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.
Ramos L  Gil VM 《Carbohydrate research》2004,339(13):2225-2232
Multinuclear ((1)H, (13)C, (17)O, (31)P, (95)Mo, (183)W) magnetic resonance spectroscopy (1D and 2D) has been used to show that 6-phospho-d-gluconic acid forms three complexes with tungsten(VI) and six complexes with molybdenum(VI) in aqueous solution, depending on pH and concentration. Two isomeric 1:2 (metal-ligand) complexes are detected both with tungstate(VI) and molybdate(VI), having MO(2)(2+) centres and involving the carboxylate and the adjacent OH groups in addition to one 2:1 (metal-ligand) complex possessing a M(2)O(5)(2+) centre, with the ligand being coordinated by the carboxylate group and the three consecutive OH groups in positions 2, 3 and 4. Molybdate(VI) forms three additional species, which are not detected with tungstate. One of them is a 2:1 complex with a Mo(2)O(5)(2+) centre, with the ligand being tetradentate via O-3, O-4, O-5 and the phosphate group. The other two are 12:4 species, which can be seen as two 1:2 complexes bound together in a ring through two diphosphomolybdate moieties each derived from heptamolybdate by inclusion of two phosphate groups from the ligands.  相似文献   

2.
The 1H NMR relaxation effects produced by paramagnetic Cr(III) complexes on nucleoside 5'-mono- and -triphosphates in D2O solution at pH' = 3 were measured. The paramagnetic probes were [Cr(III)(H2O)6]3+, [Cr(III)(H2O)3(HATP)], [Cr(III)(H2O)3(HCTP)] and [Cr(III)(H2O)3(UTP)-, while the matrix nucleotides (0.1 M) were H2AMP, HIMP-, and H2ATP2-. For the aromatic base protons, the ratios of the transverse to longitudinal paramagnetic relaxation rates (R2p/R1p) for the [Cr(III)(H2O)6]3+/H2ATP2-, [Cr(III)(H2O)3(HATP)]/H2ATP2-, [Cr(III)(H2O)3(HCTP)]/H2ATP2 and [Cr(III)(H2O)3(UTP)]-/H2ATP2 systems were below 2.33 so the dipolar term predominates. For a given nucleotide, R1p for the purine H(8) signal was larger than for the H(2) signal with the [Cr(III)(H2O)6]3+ probe, while R1p for the H(2) signal was larger with all the other Cr(III) probes. Molecular mechanics computations on the [Cr(III)(H2O)4(HPP)(alpha,beta)], [Cr(III)(NH3)4(HPP)(alpha,beta)], [Co(III)(NH3)3(H2PPP)(alpha,beta,gamma)] and [Co(III)(NH3)4(HPP)(alpha,beta)] complexes gave calculated energy-minimized geometries in good agreement with those reported in crystal structures. The molecular mechanics force constants found were then used to calculate the geometry of the inner sphere [Cr(III)(H2O)6]3+ and [Cr(III)(H2O)3(HATP)(alpha,beta,gamma)] complexes as well as the structures of the outer sphere [Cr(III)(H2O)6]3(+)-(H2AMP) and [Cr(III)(H2O)6]-(HIMP)- species. The gas-phase structure of the [Cr(III)(H2O)3(HATP)(alpha,beta,gamma)] complex shows the existence of a hydrogen bond interaction between a water ligand and the adenine N(7)(O...N = 2.82 A). The structure is also stabilized by intramolecular hydrogen bonds involving the -O(2')H group and the adenine N(3) (O...N = 2.80 A) as well as phosphate oxygen atoms and a water molecule (O...O = 2.47 A). The metal center has an almost regular octahedral coordination geometry. The structures of the two outer-sphere species reveal that the phosphate group interacts strongly with the hexa-aquochromium probe. In both complexes, the nucleotides have a similar "anti" conformation around the N(9)-C(1') glycosidic bond. However, a very important difference characterizes the two structures. For the (HIMP)- complex, strong hydrogen bond interactions exist between one and two water ligands and the inosine N(7) and O(6) atoms, respectively (O...O = 2.63 A; O...N = 2.72, 2.70 A). For the H2AMP complex, the [Cr(III)(H2O)6]3+ cation does not interact with N(7) since it is far from the purine system. Hydrogen bonds occur between water ligands and phosphate oxygens. The Cr-H(8) and Cr-H(2) distances revealed by the energy-minimized geometries for the two outer sphere species were used to calculate the R1p values for the H(8) and H(2) signals for comparison with the observed R1p values: 0.92(c), 1.04(ob) (H(8)) and 0.06(c), 0.35(ob) (H(2)) for H2AMP; and 3.76(c), 4.53(ob) (H(8)) and 0.16(c), 0.77(ob) s-1 (H(2)) for HIMP-.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

3.
Primary rat microglia stimulated with either ATP or 2'- and 3'-O-(4-benzoylbenzoyl)-ATP (BzATP) release copious amounts of superoxide (O(2)(-)*). ATP and BzATP stimulate O(2)(-)* production through purinergic receptors, primarily the P2X(7) receptor. O(2)(-)* is produced through the activation of the NADPH oxidase. Although both p42/44 MAPK and p38 MAPK were activated rapidly in cells stimulated with BzATP, only pharmacological inhibition of p38 MAPK attenuated O(2)(-)* production. Furthermore, an inhibitor of phosphatidylinositol 3-kinase attenuated O(2)(-)* production to a greater extent than an inhibitor of p38 MAPK. Both ATP and BzATP stimulated microglia-induced cortical cell death indicating this pathway may contribute to neurodegeneration. Consistent with this hypothesis, P2X(7) receptor was specifically up-regulated around beta-amyloid plaques in a mouse model of Alzheimer's disease (Tg2576).  相似文献   

4.
The molybdenum cofactor (Moco)-containing enzymes are divided into three classes that are named after prototypical members of each family, viz. sulfite oxidase, DMSO reductase and xanthine oxidase. Functional or structural models have been prepared for these three prototypical enzymes: (i) The complex [MoO2(mnt)2]2- (mnt2- = 1,2-dicyanoethylenedithiolate) has been found to be able to oxidize hydrogen sulfite to HSO4- and is thus a functional model of sulfite oxidase. Kinetic and computational studies indicate that the reaction proceeds via attack of the substrate at one of the oxo ligands of the complex, rather than at the metal. (ii) The coordination geometries of the mono-oxo [Mo(VI)(O-Ser)(S2)2] entity (S2 = dithiolene moiety of molybdopterin) found in the crystal structure of R. sphaeroides DMSO reductase and the corresponding des-oxo Mo(IV) unit have been reproduced in the complexes [M(VI)O(OSiR3)(bdt)2] and [M(VI)O(OSiR3)(bdt)2] (M = Mo,W; bdt = benzene dithiolate). (iii) A facile route has been developed for the preparation of complexes containing a cis-Mo(VI)OS molybdenum oxo, sulfido moiety similar to that detected in the oxidized form of xanthine oxidase.  相似文献   

5.
Resonance Raman spectra were investigated for the sulfo and desulfo forms of cow's milk xanthine oxidase, with various visible excitation lines between 400 and 650 nm, and Mo(VI)-ligand vibrations were observed for the first time. The Mo(VI)=S stretch was identified at 474 and 462 cm(-1 )for the (32)S- and (34)S-sulfo forms, respectively, but was absent in the reduced state and in the desulfo form. The Mo(VI)=O stretch was weakly observed at 899 cm(-1 )for the sulfo form and shifted to 892 cm(-1) with very weak intensity for the dioxo desulfo form. In measurements of an excitation profile, the two bands at 474 and 899 cm(-1) showed maximum intensity at similar excitation wavelengths, suggesting that the Raman intensity of the metal-ligand modes is due to the Mo(VI)<--S charge transfer transition, and that this is the origin of the intrinsically weak features of the Mo(VI)-ligand Raman bands. When the sulfo form was regenerated from the desulfo form, the 899 cm(-1) band reappeared. However, the band at 899 cm(-1) showed no frequency shift when regeneration was conducted in H(2)(18)O, or after several turnovers in the presence of xanthine in H(2)(18)O. When the sulfo form was reduced and reoxidized in H(2)(18)O buffer, the 899 cm(-1) band reappeared without any frequency shift. These observations suggest that the oxo oxygen in the Mo center of xanthine oxidase is not labile. Low-frequency vibrations of the Mo center were observed together with those of the Fe(2)S(2) center with some overlaps, while FAD modes were observed clearly. The absence of dithiolene modes in XO is in contrast to the Mo(VI) centers of DMSO reductase and sulfite oxidase.  相似文献   

6.
The stereospecific formation and absolute configuration of R-homocitrate coordinated FeMo-co in nitrogenase was mimicked through the structural analyses of a collection of enantiomeric and mesomeric mandelato molybdenum complexes, i.e., (NH(4))(2)[Mo(Delta)O(2)(R-mand)(2)]x3H(2)O (1a), (NH(4))(2)[Mo(Lambda)O(2)(S-mand)(2)]x3H(2)O (1b), (NH(4))(4)[Mo(Delta)O(2)(RS-mand)(2)][Mo(Lambda)O(2)(RS-mand)(2)]x8H(2)O (2), (NH(4))(2)[W(Delta)O(2)(R-mand)(2)]x2H(2)O (3a), (NH(4))(2)[W(Lambda)O(2)(S-mand)(2)]x2H(2)O (3b) (H(2)mand=mandelic acid, C(8)H(8)O(3)), which have been characterized by elemental analyses, optical rotation, circular dichroism, IR, NMR spectroscopes and X-ray single crystal studies. The R and S chiral mandelic acids induce the formations of the enantiomeric pair of chiral complexes, which are supported by the characterizations of optical rotation and circular dichroism. The configuration of the resulted metal center could be assigned as Delta or Lambda. While the RS racemic reagent yields only mesomeric compound. The Delta(R,R)-complexes 1a and 3a are enantiomers of Lambda(S,S)-1b and 3b, respectively. Of the five complexes, Mo and W atoms are all hexa-coordinated by two cis-oxo groups and two bidentate mandelate ligands through the deprotonated alpha-alkoxyl and alpha-carboxyl groups, forming a stable five-membered chelated rings. The average Mo(VI)-O bond distances with alpha-alkoxyl and alpha-carboxyl are 1.944 and 2.210 A, respectively. Further comparison indicates that bonds of alpha-alkoxyl groups in the hydroxycarboxylato molybdenum complexes are much sensitive to the change in the oxidation state of molybdenum, which support the possible Mo activation model in FeMo-co through the protonation and cleavage of alpha-alkoxyl group in homocitrate ligand.  相似文献   

7.
The concentration dependence of the chemical shifts for the protons H-2, H-8 and H-1' of ATP has been measured in D2O at 27 degrees C under several degrees of protonation in the pD range from 1.5 to 8.4. The results at pD greater than 4.5 are consistent with the isodesmic model of indefinite noncooperative stacking, while those at pD less than 4.5 indicate a preference for the formation of dimeric stacks. The stacking tendency follows the series, ATP4- (K = 1.3 M-1) less than D(ATP)3- (2.1 M-1) less than 1:1 ratio of D(ATP)3-/D2(ATP)-2- (6.0 M-1) much less than D2(ATP)2- (approximately 200 M-1) much greater than D3(ATP)- (K approximately less than 17 M-1) (for reasons of comparison all constants are expressed in the isodesmic model). These results are compared with previous data for adenosine [Ado (K = 15 M-1) greater than 1:1 ratio of Ado/D(Ado)+ (6.0 M-1) greater than D(Ado)+ (0.9 M-1)] and AMP [AMP2- (K = 2.1 M-1) less than D(AMP)- (3.4 M-1) less than 1:1 ratio of D(AMP)-/D2(AMP) +/- (5.6 M-1) greater than D2(AMP) +/- (approximately equal to 2 M-1) greater than D3(AMP)+ (K less than or equal to 1 M-1)] to facilitate the interpretation of the results for the ATP systems. Stack formation of H2(ATP)2- is clearly favored by additional ionic interactions; this is confirmed by measuring via potentiometric pH titrations the acidity constants of H2(ATP)2- in solutions containing different concentrations of ATP. It is suggested that in the [H2(ATP)]4-(2) dimer intermolecular ion pairs (and hydrogen bonds) are formed between the H+(N-1) site of one H2(ATP)2- and the gamma-P(OH)(O)-2 group of the other; in this way (a) the stack is further stabilized, and (b) the positive charges at the adenine residues are compensated (otherwise repulsion would occur as is evident from the adenosine systems). A detailed structure for the [H2(ATP)4-(2) dimer is proposed and some implications of the described stacking properties of ATP for biological systems are indicated.  相似文献   

8.
Several lines of functional evidence have shown that anion binding to a nonsynergistic site is a prerequisite for the anion-mediated iron release mechanism of transferrins. We report here structural evidence of the location of sulfate anion binding sites of the ovotransferrin N-lobe via the 1.90 A resolution apo crystal structure. The crystals were grown in an ammonium sulfate solution and belonged to space group P6(3)22 with the following unit cell dimensions: a = b = 125.17 A and c = 87.26 A. The structural determination was performed by isomorphous replacement, using Pt and Au derivatives. The structure refinement gave an R-factor of 0.187 in the resolution range of 7.0-1.90 A for the final model. From the electron density map, the existence of four bound SO(4)(2)(-) anions was detected. Three of them that exhibited reasonably low B-factors were all located in the opened interdomain cleft (sites 1-3). In site 1, the bound anion directly interacts with an Fe(3+)-coordinating ligand; SO(4) O1 and SO(4) O3 form hydrogen bonds with His250 NE2. Oxygen atom O3 of the same sulfate anion makes a hydrogen bond with Ser91 OG in a hinge strand. The sulfate anion in site 2 partially occupies the synergistic anion binding sites; SO(4) O2 and SO(4) O3 are hydrogen bonded to Arg121 NE and NH2, respectively, both of which are consensus anchor groups for CO(3)(2)(-) anion in holotransferrins. The former oxygen atom of SO(4)(2)(-) is also hydrogen bonded to Ser122 N, which forms a hydrogen bond with Fe(3+)-coordinating ligand Asp60 OD2 in holotransferrins. Some of the SO(4)(2)(-) oxygen atoms in sites 1 and 2 interact indirectly through H(2)O molecules with functionally important protein groups, such as the other Fe(3+)-coordinating ligands, Tyr92 OH and Tyr191 OH, and a dilysine trigger group, Lys209 NZ. In site 3, SO(4) O1 and SO(4) O4 form hydrogen bonds with Ser192 OG and Tyr191 N, respectively, and SO(4) O2 forms hydrogen bonds with Ser192 N and Ser192 OG. These structural data are consistent with the view that the anion bindings to the interdomain cleft, especially to sites 1 and 2, play crucial roles in the domain opening and synergistic carbonate anion release in the iron release mechanism of the ovotransferrin N-lobe.  相似文献   

9.
The preparation of platinum(II) complexes containing L-serine using K(2)[PtCl(4)] and KI as raw materials was undertaken. The cis-trans isomer ratio of the complexes in the reaction mixture differed significantly depending on whether KI was present or absent in the reaction mixture. One of the two [Pt(L-ser-N,O)(2)] complexes (L-ser=L-serinate anion) prepared using KI crystallizes in the monoclinic space group P2(1)2(1)2(1) with unit cell dimensions a=8.710(2) A, b=9.773(3) A, c=11.355(3) A, Z=4. The crystal data revealed that this complex has a cis configuration. The other [Pt(L-ser-N,O)(2)] complex also crystallizes in the monoclinic space group P2(1)2(1)2(1) with unit cell dimensions a=7.0190(9) A, b=7.7445(6) A, c=20.946(2) A, Z=4. The crystal data revealed that this complex has a trans configuration. The 195Pt NMR chemical shifts of trans-[Pt(L-ser-N,O)(2)] and cis-[Pt(L-ser-N,O)(2)] complexes are -1632 and -1832 ppm, respectively. 195Pt NMR and HPLC measurements were conducted to monitor the reactions of the two [Pt(L-ser-N,O)(2)] complexes with HCl. Both 195Pt NMR and HPLC showed that the reactivities of cis- and trans-[Pt(L-ser-N,O)(2)] toward HCl are different: coordinated carboxyl oxygen atoms of trans-[Pt(L-ser-N,O)(2)] were detached faster than those for cis-[Pt(L-ser-N,O)(2)].  相似文献   

10.
A novel bioluminescent enzymatic cycling assay for ATP and AMP with concomitant use of firefly luciferase and pyruvate orthophosphate dikinase (PPDK) was developed. In this system, AMP and pyrophosphate produced from ATP by firefly luciferase were converted back into ATP by PPDK. This resulted in constant luminescence once the stable phase had been reached. Background luminescence of the reagent was reduced with adenosine phosphate deaminase by degrading ATP and AMP in the reagent. The maximum recycling ratio calculated from the integrated luminescence value was 2.64 cycles/min. The measurable ranges for ATP and AMP were equal and were between 4 x 10(-13) and 4 x 10(-17) mol/assay. The amount of yeast RNA could be estimated in the range of 1 x 10(-8) to 1 x 10(-12) g/assay by estimating the amount of AMP resulting from the degradation of RNA with nuclease P1. Various food samples were subjected to measurement of the amount of ATP + AMP + RNA to provide an index for hygiene monitoring. For beef extract, sensitivity was improved by more than 20 million compared to the previous methods relying only on the amount of ATP as an index.  相似文献   

11.
ATP hydrolysis in the presence of polyoxomolybdates at pH levels of 6, 4, and 2 has been investigated with a help of high pressure liquid chromatography (HPLC) analyses, 31P- and 1H NMR measurements, and isothermal titration calorimetry (ITC). The polyoxomolybdates-induced ATP-hydrolysis proceeded satisfactorily in pH < 6 media at 20 degrees C with an optimum pH level of 4, while it was significantly depressed at low temperature of < or = 5 degrees C. At pH levels of 6 and 4, ADP was a main product, and the involvement of [(PO4)2Mo5O15](6-)-like ATP-molybdate complex as an intermediate was implied. At pH 2 ATP was decomposed to AMP with small generation of ADP through the formation of the ATP-molybdate complex isostructural with [(O3POPO3)Mo6O18(H2O)4]4- as an intermediate. The ITC result at pH 4 showed an occurrence of two types of the exothermic binding reactions between molybdate and ATP with binding constants (K) of 6.61x10(4) and 9.40x10(3) M(-1) and molar enthalpy values (deltaH) of -6.32x10(4) and -4.73x10(3) J mol(-1), respectively. Together with the results of 1H NMR measurements, it is deduced that the molybdates interact with not only phosphate sites in the ATP side-chain, but also adenine-ring with an accompanying aggregation of molybdates at pH 4.  相似文献   

12.
Regulation of human neutrophil functions by adenine nucleotides   总被引:3,自引:0,他引:3  
Previous work has shown that platelet-derived adenine nucleotides modulate neutrophil superoxide anion (O2-) generation. Additional studies were undertaken to characterize the effects of authentic adenosine (ADO) and its nucleotide derivatives on the inflammatory functions of human neutrophils. Stimulus-specific inhibition of neutrophil O2- generation by ADO in response to FMLP was verified. In addition, the ability of ATP, ADP, and AMP to limit neutrophil O2- generation induced by FMLP (0.2 to 0.5 microM) was demonstrated. The concentration producing 50% inhibition for nucleotide inhibition of neutrophil O2- generation was in the rank order of ADO (0.1 microM) less than AMP (0.5 microM) less than ADP less than or equal to ATP (5 microM). Guanine and inosine nucleotides (0.01 to 100 microM) did not inhibit FMLP-stimulated neutrophil O2- generation. Neutrophil degranulation in response to FMLP was only modestly inhibited by adenine nucleotides and ADO. Adenosine and ADP failed to affect chemotaxis of neutrophils stimulated with FMLP. The inability of non-metabolizable analogs to mimic the inhibitory effects of authentic ATP or ADP on the neutrophil O2- response suggested that metabolism of added nucleotides is necessary for their effectiveness. Both TLC and HPLC confirmed that ATP and ADP were converted to AMP and ADO after their incubation with unstimulated or FMLP-activated neutrophils. The addition of adenosine deaminase to neutrophil reaction mixtures in which conversion of added nucleotides was apparent removed detectable ADO but failed to completely abrogate the inhibition of neutrophil O2- generation by accumulated AMP. The kinetics of inhibition of FMLP-induced neutrophil O2- generation by ATP and ADP also indicated that conversion of these nucleotides to ADO and/or AMP may be essential for their ability to reduce neutrophil responses.  相似文献   

13.
The application of Mo(VI) complexes as anti-diabetic agents is a subject of considerable recent interest. The stability and speciation of [Mo(VI)O(4)](2-) and three analogs of known anti-diabetic V(IV) complexes ([Mo(VI)O(2)L(2)]; where LH=2,4-pentanedione, l-cysteine ethyl ester or N,N-diethyldithiocarbamic acid) in natural and simulated biological fluids (including blood and its components, cell culture media, and artificial digestion systems) were studied using MoK-edge XANES (X-ray absorption near-edge structure) spectroscopy of freeze-dried samples at 20K. All of the studied [MoO(2)L(2)] complexes decomposed extensively under simulated gastric and intestinal digestion conditions (3 h at 310 K), as well as in blood plasma or in cell culture medium (24 h at 310 K). The reaction products of [MoO(4)](2-) and [MoO(2)L(2)] with biological fluids could be satisfactorily modelled (using multiple linear regression analyses) as mixtures of tetrahedral and octahedral Mo(VI) species (with O-donor ligands) in various ratios, which were dependent on the nature of the medium rather than that of the initial Mo(VI) compounds. Red blood cells take up Mo(VI) predominantly in the form of [MoO(4)](2-). Implications of these results to the development of Mo(VI)-based anti-diabetics and to the mechanisms of natural uptake and metabolism of Mo(VI) are discussed.  相似文献   

14.
The inhibiting effect of adenosine, AMP, ADP, ATP, gamma-thio ATP (I), beta,gamma-imine ATP (II), beta,gamma-methylene ATP (III), P1,P3-di(adenosine-5') triphosphate (IV), P1,P4-di(adenosine-5') tetraphosphate (V) and adenosine 5'-tetraphosphate (VI) on the first step of the T4 RNA ligase reaction was studied. All the compounds tested, with the exception of adenosine, appeared to be competitive inhibitors of the first step of the enzymatic reaction. The inhibition constants (Ki) for the ATP analogs were determined. The data obtained suggest that the efficiency of inhibition depends on the number of phosphate groups and on the structure of ATP analogs. All the compounds under study (I-VI), except for AMP and ADP, form covalent AMP-RNA ligase complexes.  相似文献   

15.
The Mg(2+)-dependent adenylylation of the T4 DNA and RNA ligases was studied in the absence of a DNA substrate using transient optical absorbance and fluorescence spectroscopy. The concentrations of Mg(2+), ATP, and pyrophosphate were systematically varied, and the results led to the conclusion that the nucleotidyl transfer proceeds according to a two-metal ion mechanism. According to this mechanism, only the di-magnesium-coordinated form Mg(2)ATP(0) reacts with the enzyme forming the covalent complex E.AMP. The reverse reaction (ATP synthesis) occurs between the mono-magnesium-coordinated pyrophosphate form MgP(2)O(7)(2-) and the enzyme.MgAMP complex. The nucleotide binding rate decreases in the sequence ATP(4-) > MgATP(2-) > Mg(2)ATP(0), indicating that the formation of the non-covalent enzyme.nucleotide complex is driven by electrostatic interactions. T4 DNA ligase shows notably higher rates of ATP binding and of subsequent adenylylation compared with RNA ligase, in part because it decreases the K(d) of Mg(2+) for the enzyme-bound Mg(2)ATP(0) more than 10-fold. To elucidate the role of Mg(2+) in the nucleotidyl transfer catalyzed by T4 DNA and RNA ligases, we propose a transition state configuration, in which the catalytic Mg(2+) ion coordinates to both reacting nucleophiles: the lysyl moiety of the enzyme that forms the phosphoramidate bond and the alpha-beta-bridging oxygen of ATP.  相似文献   

16.
Bardiya N  Hwang YW  Bae JH 《Anaerobe》2004,10(1):7-11
Thiosulfate (S(2)O(3)(2-)) contained in the media for autotrophic Cr (VI) reduction was found to interfere with Cr (VI) measurement following the 1,5-diphenylcarbazide (DPC) method. The interference was confirmed at several abiotic and biotic conditions, and was influenced by S(2)O(3)(2-) concentration, pH, and the media components. At neutral to alkaline pH, 500 mg/L of S(2)O(3)(2-) did not cause interference, while 4 mg/L of S(2)O(3)(2-) resulted in the interference at pH 2.0. Atomic absorption spectrophotometry could be an alternative method when the interference by S(2)O(3)(2-) is expected.  相似文献   

17.
Song JF  He P  Guo W 《Analytical biochemistry》2002,304(2):212-219
The polarographic catalytic wave of vitamin P in the presence of persulfate was studied by linear potential scan polarography and cyclic voltammetry. Vitamin P yielded a single reduction wave in acidic aqueous solution, which was ascribed to a 2e(-), 2H(+) reduction of the carbonyl group in the C-4 position. Actually, the carbonyl group C=O first underwent a 1e(-), 1H(+) reduction to form a neutral free radical, and the further 1e(-), 1H(+) reduction of the free radical was simultaneous with its following chemical reactions. When S(2)O(2-)(8) was present, the free radical of vitamin P was oxidized by both S(2)O(2-)(8) and its reduction intermediate, the sulfate radical anion SO(*-)(4), to regenerate the original, which resulted in the production of a polarographic catalytic wave of vitamin P. Based on this catalytic wave, a novel method for the determination of vitamin P was proposed. In 0.02 M tartaric acid-sodium tartrate (pH 3.3) buffer containing 5.0 x 10(-3) M K(2)S(2)O(8), the peak potential of the catalytic wave was -1.42 V (vs SCE) and the peak current was rectilinear to the vitamin P concentration in the range of 8.0 x 10(-9)-1.0 x 10(-6) M (r = 0.9994, n = 13). The catalytic wave of 2.0 x 10(-7) M vitamin P enhanced the polarographic current 70 times compared with the corresponding reduction wave. The detection limit was 2.0 x 10(-9) M, and the relative standard deviation at the 2.0 x 10(-7) M level was 0.7% (n = 15). The proposed method was used for the determination of vitamin P content in the pharmaceutical preparation of tablets and the medicinal plant Sophora japonica L. without previous separation.  相似文献   

18.
The O-dealkylation of 7-alkoxyresorufins to the highly fluorescent compound, resorufin (7-hydroxyphenoxazone), provides a rapid, sensitive, and convenient assay of certain forms of liver microsomal cytochrome P450. The results of this study indicate that NADPH-cytochrome P450 reductase catalyzes the reduction of resorufin (and the 7-alkoxyresorufins) to a colorless, nonfluorescent compound(s). The reduction of resorufin by NADPH-cytochrome P450 reductase was supported by NADPH but not NADH, and was not inhibited by dicumarol, which established that the reaction was not catalyzed by contaminating DT-diaphorase (NAD[P]H-quinone oxidoreductase). In addition to the rate of reduction, the extent of reduction of resorufin was dependent on the concentration of NADPH-cytochrome P450 reductase. The maintenance of steady-state levels of reduced resorufin required the continuous oxidation of NADPH, during which molecular O2 was consumed. When NADPH was completely consumed, the spectroscopic and fluorescent properties of resorufin were fully restored. These results indicate that the reduction of resorufin by NADPH-cytochrome P450 reductase initiates a redox cycling reaction. Stoichiometric measurements revealed of 1:1:1 relationship between the amount of NADPH and O2 consumed and the amount of H2O2 formed (measured fluorometrically). The amount of O2 consumed during the redox cycling of resorufin decreased approximately 50% in the presence of catalase, whereas the rate of O2 consumption decreased in the presence of superoxide dismutase. These results suggest that, during the reoxidation of reduced resorufin, O2 is converted to H2O2 via superoxide anion. Experiments with acetylated cytochrome c further implicated superoxide anion as an intermediate in the reduction of O2 to H2O2. However, the ability of reduced resorufin to reduce acetylated cytochrome c directly (i.e., without first reducing O2 to superoxide anion) precluded quantitative measurements of superoxide anion formation. Superoxide dismutase, but not catalase, increased the steady-state level of reduced resorufin and considerably delayed its reoxidation. This indicates that superoxide anion is not only capable of reoxidizing reduced resorufin, but is considerably more effective than molecular O2 in this regard. Overall, these results suggest that NADPH-cytochrome P450 reductase catalyzes the one-electron reduction of resorufin (probably to the corresponding semiquinoneimine radical) which can either undergo a second, one-electron reduction (presumably to the corresponding dihydroquinoneimine) or a one-electron oxidation by reducing molecular O2 to superoxide anion.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
S P Harnett  G Lowe  G Tansley 《Biochemistry》1985,24(12):2908-2915
The activation of L-phenylalanine by yeast phenylalanyl-tRNA synthetase using adenosine 5'-[(S)-alpha-17O,alpha,alpha-18O2]triphosphate is shown to proceed with inversion of configuration at P alpha of ATP. This observation taken together with the lack of positional isotope exchange when adenosine 5'-[beta,beta-18O2]triphosphate is incubated with the enzyme in the absence of phenylalanine and in the presence of the competitive inhibitor phenylalaninol indicates that activation of phenylalanine occurs by a direct "in-line" adenylyl-transfer reaction. In the presence of Zn2+, yeast phenylalanyl-tRNA synthetase also catalyzes the phenylalanine-dependent hydrolysis of ATP to AMP and the synthesis of P1,P4-bis(5'-adenosyl) tetraphosphate (Ap4A). With adenosine 5'-[(S)-alpha-17O,alpha,alpha-18O2]triphosphate, the formation of AMP and Ap4A is shown to occur with inversion and retention of configuration, respectively. It is concluded that phenylalanyl adenylate is an intermediate in both processes, Zn2+ promoting AMP formation by hydrolytic cleavage of the C-O bond and Ap4A formation by displacement at phosphorus of phenylalanine by ATP.  相似文献   

20.
Complexes formed in aqueous solution between cisplatin or hydrolysis species and 5′ adenosine monophosphate (AMP) or 5′ adenosine triphosphate (ATP), the latter with and without chloride ions, have been determined using 195Pt, 31P, 13C and 1H NMR. The present results lead to the conclusion that the only monodentate complexes with AMP are cis-Pt(NH3)2(AMP-N7)Cl at acid pH and cis-Pt(NH3)2(AMP-N7)OH at neutral and basic pH. Other bidentate complexes were identified as cis-Pt(NH3)2(AMP-N7)2 and cis-Pt(NH3)2(AMP-N7)(AMP-PO). Also discussed herein are the binding of platinum to the phosphate group Pγ with ATP and at acid pH, and the formation of the [cis-Pt(NH3)2(ATP-N7)H2O]+ complex. In neutral and basic pH ranges, the phosphate moiety of ATP is the most reactive site. In the presence of an excess of chloride ions, the complexation rates between the ATP and the cisplatin are decreased. Furthermore, in the experimental conditions used neither the ATP nor the AMP have shown binding to N1.  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