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 390 毫秒
1.
Tributyl phosphate (TBP) and other alkyl phosphates represent a class of persistent organophosphorus compounds of widespread use. Biodegradation of the phosphotriesters is postulated to occur through sequential hydrolytic cleavages via the phosphodiester and monoester intermediates to alcohol and inorganic phosphate (Pi). Immobilized cells of aCitrobacter sp. liberated Pi upon challenge with TBP but the reaction was short-lived. In vitro studies with purified phosphomonoesterase (phosphatase) used31P nuclear magnetic resonance to demonstrate Pi transfer onto ethanol (phosphotransferase activity). This suggested that in vivo the onset of a futile phosphohydrolytic and transphosphorylation cycle would limit the extent of phosphate production. A mutant deficient in the transphosphorylating phosphomonoesterase showed an extended release of Pi under challenge with TBP that was not subject to the complete and premature reaction termination that precluded application of the parent strain to possible industrial processes for alkyl phosphate biodegradation.  相似文献   

2.
The reaction of [IrCl(dmso)3] with trisphosphinomethylborato ligand Li(THF){PhB(CH2PiPr2)3} at room temperature results in intramolecular C-H activation of one of the iPr substituents affording two diastereomers of cyclometalated iridium(III) complex [Ir(H)(dmso){PhB(CH2PiPr2)2(CH2PiPrCHMeCH2)}] (1) in high yield in approximately equimolar ratio. NMR spectroscopic characterization indicates that only the diastereomers with the hydride ligand in cis position with respect to the metalacyclic phosphorous atom are formed as confirmed by single crystal X-ray diffraction. Facile ring opening with H2 at room temperature gives dihydride [Ir(H)2(dmso){PhB(CH2PiPr2)3}] (2). However, C-H activation of benzene was not observed.  相似文献   

3.
We have studied crystallization behavior in M21O-P2O5-V2O5-CaO and M21O-P2O5-V2O5-CaO-YF3 flux systems (M1= Na, K) for discrete ratios P/V = 0.04-0.14 and M1/(P+V) = 0.7, 1.0, 1.2, 1.4, w(CaO) = w(YF3) = 5 wt.%. We established the conditions for the formation of phosphate vanadates based on apatite, whitlockite, NaCaPO4, KCaY(PO4)2, and YPO4 structural types.  相似文献   

4.
Reaction of HCCUr (Ur = uracil) with [RhCl(PiPr3)2] results in the formation of the vinylidene complex [RhCl(PiPr3)2(CC{H}Ur)]. In the solid state this complex forms a hydrogen bonded network which consists of complementary interactions between uracil groups on neighbouring rhodium complexes and with the methanol of crystallisation. The η2-alkyne complexes [RhCl(PiPr3)22-PhCCUr)] and [Rh(η5-C5H5)(PiPr3)(η2-PhCCUr)] have also been prepared. In contrast to the behaviour of [Rh(η5-C5H5)(PiPr3)(η2-PhCCUr)], [RhCl(PiPr3)22-PhCCUr)] shows little evidence for the formation of hydrogen bonded aggregates in solution. The difference in behaviour between the two species is rationalised on the basis of steric effects.  相似文献   

5.

Equilibrium reactions of iron(III) with phosphate were studied spectrophotometrically by UV-Vis in the pH range of ~ 1.0-2.20. The STAR-94 Program was used to determine the number of absorbing species as well as the stoichiometries and formation constants of the complex species. Some literature values were further confirmed and new values of different stoichiometries were obtained. The kinetics and mechanism of Fe(III) with phosphate were studied in acidic medium. The reactive phosphate species were found to be only H3PO4 and H2PO? 4 and for Fe(III) were only Fe3+, FeOH2+ and Fe(OH)+ 2. The observed rate constants were pH as well as Tphos (total concentration of phosphate) dependent, i.e. Kobs,i = A i + B i Tphos + C i T2 phos (at a given pH).  相似文献   

6.
The kinetics of 5′-ATP hydrolysis catalyzed by the Cu2+ ion has been investigated by HPLC in the pH range 5.6–7.8 at 25°C. Two series of experiments differing in the initial [Cu · ATP]0 (1: 1) concentration have been carried out. The reaction was being conducted up to ≈40% ATP conversion. The (CuATP2?)2OH??ub;DOH??ub; complex, which consists of two monomeric Cy(CuATP2?) molecules (in which the N7 atom and the γ-phosphate group are coordinated to Cu2+), is responsible for the formation of CuADP? + Pi (Pi is an inorganic phosphate). The highest possible DOH? concentration at a given pH is reached at the initial stage of hydrolysis. The pH value at which the highest initial rate of ADP formation is reached (pHmax (w 0, ADP)) decreases as the D concentration increases. At pH > pHmax, the decrease in the ADP formation rate in the course of the processes is pH-independent and, once an ATP conversion of 20–26% is reached, hydrolysis proceeds in a steady-state regime such that ADP and AMP form from ATP by parallel reactions. The participation of the OH? ion in the catalysis of the formation of hydrolysis intermediates is considered.  相似文献   

7.
Oxidative addition of the P–P single bond of an ortho-carborane-derived 1,2-diphosphetane (1,2-C2(PMes)2B10H10) (Mes = 2,4,6-Me3C6H2) to cobalt(−i) and nickel(0) sources affords the first heteroleptic complexes of a carborane-bridged bis(phosphanido) ligand. The complexes also incorporate labile ligands suitable for further functionalisation. Thus, the cobalt(i) complex [K([18]crown-6)][Co{1,2-(PMes)2C2B10H10}(cod)] (cod = 1,5-cyclooctadiene) bearing a labile cyclooctadiene ligand undergoes facile ligand exchange reactions with isonitriles and tert-butyl phosphaalkyne with retention of the bis(phosphanido) ligand. However, in the reaction with one equivalent of P4, the electron-rich bis(phosphanido) moiety abstracts a single phosphorus atom with formation of a new P3 chain, while the remaining three P atoms derived from P4 form an η3-coordinating cyclo-P3 ligand. In contrast, when the same reaction is performed with two equivalents of the cobalt(i) complex, a dinuclear product is formed which features an unusual P4 chain in its molecular structure.

Cobalt and nickel complexes of a new carborane-substituted bis(phosphanido) ligand have been prepared. Reaction of the cobalt complex with white phosphorus (P4) results in a remarkable fragmentation of P4 into P3 and P1 units.  相似文献   

8.
This contribution reports on a new family of NiII pincer complexes featuring phosphinite and functional imidazolyl arms. The proligands RPIMCHOPR′ react at room temperature with NiII precursors to give the corresponding complexes [(RPIMCOPR′)NiBr], where RPIMCOPRPCP‐{2‐(R′2PO),6‐(R2PC3H2N2)C6H3}, R=iPr, R′=iPr ( 3 b , 84 %) or Ph ( 3 c , 45 %). Selective N‐methylation of the imidazole imine moiety in 3 b by MeOTf (OTf=OSO2CF3) gave the corresponding imidazoliophosphine [(iPrPIMIOCOPiPr)NiBr][OTf], 4 b , in 89 % yield (iPrPIMIOCOPiPrPCP‐{2‐(iPr2PO),6‐(iPr2PC4H5N2)C6H3}). Treating 4 b with NaOEt led to the NHC derivative [(NHCCOPiPr)NiBr], 5 b , in 47 % yield (NHCCOPiPrPCC‐{2‐(iPr2PO),6‐(C4H5N2)C6H3)}). The bromo derivatives 3–5 were then treated with AgOTf in acetonitrile to give the corresponding cationic species [(RPIMCOPR)Ni(MeCN)][OTf] [R=Ph, 6 a (89 %) or iPr, 6 b (90 %)], [(RPIMIOCOPR)Ni(MeCN)][OTf]2 [R=Ph, 7 a (79 %) or iPr, 7 b (88 %)], and [(NHCCOPR)Ni(MeCN)][OTf] [R=Ph, 8 a (85 %) or iPr, 8 b (84 %)]. All new complexes have been characterized by NMR and IR spectroscopy, whereas 3 b , 3 c , 5 b , 6 b , and 8 a were also subjected to X‐ray diffraction studies. The acetonitrile adducts 6 – 8 were further studied by using various theoretical analysis tools. In the presence of excess nitrile and amine, the cationic acetonitrile adducts 6 – 8 catalyze hydroamination of nitriles to give unsymmetrical amidines with catalytic turnover numbers of up to 95.  相似文献   

9.
The inhibitory constants of complexes of trypsin and its soluble or immobilized inhibitors were determined from volumes in which trypsin emerged from the column of its immobilized inhibitor (p-aminobenzamidine coupled through hexamethylenediamine to hydroxyalkyl methacrylate gel, Spheron) eluted by solutions of soluble trypsin inhibitors (benzylamine, benzoyl-L-arginine, N-butylamine, benzamidine, and p-aminobenzamidine). The values of constants obtained by affinity chromatography in the zonal and frontal analysis arrangement were in good agreement and in accordance with data obtained kinetically. The plot of l/(Vi-V0) versus l/KI (determined by zonal analysis) or of Vi versus KI(V-Vi) (determined by frontal analysis) for identical concentrations of various inhibitors was linear. The fact that the dissociation constant of the complex of trypsin and immobilized p-aminobenzamidine (1.6-3.7 x 106 M) is lower than the dissociation constant of the complex of trypsin and free p-aminobenzamidine (1.9 x 10-5 M) seems to indicate possibilities of nonspecific adsorption in the binding of trypsin to p-aminobenzamidine-NH2-Spheron. Submitted as RNDr.-Thesis at the Faculty of Natural Sciences, Charles University, Prague, May 1977.  相似文献   

10.
In the Bi2O3-SiO2-V25++O5 system, single crystal solid solutions of the sillenite family of the general composition Bi24(Bi,Si,V)2O40 are obtained by a hydrothermal method and for the first time characterized by neutron and X-ray diffraction analysis. The tetrahedral position is found to contain vanadium ions with different formal charges (V4+ and V5+) responsible for green and orange colors, respectively, of the samples. For the first time, for some sillenites of this system dissymmetrization of the structure (a transition from the I23 space group into P23) is revealed, which is caused by the presence of several atoms in one crystallographic position and also by crystal growth conditions.  相似文献   

11.
The aromatic osmacyclopropenefuran bicycles [OsTp{κ3‐C1,C2,O‐(C1H2C2CHC(OEt)O)}(PiPr3)]BF4 (Tp=hydridotris(1‐pyrazolyl)borate) and [OsH{κ3‐C1,C2,O‐(C1H2C2CHC(OEt)O)}(CO)(PiPr3)2]BF4, with the metal fragment in a common vertex between the fused three‐ and five‐membered rings, have been prepared via the π‐allene intermediates [OsTp(η2‐CH2=CCHCO2Et)(OCMe2)(PiPr3)]BF4 and [OsH(η2‐CH2=CCHCO2Et)(CO)(OH2)(PiPr3)2]BF4, and their aromaticity analyzed by DFT calculations. The bicycle containing the [OsH(CO)(PiPr3)2]+ metal fragment is a key intermediate in the [OsH(CO)(OH2)2(PiPr3)2]BF4‐catalyzed regioselective anti‐Markovnikov hydration of ethyl buta‐2,3‐dienoate to ethyl 4‐hydroxycrotonate.  相似文献   

12.
A family of germyl rhodium complexes derived from the PGeP germylene 2,2’-bis(di-isopropylphosphanylmethyl)-5,5’-dimethyldipyrromethane-1,1’-diylgermanium(II), Ge(pyrmPiPr2)2CMe2 ( 1 ), has been prepared. Germylene 1 reacted readily with [RhCl(PPh3)3] and [RhCl(cod)(PPh3)] (cod=1,5-cyclooctadiene) to give, in both cases, the PGeP-pincer chloridogermyl rhodium(I) derivative [Rh{κ3P,Ge,P-GeCl(pyrmPiPr2)2CMe2}(PPh3)] ( 2 ). Similarly, the reaction of 1 with [RhCl(cod)(MeCN)] afforded [Rh{κ3P,Ge,P-GeCl(pyrmPiPr2)2CMe2}(MeCN)] ( 3 ). The methoxidogermyl and methylgermyl rhodium(I) complexes [Rh{κ3P,Ge,P-GeR(pyrmPiPr2)2CMe2}(PPh3)] (R=OMe, 4 ; Me, 5 ) were prepared by treating complex 2 with LiOMe and LiMe, respectively. Complex 5 readily reacted with CO to give the carbonyl rhodium(I) derivative [Rh{κ3P,Ge,P-GeR(pyrmPiPr2)2CMe2}(CO)] ( 6 ), with HCl, HSnPh3 and Ph2S2 rendering the pentacoordinate methylgermyl rhodium(III) complexes [RhHX{κ3P,Ge,P-GeMe(pyrmPiPr2)2CMe2}] (X=Cl, 7 ; SnPh3, 8 ) and [Rh(SPh)23P,Ge,P-GeMe(pyrmPiPr2)2CMe2}] ( 9 ), respectively, and with H2 to give the hexacoordinate derivative [RhH23P,Ge,P-GeMe(pyrmPiPr2)2CMe2}(PPh3)] ( 10 ). Complexes 3 and 5 are catalyst precursors for the hydroboration of styrene, 4-vinyltoluene and 4-vinylfluorobenzene with catecholborane under mild conditions.  相似文献   

13.
To determine the mechanism responsible for the formation of electrolytic sodium–vanadium oxide bronze e-Na x V2O5, synthesized earlier from acid vanadyl sulfate electrolyte, -bronze i-Na x V2O5is synthesized by exposing electrolytic oxide e-V2O5in the same sodium-containing electrolyte under open-circuit conditions, with a subsequent annealing of the sample. It is established that the two modifications of -bronze (e-Na x V2O5and i-Na x V2O5) are identical and that electrolytic precursors of -bronze Na x V2O5form via an ion-exchange mechanism.  相似文献   

14.
A comparative study of the decay kinetics of photogenerated transients from small (60 kDalton) and native (124 kDalton) oat phytochrome in the red-absorbing form (Pr) in phosphate buffer containing 5 mM ethylenediamine tetraacetic acid, pH 7.8, (PB) and in PB containing 20% ethylene glycol, has been carried out in the temperature range 275–298 K. The analysis confirmed that at least two primary photoproducts, intermediates Ii700s and Ii7oo are formed from Pr. The kinetic parameters, as observed in PB at 695 nm and 275 K, are similar for the I700 intermediates of both small and native phytochrome. Namely, the lifetimes are about 21 μs (component percentages 38%) for the I Ii700s and about 200 μ.s (62%) for the Ii700S- Arrhenius preexponential factors (A) of about 1016 and 1015 s-1and activation energies of about 61 and 56 kJ/mol were measured for the absorbance decays of the I700S of small and native phytochrome, respectively. The kinetic data favour parallel paths for the formation of the Ii700s from Pr, and the activation parameters indicate that the primary photoreactions of the transformation from Pr to the far-red-absorbing form are restricted to the chromophore within the protein. Moreover, the relatively modest temperature dependence of the lifetimes of the Ii700S from small and native Pr supports the working hypothesis that the ground state reactions to the Iibl, intermediates–although somewhat influenced by the polypeptide fragment that is removed upon degradation of native to small Pr–are localized to the chromophore, as is most probably the case also for the primary photoreactions. The effect of the addition of 20% ethylene glycol on the pre-exponential factors of the time-dependent decay functions is discussed in similar terms of the early stages of the phototransformation.  相似文献   

15.
The complex trans-[Rh(CO)(NH3)(PiPr3)2]PF6 (2) was prepared from [(η3-C3H5)Rh(PiPr3)2] (1), NH4PF6 and CO or from 1 and NH4PF6 in presence of an excess of methanol. With an excess of CO, the dicarbonyl and tricarbonyl compounds trans-[Rh(CO)2(PiPr3)2]PF6 (3) and [Rh(CO)3(PiPr3)2]PF6 (4) were obtained. Displacement of one CO ligand in 3 by pyridine and acetone led to the formation of trans-[Rh(CO)(py)PiPr3)2]PF6 (5a) and trans-[Rh(CO) (O=CMe2(PiPr3)2]PF6 (6), respectively. Treatment of 1 with [pyH]BF4 and pyridine gave trans-[Rh(py)2(PiPr3)2]BF4 (7); in presence of H2 the dihydrido complex [RhH2(py)2(PiPr3)2]BF4 (8) was formed. The reaction of 1 with NH4PF6 and ethylene produced trans [Rh(C2H4(NH3(PiPr3)2]PF6(9) whereas with methylvinylketone and acetophenone the octahedral hydridorhodium(III) complexes [RhH(η2-CH=CHC(=O)CH3 (NH3(PiPr3)2]PF6(11) and [RhH(η2-C6H4C(=O)CH3(NH3(Pipr3)2]PF6 (13) were obtained. The synthesis of the cationic vinylidenerhodium(I) compounds trans-[Rh(=C=CHR)(py)(PiPr3)2]BF4 (14–16) and trans-[Rh(=C=CHR)(NH3)(PiPr3) 2]PF6 (17–19) was achieved either on treatment of 1 with [pyH]BF4 or NH4PF6 in presence of 1-alkynes or by ethylene displacement from 9 by HCCR. With tert-butylacetylene as substrate, the alkinyl(hydrido)rhodium(III) complex [RhH(CCtBu)(NH3)(O=CMe2)(PiPr3) 2]PF6 (20) was isolated which in CH2Cl2 solution smoothly reacted to give 19 (R =tBu). The cationic but-2-yne compound trans-[Rh(MeCCMe)(NH3)(Pi Pr3)2]PF6 (21) was prepared from 1, NH4PF6 and C2Me2. The molecular structures of 3 and 14 were determined by X-ray crystallography; in both cases the square-planar coordination around the metal and the trans disposition of the phosphine ligands was confirmed.

Abstract

Der Komplex trans-[Rh(CO)(NH3)(PiPr3)2]PF6 (2) wurde aus [(η3-C3H5)Rh(PiPr3)2] (1), NH4PF6 und CO oder aus 1, NH4PF6 und Methanol hergestellt. In Gegenwart von überschüssigem CO wurden die Dicarbonyl- und Tricarbonyl-Verbindungen trans-[Rh(CO)2(PiPr3)2]PF6 (3) und [Rh(CO)3(PiPr3)2]PF6 (4) erhalten. Die Verdrängung eines CO-Liganden in 3 durch Pyridin oder Aceton führte zur Bildung von trans-[Rh(CO)(py)(PiPr3)2]PF6 (5a) bzw. trans-[Rh(CO)(O=CMe2)(PiPr3)2]PF6 (6). Bei Einwirkung von [pyH]BF4 und Pyridin auf 1 entstand trans-[Rh(py)2(PiPr3)2]BF4 (7); in Gegenwart von H2 bildete sich der Dihydrido-Komplex [RhH2(py)2(PiPr3) 2]BF4 (8). Die Reaktion von 1 mit NH4PF6 und Ethen lieferte trans-[Rh(C2H4)(NH3)(PiPr3)2] PF6 (9) während mit Methylvinylketon und Acetophenon die oktaedrischen Hydridorhodium(III)-Komplexe [RhH(η2-CH=CHC(=O)CH3 (NH3)-(PiPr3)2]PF6 (11) und [RhH(η-2-C6H4C(=O)CH3(NH3)(PiPr3)2)2]PF6 (13) erhalten wurden. Die Synthese der kationischen Vinyli-denrhodium(I)-Verbindungen trans-[Rh(=C=CHR(py)(PiPr3)2]BF4 (14–16) und trans-[Rh(=C=CHR)(NH3)(PiPr3)2]PF6 (17–19) gelang durch Einwirkung von [pyH]BF4 bzw. NH4PF6 auf 1 in Gegenwart von 1-Alkinen oder durch Ethen-Verdrängung aus 9 mit HCCR. Mit tert-Butylacetylen als Reaktionspartner wurde der Alkinyl(hydrido)rhodium(III)-Komplex [RhH(CCtBu)(NH3(O=CMe2)(PiPr3)2]PF6 (20) isoliert, der in CH2Cl2-Lösung sofort zu 19 (R =tBu) reagiert. Die kationische 2-Butin-Verbindung trans -[Rh(MeCCMe)(NH3)PiPr3)2]PF6 (21) wurde aus 1, NH4PF6 und C2Me2 hergestellt. Die Strukturen von 3 und 14 wurden kristallographisch bestimmt; in beiden Fa len ließ sich die quadratisch-planare Koordination des Metalls und die trans-Anordnung der Phosphanliganden bestätigen.  相似文献   

16.
The mixing of [V10O28]6− decavanadate anions with a dicationic gemini surfactant ( gem ) leads to the spontaneous self-assembly of surfactant-templated nanostructured arrays of decavanadate clusters. Calcination of the material under air yields highly crystalline, sponge-like V2O5 ( gem -V2O5 ). In contrast, calcination of the amorphous tetrabutylammonium decavanadate allows isolation of a more agglomerated V2O5 consisting of very small crystallites ( TBA -V2O5 ). Electrochemical analysis of the materials’ performance as lithium-ion intercalation electrodes highlights the role of morphology in cathode performance. The large crystallites and long-range microstructure of the gem -V2O5 cathode deliver higher initial capacity and superior capacity retention than TBA -V2O5 . The smaller crystallite size and higher surface area of TBA -V2O5 allow faster lithium insertion and superior rate performance to gem -V2O5 .  相似文献   

17.
The synthesis and reactivity of a CoI pincer complex [Co(?3P,CH,P‐P(CH)PNMeiPr)(CO)2]+ featuring an η2‐ Caryl?H agostic bond is described. This complex was obtained by protonation of the CoI complex [Co(PCPNMeiPr)(CO)2]. The CoIII hydride complex [Co(PCPNMeiPr)(CNtBu)2(H)]+ was obtained upon protonation of [Co(PCPNMeiPr)(CNtBu)2]. Three ways to cleave the agostic C?H bond are presented. First, owing to the acidity of the agostic proton, treatment with pyridine results in facile deprotonation (C?H bond cleavage) and reformation of [Co(PCPNMeiPr)(CO)2]. Second, C?H bond cleavage is achieved upon exposure of [Co(?3P,CH,P‐P(CH)PNMeiPr)(CO)2]+ to oxygen or TEMPO to yield the paramagnetic CoII PCP complex [Co(PCPNMeiPr)(CO)2]+. Finally, replacement of one CO ligand in [Co(?3P,CH,P‐P(CH)PNMeiPr)(CO)2]+ by CNtBu promotes the rapid oxidative addition of the agostic η2‐Caryl?H bond to give two isomeric hydride complexes of the type [Co(PCPNMeiPr)(CNtBu)(CO)(H)]+.  相似文献   

18.
The synthesis and reactivity of a CoI pincer complex [Co(ϰ3P,CH,P‐P(CH)PNMeiPr)(CO)2]+ featuring an η2‐ Caryl−H agostic bond is described. This complex was obtained by protonation of the CoI complex [Co(PCPNMeiPr)(CO)2]. The CoIII hydride complex [Co(PCPNMeiPr)(CNtBu)2(H)]+ was obtained upon protonation of [Co(PCPNMeiPr)(CNtBu)2]. Three ways to cleave the agostic C−H bond are presented. First, owing to the acidity of the agostic proton, treatment with pyridine results in facile deprotonation (C−H bond cleavage) and reformation of [Co(PCPNMeiPr)(CO)2]. Second, C−H bond cleavage is achieved upon exposure of [Co(ϰ3P,CH,P‐P(CH)PNMeiPr)(CO)2]+ to oxygen or TEMPO to yield the paramagnetic CoII PCP complex [Co(PCPNMeiPr)(CO)2]+. Finally, replacement of one CO ligand in [Co(ϰ3P,CH,P‐P(CH)PNMeiPr)(CO)2]+ by CNtBu promotes the rapid oxidative addition of the agostic η2‐Caryl−H bond to give two isomeric hydride complexes of the type [Co(PCPNMeiPr)(CNtBu)(CO)(H)]+.  相似文献   

19.

Background

Although the necessity of divalent magnesium and manganese for full activity of sugar nucleotidyltransferases and glycosyltransferases is well known, the role of these metal cations in binding the substrates (uridine 5'-triphosphate, glucose-1-phosphate, N-acetylglucosamine-1-phosphate, and uridine 5'-diphosphate glucose), products (uridine 5'-diphosphate glucose, uridine 5'-diphosphate N-acetylglucosamine, pyrophosphate, and uridine 5'-diphosphate), and/or enzyme is not clearly understood.

Results

Using isothermal titration calorimetry we have studied the binding relationship between the divalent metals, magnesium and manganese, and uridine 5'-phosphates to determine the role these metals play in carbohydrate biosynthesis. It was determined from the isothermal titration calorimetry (ITC) data that Mg+2 and Mn+2 are most tightly bound to PP i , Kb = 41,000 ± 2000 M-1 and 28,000 ± 50,000 M-1 respectively, and UTP, Kb = 14,300 ± 700 M-1 and 13,000 ± 2,000 M-1 respectively.

Conclusion

Our results indicate that the formal charge state of the phosphate containing substrates determine the binding strength. Divalent metal cations magnesium and manganese showed similar trends in binding to the sugar substrates. Enthalpy of binding values were all determined to be endothermic except for the PP i case. In addition, entropy of binding values were all found to be positive. From this data, we discuss the role of magnesium and manganese in both sugar nucleotidyltransferase and glycosyltransferase reactions, the differences in metal-bound substrates expected under normal physiological metal concentrations and those of hypomagnesaemia, and the implications for drug design.  相似文献   

20.
The substitution equilibria AuCl 2 ? + iNH 4 + = Au(NH3)iCl2 ? i + iCl? + iH+, β i * . were studied pH-metrically at 25°C and I = 1 mol/L (NaCl) in aqueous solution. It was found that logβ 1 * = ?5.10±0.15 and logβ 2 * = ?10.25±0.10. For equilibrium AuNH3Clsolid = AuNH3Cl, log K s = ?3.1±0.3. Taking into account the protonation constants of ammonia (log K H = 9.40), the obtained results show that for equilibria AuCl 2 ? + iNH3 = Au(NH3)iCl2 ? i + iCl?, logβ1 = 4.3±0.2, and logβ2 = 8.55±0.15. The standard potentials E 0 1/0 of AuNH3Cl0 and Au(NH3) 2 + species are equal to 0.90±0.02 and 0.64±0.01 V, respectively.  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