., Alden, R. A., Freer, S. T., Hansch, C., Kaufman, B. T., & Kraut, J. (1982) J. Biol. Chem. 257, 2528-2536]. The structure contains an ordered water molecule hydrogen-bonded to both hydroxyls of the biopterin dihydroxypropyl group as well as to O4 and N5 of the biopterin pteridine ring. This water molecule, not observed in previously determined DHFR structures, is positioned to complete a proposed route for proton transfer from the side-chain carboxylate of E30 to N5 of the pteridine ring. Protonation of N5 is believed to occur during the reduction of dihydropteridine substrates. The positions of the NADP+ nicotinamide and biopterin pteridine rings are quite similar to the nicotinamide and pteridine ring positions in the Escherichia coli DHFR.NADP+.folate complex [Bystroff, C., Oatley, S. J., & Kraut, J. (1990) Biochemistry 29, 3263-3277], suggesting that the reduction of biopterin and the reduction of folate occur via similar mechanisms, that the binding geometry of the nicotinamide and pteridine rings is conserved between DHFR species, and that the p-aminobenzoylglutamate moiety of folate is not required for correct positioning of the pteridine ring in ground-state ternary complexes. Instead, binding of the p-aminobenzoylglutamate moiety of folate may induce the side chain of residue 31 (tyrosine or phenylalanine) in vertebrate DHFRs to adopt a conformation in which the opening to the pteridine binding site is too narrow to allow the substrate to diffuse away rapidly. A reverse conformational change of residue 31 is proposed to be required for tetrahydrofolate release.  相似文献   

2.
3.
Cody V  Galitsky N  Rak D  Luft JR  Pangborn W  Queener SF 《Biochemistry》1999,38(14):4303-4312
Structural data from two independent crystal forms (P212121 and P21) of the folate (FA) binary complex and from the ternary complex with the oxidized coenzyme, NADP+, and recombinant Pneumocystis carinii dihydrofolate reductase (pcDHFR) refined to an average of 2.15 A resolution, show the first evidence of ligand-induced conformational changes in the structure of pcDHFR. These data are also compared with the crystal structure of the ternary complex of methotrexate (MTX) with NADPH and pcDHFR in the monoclinic lattice with data to 2.5 A resolution. Comparison of the data for the FA binary complex of pcDHFR with those for the ternary structures reveals significant differences, with a >7 A movement of the loop region near residue 23 that results in a new "flap-open" position for the binary complex, and a "closed" position in the ternary complexes, similar to that reported for Escherichia coli (ec) DHFR complexes. In the orthorhombic lattice for the binary FA pcDHFR complex, there is also an unwinding of a short helical region near residue 47 that places hydrophobic residues Phe-46 and Phe-49 toward the outer surface, a conformation that is stabilized by intermolecular packing contacts. The pyrophosphate moiety of NADP+ in the ternary folate pcDHFR complexes shows significant differences in conformation compared with that observed in the MTX-NADPH-pcDHFR ternary complex. Additionally, comparison of the conformations among these four pcDHFR structures reveals evidence for subdomain movement that correlates with cofactor binding states. The larger binding site access in the new "flap-open" loop 23 conformation of the binary FA complex is consistent with the rapid release of cofactor from the product complex during catalysis as well as the more rapid release of substrate product from the binary complex as a result of the weaker contacts of the closed loop 23 conformation, compared to ecDHFR.  相似文献   

4.
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   

5.
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   

6.
The flavoenzyme ferredoxin-NADP+ reductase (FNR) catalyses the production of NADPH in photosynthesis. The three-dimensional structure of FNR presents two distinct domains, one for binding of the FAD prosthetic group and the other for NADP+ binding. In spite of extensive experiments and different crystallographic approaches, many aspects about how the NADP+ substrate binds to FNR and how the hydride ion is transferred from FAD to NADP+ remain unclear. The structure of an FNR:NADP+ complex from Anabaena has been determined by X-ray diffraction analysis of the cocrystallised units to 2.1 A resolution. Structural perturbation of FNR induced by complex formation produces a narrower cavity in which the 2'-phospho-AMP and pyrophosphate portions of the NADP+ are perfectly bound. In addition, the nicotinamide mononucleotide moiety is placed in a new pocket created near the FAD cofactor with the ribose being in a tight conformation. The crystal structure of this FNR:NADP+ complex obtained by cocrystallisation displays NADP+ in an unusual conformation and can be considered as an intermediate state in the process of coenzyme recognition and binding. Structural analysis and comparison with previously reported complexes allow us to postulate a mechanism which would permit efficient hydride transfer to occur. Besides, this structure gives new insights into the postulated formation of the ferredoxin:FNR:NADP+ ternary complex by prediction of new intermolecular interactions, which could only exist after FNR:NADP+ complex formation. Finally, structural comparison with the members of the broad FNR structural family also provides an explanation for the high specificity exhibited by FNR for NADP+/H versus NAD+/H.  相似文献   

7.
Dihydrofolate reductase (DHFR) has several flexible active site loops that facilitate ligand binding and catalysis. Previous studies of backbone dynamics in several complexes of DHFR indicate that the time scale and amplitude of motion depend on the conformation of the active site loops. In this study, information on dynamics is extended to methyl-containing side chains. To understand the role of side chain dynamics in ligand binding and loop conformation, methyl deuterium relaxation rates of Escherichia coli DHFR in binary folate and ternary folate:NADP+ complexes have been measured, together with chi(1) rotamer populations for threonine, isoleucine, and valine residues, determined from measurements of 3J(CgammaCO) and 3J(CgammaN) coupling constants. The results indicate that, in addition to backbone motional restriction in the adenosine-binding site, side chain flexibility in the active site and the surrounding active site loops is diminished upon binding NADP+. Resonances for several methyls in the active site and the surrounding active site loops were severely broadened in the folate:NADP+ ternary complex, suggesting the presence of motion on the chemical shift time scale. The side chains of Ile14 and Ile94, which pack against the nicotinamide and pterin rings of the cofactor and substrate, respectively, exhibit rotamer disorder in the ternary folate:NADP+ complex. Conformational fluctuations of these side chains may play a role in transition state stabilization; the observed line broadening for Ile14 suggests motions on a microsecond/millisecond time scale.  相似文献   

8.
The function of a highly mobile loop in Escherichia coli dihydrofolate reductase was studied by constructing a mutant (DL1) using cassette mutagenesis that had four residues deleted in the middle section of the loop (Met16-Ala19) and a glycine inserted to seal the gap. This part of the loop involves residues 16-20 and is disordered in the X-ray crystal structures of the apoprotein and the NADP+ binary complex but forms a hairpin turn that folds over the nicotinamide moiety of NADP+ and the pteridine moiety of folate in the ternary complex [Bystroff, C., & Kraut, J. (1991) Biochemistry 30, 2227-2239]. The steady-state and pre-steady-state kinetics and two-dimensional 1H NMR spectra were analyzed and compared to the wild-type protein. The kinetics on the DL1 mutant enzyme show that the KM value for NADPH (5.3 microM), the KM for dihydrofolate (2 microM), the rate constant for the release of the product tetrahydrofolate (10.3 s-1), and the intrinsic pKa value (6.2) are similar to those exhibited by the wild-type enzyme. However, the hydride-transfer rate declines markedly from the wild-type value of 950 s-1 to 1.7 s-1 for the DL1 mutant and when taken with data for substrate binding indicates that the loop contributes to substrate flux by a factor of 3.5 x 10(4). Thus, the mobility of loop I may provide a mechanism of recruiting hydrophobic residues which can properly align the nicotinamide and pteridine rings for the hydride-transfer process (a form of transition-state stabilization).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
The NADPH-dependent reduction of ketopantoate to pantoate, catalyzed by ketopantoate reductase (KPR; EC 1.1.1.169), is essential for the biosynthesis of pantothenate (vitamin B(5)). Here we present the crystal structure of Escherichia coli KPR with NADP(+) bound, solved to 2.1 A resolution. The cofactor is bound in the active site cleft between the N-terminal Rossmann-fold domain and the C-terminal alpha-helical domain. The thermodynamics of cofactor and substrate binding were characterized by isothermal titration calorimetry. The dissociation constant for NADP(+) was found to be 6.5 muM, 20-fold larger than that for NADPH (0.34 muM). The difference is primarily due to the entropic term, suggesting favorable hydrophobic interactions of the more lipophilic nicotinamide ring in NADPH. Comparison of this binary complex structure with the previously studied apoenzyme reveals no evidence for large domain movements on cofactor binding. This observation is further supported both by molecular dynamics and by calorimetric analysis. A model of the ternary complex, based on the structure presented here, provides novel insights into the molecular mechanism of enzyme catalysis. We propose a conformational switch of the essential Lys176 from the "resting" state observed in our structure to an "active" state, to bind ketopantoate. Additionally, we identify the importance of Asn98 for substrate binding and enzyme catalysis.  相似文献   

10.
The novel flavin-dependent thymidylate synthase, ThyX, is absent in humans but several pathogenic bacteria depend exclusively on ThyX activity to synthesize thymidylate. Reduction of the enzyme-bound FAD by NADPH is suggested to be the critical first step in ThyX catalysis. We soaked Mycobacterium tuberculosis ThyX-FAD-BrdUMP ternary complex crystals in a solution containing NADP+ to gain structural insights into the reductive step of the catalytic cycle. Surprisingly, the NADP+ displaced both FAD and BrdUMP from the active site. In the resultant ThyX-NADP+ binary complex, the AMP moiety is bound in a deep pocket similar to that of the same moiety of FAD in the ternary complex, while the nicotinamide part of NADP+ is engaged in a limited number of contacts with ThyX. The additional 2'-phosphate group attached to the AMP ribose of NADP+ could be accommodated with minor rearrangement of water molecules. The newly introduced 2'-phosphate groups are engaged in water-mediated interactions across the non-crystallographic 2-fold axis of the ThyX tetramer, suggesting possibilities for design of high-affinity bivalent inhibitors of this intriguing enzyme.  相似文献   

11.
The chemical shifts of all the aromatic proton and anomeric proton resonances of NADP+, NADPH, and several structural analogues have been determined in their complexes with Lactobacillus casei dihydrofolate reductase by double-resonance (saturation transfer) experiments. The binding of NADP+ to the enzyme leads to large (0.9-1.6 ppm) downfield shifts of all the nicotinamide proton resonances and somewhat smaller upfield shifts of the adenine proton resonance. The latter signals show very similar chemical shifts in the binary and ternary complexes of NADP+ and the binary complexes of several other coenzymes, suggesting that the environment of the adenine ring is similar in all cases. In contrast, the nicotinamide proton resonances show much greater variability in position from one complex to another. The data show that the environments of the nicotinamide rings of NADP+, NADPH, and the thionicotinamide and acetylpyridine analogues of NADP+ in their binary complexes with the enzyme are quite markedly different from one another. Addition of folate or methotrexate to the binary complex has only modest effects on the nicotinamide ring of NADP+, but trimethoprim produces a substantial change in its environment. The dissociation rate constant of NADP+ from a number of complexes was also determined by saturation transfer.  相似文献   

12.
The ferredoxin nicotinamide adenine dinucleotide phosphate reductase from Pseudomonas aeruginosa ( pa-FPR) in complex with NADP (+) has been characterized by X-ray crystallography and in solution by NMR spectroscopy. The structure of the complex revealed that pa-FPR harbors a preformed NADP (+) binding pocket where the cofactor binds with minimal structural perturbation of the enzyme. These findings were complemented by obtaining sequential backbone resonance assignments of this 29518 kDa enzyme, which enabled the study of the pa-FPR-NADP complex by monitoring chemical shift perturbations induced by addition of NADP (+) or the inhibitor adenine dinucleotide phosphate (ADP) to pa-FPR. The results are consistent with a preformed NADP (+) binding site and also demonstrate that the pa-FPR-NADP complex is largely stabilized by interactions between the protein and the 2'-P AMP portion of the cofactor. Analysis of the crystal structure also shows a vast network of interactions between the two cofactors, FAD and NADP (+), and the characteristic AFVEK (258) C'-terminal extension that is typical of bacterial FPRs but is absent in their plastidic ferredoxin NADP (+) reductase (FNR) counterparts. The conformations of NADP (+) and FAD in pa-FPR place their respective nicotinamide and isoalloxazine rings 15 A apart and separated by residues in the C'-terminal extension. The network of interactions among NADP (+), FAD, and residues in the C'-terminal extension indicate that the gross conformational rearrangement that would be necessary to place the nicotinamide and isoalloxazine rings parallel and adjacent to one another for direct hydride transfer between NADPH and FAD in pa-FPR is highly unlikely. This conclusion is supported by observations made in the NMR spectra of pa-FPR and the pa-FPR-NADP complex, which strongly suggest that residues in the C'-terminal sequence do not undergo conformational exchange in the presence or absence of NADP (+). These findings are discussed in the context of a possible stepwise electron-proton-electron transfer of hydride in the oxidation of NADPH by FPR enzymes.  相似文献   

13.
Tropinone reductase-II (TR-II) catalyzes the NADPH-dependent reduction of the carbonyl group of tropinone to a beta-hydroxyl group. The crystal structure of TR-II complexed with NADP+ and pseudotropine (psi-tropine) has been determined at 1.9 A resolution. A seven-residue peptide near the active site, disordered in the unliganded structure, is fixed in the ternary complex by participation of the cofactor and substrate binding. The psi-tropine molecule is bound in an orientation which satisfies the product configuration and the stereochemical arrangement toward the cofactor. The substrate binding site displays a complementarity to the bound substrate (psi-tropine) in its correct orientation. In addition, electrostatic interactions between the substrate and Glu156 seem to specify the binding position and orientation of the substrate. A comparison between the active sites in TR-II and TR-I shows that they provide different van der Waals surfaces and electrostatic features. These differences likely contribute to the correct binding mode of the substrates, which are in opposite orientations in TR-II and TR-I, and to different reaction stereospecificities. The active site structure in the TR-II ternary complex also suggests that the arrangement of the substrate, cofactor, and catalytic residues is stereoelectronically favorable for the reaction.  相似文献   

14.
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway.  相似文献   

15.
Proton-translocating transhydrogenase (TH) couples direct and stereospecific hydride transfer between NAD(H) and NADP(H), bound to soluble domains dI and dIII, respectively, to proton translocation across a membrane bound domain, dII. The reaction occurs with proton-gradient coupled conformational changes, which affect the energetics of substrate binding and interdomain interactions. The crystal structure of TH dIII from Rhodospirillum rubrum has been determined in the presence of NADPH (2.4 A) and NADP (2.1 A) (space group P6(1)22). Each structure has two molecules in the asymmetric unit, differing in the conformation of the NADP(H) binding loop D. In one molecule, loop D has an open conformation, with the B face of (dihydro)nicotinamide exposed to solvent. In the other molecule, loop D adopts a hitherto unobserved closed conformation, resulting in close interactions between NADP(H) and side chains of the highly conserved residues, betaSer405, betaPro406, and betaIle407. The conformational change shields the B face of (dihydro)nicotinamide from solvent, which would block hydride transfer in the intact enzyme. It also alters the environments of invariant residues betaHis346 and betaAsp393. However, there is little difference in either the open or the closed conformation upon change in oxidation state of nicotinamide, i.e., for NADP vs. NADPH. Consequently, the occurrence of two loop D conformations for both substrate oxidation states gives rise to four states: NADP-open, NADP-closed, NADPH-open, and NADPH-closed. Because these states are distinguished by protein conformation and by net charge they may be important in the proton translocating mechanism of intact TH.  相似文献   

16.
Type II dihydrofolate reductases (DHFRs) encoded by the R67 and R388 plasmids are sequence and structurally different from known chromosomal DHFRs. These plasmid-derived DHFRs are responsible for confering trimethoprim resistance to the host strain. A derivative of R388 DHFR, RBG200, has been cloned and its physical properties have been characterized. This enzyme has been shown to transfer the pro-R hydrogen of NADPH to its substrate, dihydrofolate, making it a member of the A-stereospecific class of dehydrogenases [Brito, R. M. M., Reddick, R., Bennett, G. N., Rudolph, F. B., & Rosevear, P. R. (1990) Biochemistry 29,9825]. Two distinct binary RBG200.NADP+ complexes were detected. Addition of NADP+ to RBG200 DHFR results in formation of an initial binary complex, conformation I, which slowly interconverts to a second more stable binary complex, conformation II. The binding of NADP+ to RBG200 DHFR in the second binary complex was found to be weak, KD = 1.9 +/- 0.4 mM. Transferred NOEs were used to determine the conformation of NADP+ bound to RBG200 DHFR. The initial slope of the NOE buildup curves, measured from the intensity of the cross-peaks as a function of the mixing time in NOESY spectra, allowed interproton distances on enzyme-bound NADP+ to be estimated. The experimentally measured distances were used to define upper and lower bound distance constraints between proton pairs in distance geometry calculations. All NADP+ structures consistent with the experimental distance bounds were found to have a syn conformation about the nicotinamide-ribose (X = 94 +/- 26 degrees) and an anti conformation about the adenine-ribose (X = -92 +/- 32 degrees) glycosidic bonds.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The three-dimensional structures of D-Phe-Pro-Arg-chloromethyl ketone-inhibited thrombin in complex with Tyr-63-sulfated hirudin (ternary complex) and of thrombin in complex with the bifunctional inhibitor D-Phe-Pro-Arg-Pro-(Gly)4-hirudin (CGP 50,856, binary complex) have been determined by X-ray crystallography in crystal forms different from those described by Skrzypczak-Jankun et al. (Skrzypczak-Jankun, E., Carperos, V.E., Ravichandran, K.G., & Tulinsky, A., 1991, J. Mol. Biol. 221, 1379-1393). In both complexes, the interactions of the C-terminal hirudin segments of the inhibitors binding to the fibrinogen-binding exosite of thrombin are clearly established, including residues 60-64, which are disordered in the earlier crystal form. The interactions of the sulfate group of Tyr-63 in the ternary complex structure explain why natural sulfated hirudin binds with a 10-fold lower K(i) than the desulfated recombinant material. In this new crystal form, the autolysis loop of thrombin (residues 146-150), which is disordered in the earlier crystal form, is ordered due to crystal contacts. Interactions between the C-terminal fragment of hirudin and thrombin are not influenced by crystal contacts in this new crystal form, in contrast to the earlier form. In the bifunctional inhibitor-thrombin complex, the peptide bond between Arg-Pro (P1-P1') seems to be cleaved.  相似文献   

18.
The binding of NADP+ to dihydrofolate reductase (EC 1.5.1.3) in the presence and absence of substrate analogs has been studied using 1H and 13C nuclear magnetic resonance (NMR). NADP+ binds strongly to the enzyme alone and in the presence of folate, aminopterin, and methotrexate with a stoichiometry of 1 mol of NADP+/mol of enzyme. In the 13C spectra of the binary and ternary complexes, separate signals were observed for the carboxamide carbon of free and bound [13CO]NADP+ (enriched 90% in 13C). The 13C signal of the NADP+-reductase complex is much broader than that in the ternary complex with methotrexate because of exchange line broadening on the binary complex signal. From the difference in line widths (17.5 +/- 3.0 Hz) an estimate of the dissociation rate constant of the binary complex has been obtained (55 +/- 10 sec-1). The dissociation rate of the NADP+-reductase complex is not the rate-limiting step in the overall reaction. In the various complexes studied large 13C chemical shifts were measured for bound [13CO]NADP+ relative to free NADP+ (upfield shifts of 1.6-4.3 ppm). The most likely origin of the bound shifts lies in the effects on the shieldings of electric fields from nearby charged groups. For the NADP+-reductase-folate system two 13C signals from bound NADP+ are observed indicating the presence of more than one form of the ternary complex. The IH spectra of the binary and ternary complexes confirm both the stoichiometry and the value of the dissociation rate constant obtained from the 13C experiments. Substantial changes in the IH spectrum of the protein were observed in the different complexes and these are distinct from those seen in the presence of NADPH.  相似文献   

19.
R67 dihydrofolate reductase (DHFR) is a type II DHFR produced by bacteria as a resistance mechanism to the increased clinical use of the antibacterial drug trimethoprim. Type II DHFRs are not homologous in either sequence or structure with chromosomal DHFRs. The type II enzymes contain four identical subunits which form a homotetramer containing a single active site pore accessible from either end. Although the crystal structure of the complex of R67 DHFR with folate has been reported [Narayana et al. (1995) Nat. Struct. Biol. 2, 1018], the nature of the ternary complex which must form with substrate and cofactor is unclear. We have performed transferred NOE and interligand NOE (ILOE) studies to analyze the ternary complexes formed from NADP(+) and folate in order to probe the structure of the ternary complex. Consistent with previous studies of the binary complex formed from another type II DHFR, the ribonicotinamide bond of NADP(+) was found to adopt a syn conformation, while the adenosine moiety adopts an anti conformation. Large ILOE peaks connecting NADP(+) H4 and H5 with folate H9 protons are observed, while the absence of a large ILOE connecting NADP(+) H4 and H5 with folate H7 indicates that the relative orientation of the two ligands differs significantly from the orientation in the chromosomal enzyme. To obtain more detailed insight, we prepared and studied the folate analogue 2-deamino-2-methyl-5,8-dideazafolate (DMDDF) which contains additional protons in order to provide additional NOEs. For this analogue, the exchange characteristics of the corresponding ternary complex were considerably poorer, and it was necessary to utilize higher enzyme concentrations and higher temperature in order to obtain ILOE information. The results support a structure in which the NADP(+) and folate/DMDDF molecules extend in opposite directions parallel to the long axis of the pore, with the nicotinamide and pterin ring systems approximately stacked at the center. Such a structure leads to a ternary complex which is in many respects similar to the gas-phase theoretical calculations of the dihydrofolate-NADPH transition state by Andres et al. [(1996) Bioorg. Chem. 24, 10-18]. Analogous NMR studies performed on folate, DMDDF, and R67 DHFR indicate formation of a ternary complex in which two symmetry-related binding sites are occupied by folate and DMDDF.  相似文献   

20.
Heteronuclear NMR methods have been used to probe the conformation of four complexes of Escherichia coli dihydrofolate reductase (DHFR) in solution. (1)H(N), (15)N, and (13)C(alpha) resonance assignments have been made for the ternary complex with folate and oxidized NADP(+) cofactor and the ternary complex with folate and a reduced cofactor analog, 5,6-dihydroNADPH. The backbone chemical shifts have been compared with those of the binary complex of DHFR with the substrate analog folate and the binary complex with NADPH (the holoenzyme). Analysis of (1)H(N) and (15)N chemical shifts has led to the identification of marker resonances that report on the active site conformation of the enzyme. Other backbone amide resonances report on the presence of ligands in the pterin binding pocket and in the adenosine and nicotinamide-ribose binding sites of the NADPH cofactor. The chemical shift data indicate that the enzyme populates two dominant structural states in solution, with the active site loops in either the closed or occluded conformations defined by X-ray crystallography; there is no evidence that the open conformation observed in some X-ray structures of E. coli DHFR are populated in solution.  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