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 15 毫秒
1.
To address the effects of ligand binding on the structural fluctuations of Escherichia coli dihydrofolate reductase (DHFR), the hydrogen/deuterium (H/D) exchange kinetics of its binary and ternary complexes formed with various ligands (folate, dihydrofolate, tetrahydrofolate, NADPH, NADP(+), and methotrexate) were examined using electrospray ionization mass spectrometry. The kinetic parameters of H/D exchange reactions, which consisted of two phases with fast and slow rates, were sensitively influenced by ligand binding, indicating that changes in the structural fluctuation of the DHFR molecule are associated with the alternating binding and release of the cofactor and substrate. No additivity was observed in the kinetic parameters between a ternary complex and its constitutive binary complexes, indicating that ligand binding cooperatively affects the structural fluctuation of the DHFR molecule via long-range interactions. The local H/D exchange profile of pepsin digestion fragments was determined by matrix-assisted laser desorption/ionization mass spectrometry, and the helix and loop regions that appear to participate in substrate binding, largely fluctuating in the apo-form, are dominantly influenced by ligand binding. These results demonstrate that the structural fluctuation of kinetic intermediates plays an important role in enzyme function, and that mass spectrometry on H/D exchange coupled with ligand binding and protease digestion provide new insight into the structure-fluctuation-function relationship of enzymes.  相似文献   

2.
M A Moore  F Ahmed  R B Dunlap 《Biochemistry》1986,25(11):3311-3317
The proposed mechanism of action of thymidylate synthase envisages the formation of a covalent ternary complex of the enzyme with the substrate dUMP and the cofactor 5,10-methylenetetrahydrofolate (CH2H4folate). The proposed structure of this adduct has been based by analogy on that of the covalent inhibitory ternary complex thymidylate synthase-FdUMP-CH2H4folate. Our recent success in using the protein precipitant trichloroacetic acid to trap the latter complex and covalent binary complexes of the enzyme with FdUMP, dUMP, and dTMP led to the use of this technique in attempts to trap the transient putative covalent catalytic ternary complex. Experiments performed with [2-14C]dUMP and [3',5',7,9-3H]CH2H4folate show that both the substrate and the cofactor remained bound to the protein after precipitation with trichloroacetic acid. The trapped putative covalent catalytic complex was subjected to CNBr fragmentation, and the resulting peptides were fractionated by reverse-phase high-pressure liquid chromatography. The isolated active site peptide was shown to retain the two ligands and was further characterized by a limited sequence analysis using the dansyl Edman procedure. The inhibitory ternary complex, which was formed with [14C]FdUMP and [3H]CH2H4folate, served as a control. The active site peptide isolated from the CNBr-treated inhibitory ternary complex was also subjected to sequence analysis. The two peptides exhibited identical sequences for the first four residues from the N-terminus, Ala-Leu-Pro-Pro, and the fifth amino acid residue was found to be associated with the labeled nucleotides and the cofactor.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Cobalamin-independent methionine synthase (MetE) catalyzes the transfer of the N5-methyl group of methyltetrahydrofolate (CH(3)-H(4)folate) to the sulfur of homocysteine (Hcy) to form methionine and tetrahydrofolate (H(4)folate) as products. This reaction is thought to involve a direct methyl transfer from one substrate to the other, requiring the two substrates to interact in a ternary complex. The crystal structure of a MetE.CH(3)-H(4)folate binary complex shows that the methyl group is pointing away from the Hcy binding site and is quite distant from the position where the sulfur of Hcy would be, raising the possibility that this binary complex is nonproductive. The CH(3)-H(4)folate must either rearrange or dissociate before methyl transfer can occur. Therefore, determining the order of substrate binding is of interest. We have used kinetic and equilibrium measurements in addition to isotope trapping experiments to elucidate the kinetic pathway of substrate binding in MetE. These studies demonstrate that both substrate binary complexes are chemically and kinetically competent for methyl transfer and suggest that the conformation observed in the crystal structure is indeed on-pathway. Additionally, the substrates are shown to bind synergistically, with each substrate binding 30-fold more tightly in the presence of the other. Methyl transfer has been determined to be slow compared to ternary complex formation and dissociation. Simulations indicate that nearly all of the enzyme is present as the ternary complex under physiological conditions.  相似文献   

4.
Substitution of Co(II) for the catalytic site Zn(II) of horse liver alcohol dehydrogenase (LADH) yields an active enzyme derivative, CoIIE, with characteristic Co(II) charge-transfer and d-d electronic transitions that are sensitive to the events which take place during catalysis [Koerber, S. C., MacGibbon, A. K. H., Dietrich, H., Zeppezauer, M., & Dunn, M. F. (1983) Biochemistry 22, 3424-3431]. In this study, UV-visible spectroscopy and rapid-scanning stopped-flow (RSSF) kinetic methods are used to detect and identify intermediates in the LADH catalytic mechanism. In the presence of the inhibitor isobutyramide, the pre-steady-state phase of alcohol (RCH2OH) oxidation at pH above 7 is characterized by the formation and decay of an intermediate with lambda max = 570, 640, and 672 nm for both aromatic and aliphatic alcohols (benzyl alcohol, p-nitrobenzyl alcohol, anisyl alcohol, ethanol, and methanol). By comparison with the spectrum of the stable ternary complex formed with oxidized nicotinamide adenine dinucleotide (NAD+) and 2,2',2'-trifluoroethoxide ion (TFE-), CoIIE(NAD+, TFE-), the intermediate which forms is proposed to be the alkoxide ion (RCH2O-) complex, CoIIE(NAD+, RCH2O-). The timing of reduced nicotinamide adenine dinucleotide (NADH) formation indicates that intermediate decay is limited by the interconversion of ternary complexes, i.e., CoIIE(NAD+, RCH2O-) in equilibrium CoIIE(NADH, RCHO). From competition experiments, we infer that, at pH values below 5, NAD+ and alcohol form a CoIIE(NAD+, RCH2OH) ternary complex. RSSF studies carried out as a function of pH indicate that the apparent pKa values for the ionization of alcohol within the ternary complex, i.e., CoIIE(NAD+, RCH2OH) in equilibrium CoIIE(NAD+, RCH2O-) + H+, fall in the range 5-7.5. Using pyrazole as the dead-end inhibitor, we find that the single-turnover time courses for the reduction of benzaldehyde, p-nitrobenzaldehyde, anisaldehyde, and acetaldehyde at pH above 7 all show evidence for the formation and decay of an intermediate. Via spectral comparisons with CoIIE-(NAD+, TFE-) and with the intermediate formed during alcohol oxidation, we identify the intermediate as the same CoIIE(NAD+, RCH2O-) ternary complex detected during alcohol oxidation.  相似文献   

5.
Tropomyosin (TM) binds to and regulates the actin filament. We used circular dichroism and heteronuclear NMR to investigate the secondary structure and interactions of the C terminus of striated muscle alpha-TM, a major functional determinant, using a model peptide, TM9a(251-284). The (1)H(alpha) and (13)C(alpha) chemical shift displacements show that residues 252 to 277 are alpha-helical but residues 278 to 284 are nonhelical and mobile. The (1)H(N) and (13)C' displacements suggest that residues 257 to 269 form a coiled coil. Formation of an "overlap" binary complex with a 33-residue N-terminal chimeric peptide containing residues 1 to 14 of alpha-TM perturbs the (1)H(N) and (15)N resonances of residues 274 to 284. Addition of a fragment of troponin T, TnT(70-170), to the binary complex perturbs most of the (1)H(N)-(15)N cross-peaks. In addition, there are many new cross-peaks, showing that the binding is asymmetric. Q263, in a proposed troponin T binding site, shows two sets of side-chain (15)N-(1)H cross-peaks, indicating conformational flexibility. The conformational equilibrium of the side chain changes upon formation of the binary and ternary complexes. Replacing Q263 with leucine greatly increases the stability of TM9a(251-284) and reduces its ability to form the binary and ternary complexes, showing that conformational flexibility is crucial for the binding functions of the C terminus.  相似文献   

6.
Binding energetics of phosphorus-containing inhibitors of thermolysin   总被引:1,自引:0,他引:1  
D Grobelny  U B Goli  R E Galardy 《Biochemistry》1989,28(12):4948-4951
The importance of a specific hydrogen bond between thermolysin and a phosphonamidate inhibitor, Z-NHCH2-PO(O-)-Leu-Leu (1) [Bartlett, P. A., & Marlowe, C. K. (1987) Science (Washington D.C.) 235, 569-571], has been reevaluated. We have determined the inhibition constants (binding free energies) for thermolysin of phosphonamidate n-hexyl-P(O)(O-)-Leu-Trp-NHMe (4), phosphonate n-hexyl-P-(O)(O-)OCH(iBu)CO-Trp-NHMe (5), and phosphinates n-hexyl-P(O)(O-)CH2CH(iBu)CO-Trp-NHMe (6) and Z-NHCH2PO(O-)CH2CH(iBu)CO-Leu (3). Replacement of the P-NH group by P-CH2 (1----3 and 4----6) weakens the overall binding free energy by about 1.5 kcal/mol. A negligible difference in solvation energy has been measured for these pairs, and the basicity of the P-O- ligand for zinc in each pair remains nearly unchanged as determined by pH titration of their 31P NMR resonances. Therefore, this value of 1.5 kcal/mol can be assigned to the specific hydrogen bond known to exist between the P-NH of 1 and thermolysin [Tronrud, D. E., Holden, H. M., & Matthews, B. W. (1987) Science (Washington, D.C.) 235, 871-574] and inferred to exist between 4 and the enzyme. Substitution of P-O for P-NH (1----2 [Bartlett, P. A., & Marlowe, C. K. (1987) Science (Washington, D.C.) 235, 569-571] and 4----5) weakens the overall binding free energy by 4.1 kcal/mol for each pair as the basicity of the P-O- ligand decreases by about 1.6 pH units. The measured solvation energy difference between 4 and 5 (and by inference between 1 and 2) is negligible.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Rapid-scanning cryospectroscopy of cobalt(II)-substituted carboxypeptidase A serves to identify and characterize ternary enzyme-substrate-inhibitor (IES) complexes formed by the interaction between the enzyme, a peptide substrate, and a noncompetitive inhibitor. A cobalt absorption spectrum distinct from any induced by peptide or inhibitor alone signals formation of the IES complex. Tight-binding noncompetitive inhibitors containing an aromatic ring, e.g., beta-phenylpropionate, cause the IES complex to form much more slowly than simple binary complexes of the enzyme with either peptide or inhibitor. An inhibitor such as acetate, which binds more weakly and is less bulky, permits the IES complex to form relatively quickly. Remarkably, the cobalt spectra of the IES complexes match those previously found for the steady-state ester (depsipeptide) intermediates. Chemical quenching studies have demonstrated that in these ester intermediates the scissile bond is broken [Galdes, A., Auld, D. S., & Vallee, B. L. (1986) Biochemistry 25, 646-651]. This finding, in conjunction with the present studies, implies that a peptide and a noncompetitive inhibitor of its hydrolysis occupy the same binding loci as the hydrolytic products of a depsipeptide and further indicates that breakdown of an enzyme-biproduct complex is rate-determining for the turnover of depsipeptides.  相似文献   

8.
The crystal structure of ternary and binary substrate complexes of the catalytic subunit of cAMP-dependent protein kinase has been refined at 2.2 and 2.25 A resolution, respectively. The ternary complex contains ADP and a 20-residue substrate peptide, whereas the binary complex contains the phosphorylated substrate peptide. These 2 structures were refined to crystallographic R-factors of 17.5 and 18.1%, respectively. In the ternary complex, the hydroxyl oxygen OG of the serine at the P-site is 2.7 A from the OD1 atom of Asp 166. This is the first crystallographic evidence showing the direct interaction of this invariant carboxylate with a peptide substrate, and supports the predicted role of Asp 166 as a catalytic base and as an agent to position the serine -OH for nucleophilic attack. A comparison of the substrate and inhibitor ternary complexes places the hydroxyl oxygen of the serine 2.7 A from the gamma-phosphate of ATP and supports a direct in-line mechanism for phosphotransfer. In the binary complex, the phosphate on the Ser interacts directly with the epsilon N of Lys 168, another conserved residue. In the ternary complex containing ATP and the inhibitor peptide, Lys 168 interacts electrostatically with the gamma-phosphate of ATP (Zheng J, Knighton DR, Ten Eyck LF, Karlsson R, Xuong NH, Taylor SS, Sowadski JM, 1993, Biochemistry 32:2154-2161). Thus, Lys 168 remains closely associated with the phosphate in both complexes. A comparison of this binary complex structure with the recently solved structure of the ternary complex containing ATP and inhibitor peptide also reveals that the phosphate atom traverses a distance of about 1.5 A following nucleophilic attack by serine and transfer to the peptide. No major conformational changes of active site residues are seen when the substrate and product complexes are compared, although the binary complex with the phosphopeptide reveals localized changes in conformation in the region corresponding to the glycine-rich loop. The high B-factors for this loop support the conclusion that this structural motif is a highly mobile segment of the protein.  相似文献   

9.
Interactive computer graphics was used as a tool in studying the cleavage mechanism of the model substrate Z-Phe-Phe-Leu-Trp by the zinc endopeptidase thermolysin. Two Michaelis complexes and three binding orientations of the tetrahedral intermediate to the crystal structure of thermolysin were investigated. Our results indicate that a Michaelis complex, which does not involve coordination of the scissile peptide to the zinc, is consistent with available experimental data and the most plausible of the two complexes. A tetrahedral intermediate complex wherein the two oxygens of the hydrated scissile peptide straddle the zinc in a bidentate fashion results in the most favorable interactions with the active site. The preferred tetrahedral intermediate and Michaelis complex provide a rationalization for the published substrate data. A trajectory for proceeding from the Michaelis complex to the tetrahedral intermediate is proposed. This trajectory involves a simultaneous activation of the zinc-bound water molecule concurrent with attack on the scissile peptide. A detailed ordered product release mechanism is also presented. These studies suggest some modifications and a number of extensions to the mechanism proposed earlier [Kester, W. R., & Matthews, B. W. (1977) Biochemistry 16, 2506; Holmes, M. A., & Matthews, B. W. (1981) Biochemistry 20, 6912]. The binding mode of the thermolysin inhibitor N-(1-carboxy-3-phenylpropyl)-L-leucyl-L-tryptophan [Monzingo, A. F., & Matthews, B. W. (1984) Biochemistry (preceding paper in this issue)] is compared with that of the preferred tetrahedral intermediate, providing insight into this inhibitor design.  相似文献   

10.
We have found previously (Fahien, L.A., Kmiotek, E.H., MacDonald, M. J., Fibich, B., and Mandic, M. (1988) J. Biol. Chem. 263, 10687-10697) that glutamate-malate oxidation can be enhanced by cooperative binding of mitochondrial aspartate aminotransferase and malate dehydrogenase to the alpha-ketoglutarate dehydrogenase complex. The present results demonstrate that glutamate dehydrogenase, which forms binary complexes with these enzymes, adds to this ternary complex and thereby increases binding of the other enzymes. Kinetic evidence for direct transfer of alpha-ketoglutarate and NADH, within these complexes, has been obtained by measuring steady-state rates of E2 when most of the substrate or coenzyme is bound to the aminotransferase or glutamate dehydrogenase (E1). Rates significantly greater than those which can be accounted for by the concentration of free ligand, calculated from the measured values of the E1-ligand dissociation constants, require that the E1-ligand complex serve as a substrate for E2 (Srivastava, D. K., and Bernhard, S. A. (1986) Curr. Tops. Cell Regul. 28, 1-68). By this criterion, NADH is transferred directly from glutamate dehydrogenase to malate dehydrogenase and alpha-ketoglutarate is channeled from the aminotransferase to both glutamate dehydrogenase and the alpha-ketoglutarate dehydrogenase complex. Similar evidence indicates that GTP bound to an allosteric site on glutamate dehydrogenase functions as a substrate for succinic thiokinase. The potential physiological advantages to channeling of activators and inhibitors as well as substrates within multienzyme complexes organized around the alpha-ketoglutarate dehydrogenase complex are discussed.  相似文献   

11.
Yang L  Su Y  Liu W  Jin X  Wu J 《Carbohydrate research》2002,337(16):1485-1493
The crystal structures of CaCl(2).galactitol.4 H(2)O and 2EuCl(3).galactitol.14 H(2)O were determined to compare the coordination behavior of Ca and lanthanide ions. The crystal system of the Ca-galactitol complex, CaCl(2).C(6)H(14)O(6).4 H(2)O, is monoclinic, Cc space group. Each Ca ion is coordinated to eight oxygen atoms, four from two galactitol molecules and four from water molecules. Galactitol provides O-2, -3 to coordinate to one Ca(2+), and O-4, -5 with another Ca(2+), to form a chain structure. The crystal system of the Eu-galactitol complex, 2EuCl(3).C(6)H(14)O(6).14 H(2)O, is triclinic, P1; space group. Each Eu ion is coordinated to nine oxygen atoms, three from an alditol molecule and six from water molecules. Each galactitol provides O-1, -2, -3 to coordinate with one Eu(3+) and O-4, -5, -6 with another Eu(3+). The other water molecules are hydrogen-bonded in the structure. The similar IR spectra of Pr-, Nd-, Sm-, Eu-, Dy-, and Er-galactitol complexes show that those lanthanide ions have the same coordination mode to neutral galactitol. The Raman spectra also confirm the formation of metal ion-carbohydrate complexes.  相似文献   

12.
The formation of binary and ternary complexes was investigated by ESR spectroscopy in aqueous solutions of Cu(2+), ATP and the dipeptides glygly and gly-L-pro at room temperature. Spectra and stability constants of two ternary complexes for each peptide. (GG)Cu(ATP)(3?), (GG)Cu(ATP)(4?), (GP)cu(ATP)(3?) and (GP(Cu(ATP)(5?) were determined. Assuming that complexes of similar structure show similar spectra, some conclusions could be drawn about the structure of the complexes. The characteristic difference between gly-L-pro and glygly is attributed to the lack of the peptide proton in gly-L-pro. At acidic pH Cu(2+) is bound in binary ATP complexes, at neutral to basic pH in binary peptide or in ternary peptide-Cu-ATP complexes.  相似文献   

13.
Three new ternary peptide-Cu(II)-1,10-phenanthroline (phen) complexes, [Cu(L-ala-gly)(phen)].3.5H(2)O 1, [Cu(L-val-gly)(phen)] 2 and [Cu(gly-L-trp)(phen)].2H(2)O 3, have been prepared and structurally characterised. These compounds exist as distorted square pyramidal complexes with the five co-ordination sites occupied by the tridentate peptide dianion and the two heterocyclic nitrogens of the phenanthroline ligand. The bulk of the lateral chain in the peptide moiety determines the relative disposition of the phen ligand. Thus, in [Cu(L-val-gly)(phen)] 2, the phenanthroline plane is deviated towards the opposite side of the isopropyl group of the L-valine moiety. On the other hand, in [Cu(gly-L-trp)(phen)].2H(2)O 3 the absence of stacking interactions between phen and indole rings and the presence of an intramolecular CH...pi interaction should be pointed out. These complexes exhibit significant differences in their nuclease activity which depends on the nature of the peptidic moiety, the complex [Cu(gly-L-trp) (phen)].2H(2)O 3 being the most active.  相似文献   

14.
J Ellis  C R Bagshaw  W V Shaw 《Biochemistry》1991,30(44):10806-10813
Chloramphenicol acetyltransferase (CAT) catalyzes the acetyl-CoA-dependent acetylation of chloramphenicol by a ternary complex mechanism with a rapid equilibrium and essentially random order of addition of substrates. Such a kinetic mechanism for a two-substrate reaction provides an opportunity to compare the affinity of enzyme for each substrate in the binary complexes (1/Kd) with corresponding values (1/Km) for affinities in the ternary complex where any effect of the other substrate should be manifest. The pursuit of such information for CAT involved the use of four independent methods to determine the dissociation constant (Kd) for chloramphenicol in the binary complex, techniques which included stopped-flow measurements of on and off rates, and a novel fluorometric titration method. The binary complex dissociation constant (Kd) for acetyl-CoA was measured by fluorescence enhancement and steady-state kinetic analysis. The ternary complex dissociation constant (Km) for each substrate (in the presence of the other) was determined by kinetic and fluorometric methods, using CoA or ethyl-CoA to form nonproductive ternary complexes. The results demonstrate an unequivocal decrease in affinity of CAT for each of its substrates on progression from the binary to the ternary complex, a phenomenon most economically described as negative cooperativity. The binary complex dissociation constants (Kd) for chloramphenicol and acetyl-CoA are 4 microM and 30 microM whereas the corresponding dissociation constants in the ternary complex (Km) are 12 microM and 90 microM, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
In order to understand more fully the structural features of aldo-keto reductases (AKRs) that determine their substrate specificities it would be desirable to obtain crystal structures of an AKR with a substrate at the active site. Unfortunately the reaction mechanism does not allow a binary complex between enzyme and substrate and to date ternary complexes of enzyme, NADP(H) and substrate or product have not been achieved. Previous crystal structures, in conjunction with numerous kinetic and theoretical analyses, have led to the general acceptance of the active site tyrosine as the general acid–base catalytic residue in the enzyme. This view is supported by the generation of an enzymatically inactive site-directed mutant (tyrosine-48 to phenylalanine) in human aldose reductase [AKR1B1]. However, crystallization of this mutant was unsuccessful. We have attempted to generate a trapped cofactor/substrate complex in pig aldehyde reductase [AKR1A2] using a tyrosine 50 to phenylalanine site-directed mutant. We have been successful in the generation of the first high resolution binary AKR–Y50F:NADP(H) crystal structure, but we were unable to generate any ternary complexes. The binary complex was refined to 2.2A and shows a clear lack of density due to the missing hydroxyl group. Other residues in the active site are not significantly perturbed when compared to other available reductase structures. The mutant binds cofactor (both oxidized and reduced) more tightly but shows a complete lack of binding of the aldehyde reductase inhibitor barbitone as determined by fluorescence titrations. Attempts at substrate addition to the active site, either by cocrystallization or by soaking, were all unsuccessful using pyridine-3-aldehyde, 4-carboxybenzaldehyde, succinic semialdehyde, methylglyoxal, and other substrates. The lack of ternary complex formation, combined with the significant differences in the binding of barbitone provides some experimental proof of the proposal that the hydroxyl group on the active site tyrosine is essential for substrate binding in addition to its major role in catalysis. We propose that the initial event in catalysis is the binding of the oxygen moiety of the carbonyl-group of the substrate through hydrogen bonding to the tyrosine hydroxyl group.  相似文献   

16.
The formation of binary and ternary complexes of Ni(II) with two biologically relevant molecules, 2'-deoxyguanosine 5'-monophosphate (dGMP) and l-histidine (histidine or His) was characterized by potentiometry and UV-visible spectroscopy. For dGMP, the mononuclear complexes with stoichiometries NiH(2)L(+), NiHL and NiL(-) were found. In the mixed system the ternary complexes NiH(2)LA, NiHLA(-) and NiLA(2-) were detected. In binary systems, the Ni(II) ion coordinates to dGMP through the N-7 atom of its purine ring and indirectly through a water molecule bonded to the phosphate group, while in ternary complexes Ni(II) is bonded to all three histidine donors and directly to the phosphate group of dGMP. Both binary and ternary complexes are susceptible to oxidation by H(2)O(2), with the increased formation of 8-oxo-dGMP in the ternary system. The toxicological relevance of these findings stems from possible disturbance by the major biological Ni(II)-His complex of the nucleotide pools homeostasis through the formation of ternary species and oxidation promotion, as well as from 8-oxo-dGMP capacity to inhibit enzymatic elimination of promutagenic oxidized nucleotides from such pools.  相似文献   

17.
Four ternary metal--ortho-iodohippurate (I-hip)--acyclovir (ACV) complexes, [M(I-hip)(2)(ACV)(H(2)O)(3)] where M is Co(II) (1), Ni(II) (2), Cu (3) and Zn(II) have been obtained by reaction between the corresponding binary complexes M(II)(I-hip)(2)xnH(2)O and ACV. Three ternary complexes (M=Co, Ni and Zn) and the corresponding Zn(II)--ortho-iodohippurate binary derivative have been structurally characterized by X-ray diffraction: The studies show these three ternary complexes are isostructural and present, in solid state, an interesting stacking between the nucleobase and the aryl ring of the hippurate moiety, which probably promotes the formation of ternary complexes. Moreover, the two different ligands interact between them by means of ancillary hydrogen bonds with water molecules coordinated to the metal ion. It must be mentioned that these two recognition factors, hydrogen bonds plus stacking, could explain the reason for the isostructurality of these ternary derivatives with so different three metal ions, with diverses trends in coordination numbers and geometries. In solid state, there are two enantiomeric molecules that are related by an inversion center as the crystal-building unit (as a translational motif) for the ternary complexes.  相似文献   

18.
R C Bethell  G Lowe 《Biochemistry》1988,27(4):1125-1131
A new synthetic route to adenosine 5'-([gamma(R)-17O,18O]-gamma-thiotriphosphate) is described which combines chemical methods for introducing the heavy oxygen isotopes and enzymic methods for achieving the enantiospecificity. This material was used as a substrate for the activation of glutamate catalyzed by glutamine synthetase from Salmonella typhimurium. Analysis of the chirality of the [16O,17O,18O]thiophosphate produced showed that the reaction proceeds with inversion of configuration on phosphorus. This result, taken together with the positional isotope exchange studies of Midelfort and Rose [Midelfort, C. F., & Rose, I.A. (1976) J. Biol. Chem. 251, 5881-5887], demonstrates that the activation of glutamate to form gamma-glutamyl phosphate proceeds by a direct "in-line" transfer of the phosphoryl group.  相似文献   

19.
Aree T  Chaichit N 《Carbohydrate research》2003,338(15):1581-1589
A new crystal form of beta-cyclodextrin (beta-CD)[bond]ethanol[bond]dodecahydrate inclusion complex [(C(6)H(10)O(5))(7).0.3C(2)H(5)OH.12H(2)O] belongs to monoclinic space group C2 (form II) with unit cell constants a=19.292(1), b=24.691(1), c=15.884(1) A, beta=109.35(1) degrees. The beta-CD macrocycle is more circular than that of the complex in space group P2(1) [form I: J. Am. Chem. Soc. 113 (1991) 5676]. In form II, a disordered ethanol molecule (occupancy 0.3) is placed in the upper part of beta-CD cavity (above the O-4 plane) and is sustained by hydrogen bonding to water site W-2. In form I, an ethanol molecule located below the O-4-plane is well ordered because it hydrogen bonds to surrounding O-3[bond]H, O-6[bond]H groups of the symmetry-related beta-CD molecules. In the crystal lattice of form I, beta-CD macrocycles are stacked in a typical herringbone cage structure. By contrast, the packing structure of form II is a head-to-head channel that is stabilized at both O-2/O-3 and O-6 sides of each beta-CD by direct O(CD)...O(CD) and indirect O(CD)...O(W)...(O(W))...O(CD) hydrogen bonds. The 12 water molecules are disordered in 18 positions both inside the channel-like cavity of beta-CD dimer (W-1[bond]W-6) and in the interstices between the beta-CD macrocycles (W-7[bond]W-18). The latter forms a cluster that is hydrogen bonded together and to the neighboring beta-CD O[bond]H groups.  相似文献   

20.
Circular dichroism spectra and circular dichroism difference spectra, generated when porcine heart mitochondrial and supernatant malate dehydrogenase bind coenzymes or when enzyme dihydroincotinamide nucleotide binary complexes bind substrate analogs, are presented. No significant changes are observed in protein chromophores in the 200- to 240-nm spectral range indicating that there is apparently little or no perturbation of the alpha helix or peptide backbone when binary or ternary complexes are formed. Quite different spectral perturbances occur in the two enzymes with reduced coenzyme binding as well as with substrate-analog binding by enzyme-reduced coenzyme binding. Comparison of spectral perturbations in both enzymes with oxidized or reduced coenzyme binding suggests that the dihydronicotinamide moiety of the coenzyme interacts with or perturbs indirectly the environment of aromatic amino acid residues. Reduced coenzyme binding apparently perturbs tyrosine residues in both mitochondrial malate dehydrogenase and lactic dehydrogenase. Reduced coenzyme binding perturbs tyrosine and tryptophan residues in supernatant malate dehydrogenase. The number of reduced coenzyme binding sites was determined to be two per 70,000 daltons in the mitochondrial enzyme, and the reduced coenzyme dissociation constants, determined through the change in ellipticity at 260 nm, with dihydronicotinamide adenine dinucleotide binding, were found to be good agreement with published values (Holbrook, J. J., and Wolfe, R. G. (1972) Biochemistry 11, 2499-2502) obtained through fluorescence-binding studies and indicate no apparent extra coenzyme binding sites. When D-malate forms a ternary complex with malate dehydrogenase-reduced coenzyme complexes, perturbation of both adenine and dihydronicotinamide chromophores is evident. L-Malate binding, however, apparently produces only a perturbation of the adenine chromophore in such complexes. Since the coenzyme has been found to bind in an open conformation on the surface of the enzyme and the substrate analogs bind at or very near the dihydronicotinamide moiety binding site, protein conformational changes are implicated during ternary complex formation with D-malate which can effect the adenine chromophore at some distance from the substrate binding site.  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