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 250 毫秒
1.
We conclude from X-ray diffraction studies at low resolution (7 Å) that the binding of sugar and nucleotide substrates to dimeric yeast hexokinase BII crystals exhibits both negative co-operativity and positive allosteric co-operativity. Difference electron density maps show the positions of sugar and nucleotide binding sites and extensive substrate-induced structural changes in the protein. Sugar substrates and inhibitors bind in the deep cleft that divides each subunit into two lobes and nucleotide substrates bind nearby to one site per dimer, which lies between the subunits and on the molecular symmetry axis. Although the inhibitors o- and p-iodobenzoylglucosamine and o-toluoylglucosamine bind equally to both subunits, the degree of substitution of glucose or xylose is very different for the two subunits. The substrate analog β, γ-imido ATP shows only one strong binding site per dimer. This negative co-operativity in substrate binding may result from the heterologous or non-equivalent association of the two subunits (Anderson et al., 1974), which provides non-equivalent environments for the two chemically identical subunits.Further, there is a positive allosteric interaction between the sugar and nucleotide binding sites. Sugar binding is required for nucleotide binding at the intersubunit site and the binding of nucleotide modifies the binding of sugars. These positive heterotropic interactions appear to be mediated by extensive substrate-induced structural changes in the enzyme.  相似文献   

2.
Structure and refinement of penicillopepsin at 1.8 A resolution   总被引:15,自引:0,他引:15  
Penicillopepsin, the aspartyl protease from the mould Penicillium janthinellum, has had its molecular structure refined by a restrained-parameter least-squares procedure at 1.8 Å resolution to a conventional R-factor of 0.136. The estimated co-ordinate accuracy for the majority of the 2363 atoms of the enzyme is better than 0.12 Å. The average atomic thermal vibration parameter, B, for the atoms of the enzyme is 14.5 Å2. One determining factor of this low average B value is the large central hydrophobic core, in which there are two prominent clusters of aromatic residues, one of nine, the other of seven residues. The N and C-terminal domains of penicillopepsin display an approximate 2-fold symmetry: 70 residue pairs are topologically equivalent, related by a rotation of 177 ° and a translation of 1.2 Å. The analysis of the secondary structural features of the molecule reveals non-linear hydrogen bonding. In penicillopepsin, there is no difference in the mean hydrogen-bond parameters for the elements of α-helix, parallel or antiparallel β-pleated sheet. The mean values for these structural elements are: NO, 2.90 Å; NHO, 1.95 Å; N?O, 160 °. The average hydrogen-bond parameters of the reverse β-turns and the 310 helices are distinctly different from the above values. The analysis of sidechain conformational angles χ1 and χ2 penicillopepsin and other enzyme structures refined in this laboratory shows much narrower distributions as compared with those compiled from unrefined protein structures. The close proximity of the carboxyl groups of Asp33 and Asp213 suggests that they share a proton in a tight hydrogen-bonded environment (Asp33OD2 to Asp213OD1 is 2.87 Å). There are several solvent molecules in the active site region and, in particular, O39 forms hydrogen-bonded interactions with both aspartate residues. The disposition of the two carboxyl groups suggests that neither is likely to be involved in a direct nucleophilic attack on the scissile bond of a substrate. The average atomic B-factors of the residues in this region of the molecule are between 5 and 8 Å2, confirming the proposal that conformational mobility of the active site residues has no role in the enzymatic mechanism. However, conformational mobility of neighbouring regions of the molecule e.g. the “flap” containing Tyr75, is verified by the high B-factors for those residues. The positions of 319 solvent sites per asymmetric unit have been selected from difference electron density maps and refined. Thirteen have been classified as internal, and several of these may have key roles during catalysis. The positively charged Nζ atom of Lys304 forms hydrogen bonds to the carboxylate of Asp14 (internal ion pair) and to two internal water molecules O5 and O25. The protonated side-chain of Asp300 forms a hydrogen bond to Thr214O, 2.78 Å, and is the recipient of a hydrogen bond from a surface pocket water molecule O46. There is no possibility for direct interaction between Asp300 and Lys304 without large conformational changes of their environment. The intermolecular packing involves many protein-protein contacts (66 residues) with a large number of solvent molecules involved in bridging between polar residues at the contact surface. The penicillopepsin molecules resemble an approximate hexagonal close-packing of spheres with each molecule having 12 “nearest” neighbours.  相似文献   

3.
The complex formed by porcine pancreatic kallikrein A with the bovine pancreatic trypsin inhibitor (PTI) has been crystallized at pH 4 in tetragonal crystals of space group P41212 with one molecule per asymmetric unit. Its crystal structure has been solved applying Patterson search methods and using a model derived from the bovine trypsin-PTI complex (Huber et al., 1974) and the structure of porcine pancreatic kallikrein A (Bode et al., 1983). The kallikrein-PTI model has been crystallographically refined to an R-value of 0·23 including X-ray data to 2·5 Å.The root-mean-square deviation, including all main-chain atoms, is 0·45 Å and 0·65 Å for the PTI and for the kallikrein component, respectively, compared with the refined models of the free components. The largest differences are observed in external loops of the kallikrein molecule surrounding the binding site, particularly in the C-terminal part of the intermediate helix around His172. Overall, PTI binding to kallikrein is similar to that of the trypsin complex. In particular, the conformation of the groups at the active site is identical within experimental error (in spite of the different pH values of the two structures). Ser195 OG is about 2·5 Å away from the susceptible inhibitor bond Lys15 C and forms an optimal 2·5 Å hydrogen bond with His57 NE.The PTI residues Thr11 to Ile18 and Val34 to Arg39 are in direct contact with kallikrein residues and form nine intermolecular hydrogen bonds. The reactive site Lys15 protrudes into the specificity pocket of kallikrein as in the trypsin complex, but its distal ammonium group is positioned differently to accommodate the side-chain of Ser226. Ser226 OG mediates the ionic interaction between the ammonium group and the carboxylate group of Asp189. Model-building studies indicate that an arginine side-chain could be accommodated in this pocket. The PTI disulfide bridge 14–38 forces the kallikrein residue Tyr99 to swing out of its normal position. Model-building experiments show that large hydrophobic residues such as phenylalanine can be accommodated at this (S2) site in a wedge-shaped hydrophobic cavity, which is formed by the indole ring of Trp215 and by the phenolic side-chain of Tyr99, and which opens towards the bound inhibitor/substrate chain. Arg17 in PTI forms a favorable hydrogen bond and van der Waals' contacts with kallikrein residues, whereas the additional hydrogen bond formed in the trypsin-PTI complex between Tvr39 OEH and Ile19 N is not possible The kallikrein binding site offers a qualitative explanation of the unusual binding and cleavage at the N-terminal Met-Lys site of kininogen. Model-building experiments suggest that the generally restricted capacity of kallikrein to bind protein inhibitors with more extended binding segments might be explained by steric hindrance with some extruding external loops surrounding the kallikrein binding site (Bode et al., 1983).  相似文献   

4.
In the present work, in order to investigate the electronic excited-state intermolecular hydrogen bonding between the chromophore coumarin 153 (C153) and the room-temperature ionic liquid N,N-dimethylethanolammonium formate (DAF), both the geometric structures and the infrared spectra of the hydrogen-bonded complex C153–DAF+ in the excited state were studied by a time-dependent density functional theory (TDDFT) method. We theoretically demonstrated that the intermolecular hydrogen bond C1?=?O1···H1–O3 in the hydrogen-bonded C153–DAF+ complex is significantly strengthened in the S1 state by monitoring the spectral shifts of the C=O group and O–H group involved in the hydrogen bond C1?=?O1···H1–O3. Moreover, the length of the hydrogen bond C1?=?O1···H1–O3 between the oxygen atom and hydrogen atom decreased from 1.693 Å to 1.633 Å upon photoexcitation. This was also confirmed by the increase in the hydrogen-bond binding energy from 69.92 kJ mol?1 in the ground state to 90.17 kJ mol?1 in the excited state. Thus, the excited-state hydrogen-bond strengthening of the coumarin chromophore in an ionic liquid has been demonstrated theoretically for the first time.  相似文献   

5.
The enthalpies of the hexokinase-catalyzed phosphorylation or glucose, mannose, and fructose by ATP to the respective hexose 6-phosphates have been measured calorimetrically in TRIS/TRIS HCl buffer at 25.0, 28.5, and 32.0°C. The effects on the measured enthalpy of the glucose/hexokinase reaction due to variation of pH (over the range 6.7 to 9.0) and ionic strength (over the range 0.02 to 0.25) have been examined. Correction for enthalpy of buffer protonation leads to δHo and δCpo values for the processes: eq-D-hexose + ATP4− = eq-D-hexose 6-phosphate2− + ADP3−+ H+. Results are δHo = −23.8 ± 0.7 kJ · mol−1 and δCpo = −156 ± 280 J·mol−1·K−1 for glucose. δHo = −21.9 ± 0.7 kJ·mol−1 and δCpo = 10 ± 140 J·mol−1·K−1 for mannose, and δHo = −15.0 ± 0.9 kJ·mol−1 and δCpo = −41 ± 160 J·mol−1·K−1 for fructose. Combination of these measured enthalpies with Gibbs energy data for hydrolysis of ATP4− and that for the hexose 6-phosphates lead to δSo values for the above hexokinase-catalyzed reactions.  相似文献   

6.
《Inorganica chimica acta》1986,119(2):227-232
Interaction between D-glucuronic acid and hydrated uranyl salts has been studied in aqueous solution and solid complexes of the type UO2(D- glucuronate)X·2H20 and UO2(D-glucuronate)2·2H2O, where X = CI, Br or NO3, are isolated and characterized by means of FT-IR and proton-NMR spectroscopy.On comparison with the structurally identified Ca(D-glucuronate)Br·3H2O compound, it is concluded that the UO22+ cation binds to two D- glucuronate moieties in uranylsugar complexes via O6, O5 oxygen atoms (ionized carboxyl group) of the first and O6′, 04 (non-ionized carboxyl group) of the second sugar moiety, whereas in the UO2(D- glucuronate)2·2H2O salt the uranyl ion is bonded to two sugar anions through O6, O6′ oxygen atoms of the ionized carboxyl group, resulting in a six- coordination geometry around the uranium ion. The strong intermolecular hydrogen bonding network of the free acid is rearranged upon sugar metalation and the sugar moiety showed β-anomer conformation both in the free acid and in these uranylsugar complexes.  相似文献   

7.
Ribose-5-phosphate isomerase A (RpiA) plays an important role in interconverting between ribose-5-phosphate (R5P) and ribulose-5-phosphate in the pentose phosphate pathway and the Calvin cycle. We have determined the crystal structures of the open form RpiA from Vibrio vulnificus YJ106 (VvRpiA) in complex with the R5P and the closed form with arabinose-5-phosphate (A5P) in parallel with the apo VvRpiA at 2.0 Å resolution. VvRpiA is highly similar to Eschericihia coliRpiA, and the VvRpiA-R5P complex strongly resembles the E. coli RpiA-A5P complex. Interestingly, unlike the E. coli RpiA-A5P complex, the position of A5P in the VvRpiA-A5P complex reveals a different position than the R5P binding mode. VvRpiA-A5P has a sugar ring inside the binding pocket and a phosphate group outside the binding pocket: By contrast, the sugar ring of A5P interacts with the Asp4, Lys7, Ser30, Asp118, and Lys121 residues; the phosphate group of A5P interacts with two water molecules, W51 and W82.  相似文献   

8.
《Inorganica chimica acta》1987,133(2):347-352
When crystals of [Dy(OH2)7(OHMe)] [DyCl(OH2)2(18- crown-6)]2Cl7·2H2O [1] are allowed to warm from 5 °C to ambient temperature (22 °C) under the original solvent mixture (1:3 CH3OH: CH3CN), they redissolve and the title complex can be isolated by slow evaporation of the resulting solution. The crystal structure of this complex, [Dy(OH2)8]Cl3·18-crown-6·4H2O, has been determined. It crystallizes in the monoclinic space group, P21/c, with a = 10.395(1), b = 18.684(1), c = 16.259- (3) Å, β= 102.56(1)°, and Dcalc = 1.61 g cm−3 for Z = 4. A final conventional R value of 0.041 was obtained by least-squares refinement using 3453 independent observed [Fo⩾5σ(Fo)] reflections. The [Dy(OH2)8]3+ cations and crown ether molecules are hydrogen bonded in a polymeric chain with the crown molecules separating the cations and a total of seven DyOH2···O(crown ether) hydrogen bonds. The chains are connected by a hydrogen bonding network consisting of the cations, chloride ions, and uncoordinated water molecules. The geometry of the cation is best described as a bicapped trigonal prism with distortions on the reaction pathway toward dodecahedral symmetry. The two capping atoms average 2.41(1) Å from Dy, the remaining DyO distances average 2.38(2) Å. The 18-crown-6 molecule has the D3d conformation normally observed except for a distortion of one OCCO unit containing the oxygen atom accepting two hydrogen bonds.  相似文献   

9.
《Inorganica chimica acta》1986,120(2):177-184
The dark blue dimeric complex di-μ-hydroxo- bis [(1,4,7,10-tetraazacyclododecane)chromium(III)] dithionate tetrahydrate, [Cr(C8H20N4)OH]2(S2O6)2· 4H2O or [Cr(cyclen)OH]2(S2O6)2·4H2O, has been synthesized. The crystal structure of the complex has been determined from threeodimensional counter X-ray data. The complex crystallizes in space group P21/n of the monoclinic system with two dimeric formula units in a cell of dimensions a = 8.837(5), b = 14.472(8), c= 13.943(6) Å andβ=95.83(4)o. The structure has been refined by full-matrix least- squares methods to a final value of the weighted R-factor of 0.059 on the basis of 1774 independent intensities. The geometry of the cyclen macrocycle is unsymmetrical, the observed conformations being λδδλ and its enantiomer. The strained ligand conformation leads to significant deviations from octahedral geometry at the chromium centers, and to a bridged geometry in which the CrOCr angle ø and the Cr···Cr separation of 104.1(1)o and 3.086(2) Å are the largest observed in dimers of this kind. The magnetic susceptibility of the complex indicates antiferromagnetic coupling, with the ground state singlet lying 21.56(6) cm−1 below the lowest lying triplet state. The structural parameters have been used to calculate the triplet energy by means of the Glerup- Hodgson- Pedersen (GHP) model, and the calculated value of 22.3 cme−1 is very similar to the observed value.  相似文献   

10.
The α-lytic protease was isolated from an extracellular filtrate of the soil microorganism Myxobacter 495. Trigonal crystals (space group, P3221) of this serine enzyme were grown from 1·3 m-Li2SO4 at pH 7·2. X-ray reflections from crystals of the native enzyme, comprising the 2·8 Å limiting sphere, were phased by the multiple isomorphous replacement technique. Five heavy-atom derivatives were used and the overall mean figure of merit 〈m?〉 is 0·83. The resulting native electron density map of α-lytic protease has been interpreted in conjunction with the published sequence (Olson et al., 1970) of 198 amino-acid residues.α-Lytic protease has a structural core similar to that of the pancreatic serine proteases (108 α-carbon atom positions are topologically equivalent (within 2·0 Å) to residues of porcine elastase) and its tertiary structure is even more closely related to the two other bacterial serine protease structures previously determined (James et al., 1978; Brayer et al., 1978b; Delbaere et al., 1979a). α-Lytic protease has the following distinctive features in common with the bacterial serine enzymes, Streptomyces griseus proteases A and B: an amino terminus that is exposed to solvent on the enzyme surface, a considerably shortened uranyl loop (residues 65 to 84), a major segment of polypeptide chain from the autolysis loop deleted (residues 144 to 155), a buried guanidinium group of Arg138 in an ion-pair bond with Asp194, and an altered conformation of the methionine loop (residues 168 to 182) relative to the pancreatic enzymes.At the present resolution, the members of the catalytic quartet (Ser214, Asp102, His57 and Ser195) adopt the conformation found in all members of the Gly-Asp-Ser-Gly-Gly serine protease family. The carboxylate of Asp102 is in a highly polar environment, as it is the recipient of four hydrogen bonds. The interaction between the Nε2 atom of the imidazole ring in His57 and Oγ atom of Ser195 is very weak (3·3 Å) and supports the concept that there is little, if any, enhanced nucleophilicity of the side-chain of Ser195 in the native enzyme.The molecular basis for the observed substrate specificity of α-lytic protease is clear from the distribution of amino acid side-chains in the neighborhood of the active site. An insertion of five residues at position 217, and the conformation of the side-chain of Met192 account for the fact that the specificity pocket can bind only small residues, such as Ala, Ser or Val.  相似文献   

11.
D-Serine dehydratase from Escherichia coli is a member of the β-family (fold-type II) of the pyridoxal 5′-phosphate-dependent enzymes, catalyzing the conversion of D-serine to pyruvate and ammonia. The crystal structure of monomeric D-serine dehydratase has been solved to 1.97 Å-resolution for an orthorhombic data set by molecular replacement. In addition, the structure was refined in a monoclinic data set to 1.55 Å resolution. The structure of DSD reveals a larger pyridoxal 5′-phosphate-binding domain and a smaller domain. The active site of DSD is very similar to those of the other members of the β-family. Lys118 forms the Schiff base to PLP, the cofactor phosphate group is liganded to a tetraglycine cluster Gly279-Gly283, and the 3-hydroxyl group of PLP is liganded to Asn170 and N1 to Thr424, respectively. In the closed conformation the movement of the small domain blocks the entrance to active site of DSD. The domain movement plays an important role in the formation of the substrate recognition site and the catalysis of the enzyme. Modeling of D-serine into the active site of DSD suggests that the hydroxyl group of D-serine is coordinated to the carboxyl group of Asp238. The carboxyl oxygen of D-serine is coordinated to the hydroxyl group of Ser167 and the amide group of Leu171 (O1), whereas the O2 of the carboxyl group of D-serine is hydrogen-bonded to the hydroxyl group of Ser167 and the amide group of Thr168. A catalytic mechanism very similar to that proposed for L-serine dehydratase is discussed.  相似文献   

12.
The crystal structure of zinc citrate [Zn(II) (C6H5O7)2·4NH4+] shows isolated zinc ions octahedrally coordinated to two equivalent citrates via a central hydroxyl, central carboxyl, and one terminal carboxyl from each citrate. The clusters are linked through hydrogen bonds to ammonium ions in the lattice. The structure is distinctly different from that of other divalent cation triply ionized citrate complexes, which are polymeric. Crystal data : space group P21/C, a = 8.784(3) Å, b = 13.499(4) Å, c = 9.083(3) Å, β = 113.4°(1), V = 988(1) Å3. Citrate has been identified as the low molecular weight ligand that complexes zinc in human milk; this may be of interest in relation to intestinal zinc absorption.  相似文献   

13.
Porcine pancreas kallikrein A has been crystallized in the presence of the small inhibitor benzamidine, yielding tetragonal crystals of space group P41212 containing two molecules per asymmetric unit. X-ray data up to 2·05 Å resolution have been collected using normal rotation anode as well as synchrotron radiation. The crystal structure of benzamidine-kallikrein has been determined using multiple isomorphous replacement techniques, and has subsequently been refined to a crystallographic R-value of 0·220 by applying a diagonal matrix least-squares energy constraint refinement procedure.Both crystallographically independent kallikrein molecules 1 and 2 are related by a non-integral screw axis and form open, heterologous “dimer” structures. The root-mean-square deviation of both molecules is 0·37 Å for all main-chain atoms. This value is above the estimated mean positional error of about 0·2 Å and reflects some significant conformational differences, especially at surface loops. The binding site of molecule 1 in the asymmetric unit is in contact with residues of molecule 2, whereas the binding site of the latter is free and accessible to the solvent. In both molecules the characteristic “kallikrein loop”, where the peptide chain of kallikrein A is cleaved, is only partially traceable. The carbohydrate attached to Asn95 in this loop, although detectable chemically, is not defined.A comparison of the refined structures of porcine kallikrein and bovine trypsin indicates spatial homology for these enzymes. The root-mean-square difference is 0·68 Å if we compare only main-chain atoms of internal segments. Remarkably large deviations are found in some external loops most of which surround the binding site and form a more compact rampart around it in kallikrein than in trypsin. This feature might explain the strongly reduced activity and accessibility of kallikrein towards large protein substrates and inhibitors (e.g. as shown by the model-building experiments on inhibitor complexes reported by Chen &; Bode. 1983).The conformation of the active site residues is very similar in both enzymes. Tyr99 of kallikrein, which is a leucyl residue in trypsin, protrudes into the binding site and interferes with the binding of peptide substrates (Chen &; Bode. 1983). The kallikrein specificity pocket is significantly enlarged compared with trypsin due to a longer peptide segment, 217 to 220, and to the unique outwards orientation of the carbonyl group of cis-Pro219. Further, the side-chain of Ser226 in porcine kallikrein, which is a glycyl residue in trypsin, partially covers Asp 189 at the bottom of the pocket. These features considerably affect the binding geometry and strength of binding of benzamidine.  相似文献   

14.
The high-resolution structure of bovine trypsin inhibited with DFP2 was determined by Stroud et al. (1971 and R. M. Stroud, L. M. Kay, A. Cooper &; R. E. Dickerson, Abstr. 8th Int. Congr. Biochem. 1970). The experiments reported here were designed to study the specific side-chain binding pocket of trypsin using benzamidine, which is a competitive, specific inhibitor of trypsin. High-resolution electron density syntheses and difference syntheses unambiguously identify the side-chain binding pocket, which normally recognizes and binds the side chains of arginine or lysine during proteolysis. Several important conformational differences in the protein structure are apparent between DIP- and BA-trypsins, and these are discussed with particular reference to inhibition, the binding of lysine and arginine, subsequent orientation of the target at the active site, and the enhancement of tryptic activity towards non-specific substrates seen on binding small alkyl amines or guanidines in the specific binding pocket.The BA-trypsin structure provides a good model for the binding of real substrate side chains to trypsin during catalysis, explaining the sharp trypsin specificity for lysine or arginine side chains (Weinstein &; Doolittle, 1972) and the lack of specificity for stereochemically different basic side chains. Benzamidine is shown to inhibit trypsin by steric interference with the inferred position of good substrates, even when they do not carry any side chain.Apart from the substitution of benzamidine and DIP, the most significant differences between DIP-trypsin and BA-trypsin involve complete repositioning of the side chain of Gln192, alterations in the side chains of Asp102, His57 and Ser195 at the active site, and changes in the solvent structure around this region. The carboxyl group of Asp189, which is responsible for trypsin specificity, shows no movement on binding benzamidine. The amidinium cation of benzamidine forms a salt bridge with Asp189 in BA-trypsin; a similar salt bridge can be constructed between the side chains of model substrates with lysyl or arginyl side chains and Aspl89. The γ-oxygen of Ser190 is displaced by a 120 ° rotation about its αβ bond on binding benzamidine and the binding pocket closes to sandwich the inhibitor ring between the peptide planes of 190–191 and 215–216. These contacts are presumably found in the enzyme-substrate complex with specific substrates.The active site structure at pH 8.0 is discussed with particular reference to the microscopic pKa values of Asp102 and His57, the pKa of the Asp-His system, and the mechanistic consequences of these assignments.  相似文献   

15.
Bis (difluoroboron - α - furilglyoximato) nickel (II), C20H12O8N4B2F4Ni, was prepared by cyclization of its hydrogen-bonded precursor with BF3·OEt2. The compound crystallizes in the space group P21/c with a = 11.162(2), b = 5.569(2), c = 19.527(3) Å, β = 100.08(1)°, U = 1195.1(3) Å3, and Z = 2. The structure was refined to an R value of 0.033 using 2371 unique reflections collected with a CAD4-SDP diffractometer system. Unlike the corresponding planar macrocyclic as well as hydrogen-bonded dimethylglyoximates, the title compound neither dimerizes not exhibits columnar stacked structure. The 14-member macrocycle is planar except the B atoms, and no metal-metal interactions are observed in this compound. The complexation and cyclization reactions were investigated using spectral data. The structure is compared with other macrocyclic complexes.  相似文献   

16.
《Inorganica chimica acta》1986,122(1):111-118
The title complex, prepared in 1 M NaOH, was crystallized from hot N,N-dimethylformamide/ ethanol solutions to give Na12[Ce(C6H2O2(SO3)2)4]· 9H2O·6DMF. The purple—brown crystals were examined by X-ray diffraction while inside quartz capillaries filled with DMF, (λmax 425 nm, ϵ 3664; λsh 520 nm, ϵ 2240) and belong to space group Pbca, Z=8 with a=21.846(4), b=17.348(2), c=43.103- (6) Å, V=16.335(7) Å3, Dc=1.693 gcmt−3, Do=1.725 g cmt−3. Diffractometer data were collected using Mo Kα radiation to 2θ=43o. For 7331 independent data with Fo2>3σ(Fo2) full matrix least squares refinement converged to unweighted and weighted R factors of 0.072 and 0.110, respectively, with a mixture of anisotropic and isotropic thermal parameters. The disordered DMF atom parameters were not refined. The structure consists of discrete monomeric Ce(C6H2S2O8)412− units with 12 Na+ counter cations and 10 H2O molecules (two with half occupancy), and 6 DMF molecules of solvation filling up spaces between cations and anions. Cerium(IV) is in a general position with a coordination polyhedron close to the trigonal-faced dodecahedron, D2d, with the angles between the two BAAB trapezoids of 2.3o and 3.7o. The average CeO(A) distance, 2.363(9) Å is longer than the average CeO(B) distance, 2.326(15)Å, with the reverse being true for one of the four tironato ligands. The average ring OCeO angle is 67.9(1)o. The cerium (IV) complex is found by cyclic voltammetry to undergo a quasi-reversible one-electron reduction (in strongly basic solution with excess tiron) with Ef=−497 mV vs. SCE, hence the ratio of the formation constants for tetrakis(tironato)cerate(IV) to that for tetrakis(tironato)cerate(III), KIV/KIII, is 1033. Characterization of other tiron salts is reported.  相似文献   

17.
Regioselective glycosylation of flavonoids cannot be easily achieved due to the presence of several hydroxyl groups in flavonoids. This hurdle could be overcome by employing uridine diphosphate-dependent glycosyltransferases (UGTs), which use nucleotide sugars as sugar donors and diverse compounds including flavonoids as sugar acceptors. Quercetin rhamnosides contain antiviral activity. Two quercetin diglycosides, quercetin 3-O-glucoside-7-O-rhamnoside and quercetin 3,7-O-bisrhamnoside, were synthesized using Escherichia coli expressing two UGTs. For the synthesis of quercetin 3-O-glucoside-7-O-rhamnoside, AtUGT78D2, which transfers glucose from UDP-glucose to the 3-hydroxyl group of quercetin, and AtUGT89C1, which transfers rhamnose from UDP-rhamnose to the 7-hydroxyl group of quercetin 3-O-glucoside, were transformed into E. coli. Using this approach, 67 mg/L of quercetin 3-O-glucoside-7-O-rhamnoside was synthesized. For the synthesis of quercetin 3,7-O-bisrhamnoside, AtUGT78D1, which transfers rhamnose to the 3-hydroxy group of quercetin, and AtUGT89C1 were used. The RHM2 gene from Arabidopsis thaliana was coexpressed to supply the sugar donor, UDP-rhamnose. E. coli expressing AtUGT78D1, AtUGT89C1, and RHM2 was used to obtain 67.4 mg/L of quercetin 3,7-O-bisrhamnoside.  相似文献   

18.
The gibberellin (GA) binding properties of a cytosol fraction from hypocotyls of cucumber (Cucumis sativus L. cv National Pickling) were examined using a DEAE filter paper assay, [3H]GA4, and over 20 GAs, GA derivatives and other growth regulators. The results demonstrate structural specificity of the binding protein for γ-lactonic C-19 GAs with a 3 β-hydroxyl and a C-6 carboxyl group. Additional hydroxylations of the A, C, or D ring of the ent-gibberellane skeleton and methylation of the C-6 carboxyl impede or abolish binding affinity. Bioassay data are generally supported by the in vitro results but significantly GA9 and GA36, both considered to be precursors of GA4 in cucumber, show no affinity for the binding protein. The results are discussed in relation to the active site of the putative GA4 receptor in cucumber.  相似文献   

19.
Phage lysozyme has catalytic activity similar to that of hen egg white lysozyme, but the amino acid sequences of the two enzymes are completely different.The binding to phage lysozyme of several saccharides including N-acetylglucosamine (GlcNAc), N-acetylmuramic acid (MurNAc) and (GlcNAc)3 have been determined crystallographically and shown to occupy the pronounced active site cleft. GlcNAc binds at a single location analogous to the C site of hen egg white lysozyme. MurNAc binds at the same site. (GlcNAc)3 clearly occupies sites B and C, but the binding in site A is ill-defined.Model building suggests that, with the enzyme in the conformation seen in the crystal structure, a saccharide in the normal chair configuration cannot be placed in site D without incurring unacceptable steric interference between sugar and protein. However, as with hen egg white lysozyme, the bad contacts can be avoided by assuming the saccharide to be in the sofa conformation. Also Asp20 in T4 lysozyme is located 3 Å from carbon C(1) of saccharide D, and is in a position to stabilize the developing positive charge on a carbonium ion intermediate. Prior genetic evidence had indicated that Asp20 is critically important for catalysis. This suggests that in phage lysozyme catalysis is promoted by a combination of steric and electronic effects, acting in concert, The enzyme shape favors the binding in site D of a saccharide with the geometry of the transition state, while Asp20 stabilizes the positive charge on the oxocarbonium ion of this intermediate. Tn phage lysozyme, the identity of the proton donor is uncertain. In contrast to hen egg white lysozyme, where Glu35 is 3 Å from the glycosidic DOE bond, and is in a non-polar environment, phage lysozyme has an ion pair, Glull … Arg145, 5 Å away from the glycosidic oxygen. Possibly Glull undergoes a conformational adjustment in the presence of bound substrate, and acts as the proton donor. Alternatively, the proton might come from a bound water molecule.  相似文献   

20.
Three-dimensional X-ray diffraction data were used to determine the crystal structure of sodium β-d-glucuronate monohydrate, a model system for investigating the factors involved in the binding of sodium ions to d-glucuronate residues of glycosaminoglycans. Crystals of the salt are monoclinic, space group P21, with a = 9.206(3) Å, b = 7.007(2) Å, c = 7.378(3) Å, β = 96.84(3)°, and Z = 2. Intensity data for 858 reflections were measured with an automated diffractometer. A trial structure, obtained by direct methods, was refined by least squares to R = 0.035. An outstanding feature of the crystal packing is the interaction of d-glucuronate anions with sodium ions. The sodium ion is coordinated to three symmetry-related d-glucuronate anions and to one water molecule. The d-glucuronate anion binds sodium cations through the three following sites: one that involves a carboxyl oxygen atom combined with ring oxygen O-5; one that includes a single carboxyl oxygen atom, and one composed of the O-3–O-4 pair of hydroxyl groups.  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