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.
The hydrolysis of p-nitrophenyl sulfate, p-nitrocatechol sulfate, and [35S]sodium dodecyl sulfate was examined in anoxic sediments of Wintergreen Lake, Michigan. Significant levels of sulfhydrolase activity were observed in littoral, transition, and profundal sediment samples. Rates of sulfate formation suggest that the sulfhydrolase system would represent a major source of sulfate within these sediments. Sulfate formed by ester sulfate hydrolysis can support dissimilatory sulfate reduction as shown by the incorporation of 35S from labeled sodium dodecyl sulfate into H235S. Sulfhydrolase activity varied with sediment depth, was greatest in the littoral zone, and was sensitive to the presence of oxygen. Estimations of ester sulfate concentrations in sediments revealed large quantities of ester sulfate (~30% of total sulfur). Both total sulfur and ester sulfate concentrations varied with the sediment type and were two to three orders of magnitude greater than the inorganic sulfur concentration.  相似文献   

2.
Using 4-methylumbelliferyl-β-D-N,N′,N″-triacetylchitotrioside (4-MU-GlcNAc3) as a substrate, an acidic chitinase was purified from seeds of black soybean (Glycine max Tainan no. 3) by ammonium sulfate fractionation and three successive steps of column chromatography. The purified chitinase was a monomeric enzyme with molecular mass of 20.1 kDa and isoelectric point of 4.34. The enzyme catalyzed the hydrolysis of synthetic substrates p-nitrophenyl N-acetyl chitooligosaccharides with chain length from 3 to 5 (GlcNAcn, n = 3-5), and pNp-GlcNAc4 was the most degradable substrate. Using pNp-GlcNAc4 as a substrate, the optimal pH for the enzyme reaction was 4.0; kinetic parameters K m and kcat were 245 µM and 10.31 min−1, respectively. This enzyme also showed activity toward CM-chitin-RBV, a polymer form of chitin, and N-acetyl chitooligosaccharides, an oligomer form of chitin. The smallest oligomer substrate was an N-acetylglucosamine tetramer. These results suggested that this enzyme was an endo-splitting chitinase with short substrate cleavage activity and useful for biotechnological applications, in particular for the production of N-acetyl chitooligosaccharides.  相似文献   

3.
The human sulfatase family has 17 members, 13 of which have been characterized biochemically. These enzymes specifically hydrolyze sulfate esters in glycosaminoglycans, sulfolipids, or steroid sulfates, thereby playing key roles in cellular degradation, cell signaling, and hormone regulation. The loss of sulfatase activity has been linked to severe pathophysiological conditions such as lysosomal storage disorders, developmental abnormalities, or cancer. A novel member of this family, arylsulfatase K (ARSK), was identified bioinformatically through its conserved sulfatase signature sequence directing posttranslational generation of the catalytic formylglycine residue in sulfatases. However, overall sequence identity of ARSK with other human sulfatases is low (18–22%). Here we demonstrate that ARSK indeed shows desulfation activity toward arylsulfate pseudosubstrates. When expressed in human cells, ARSK was detected as a 68-kDa glycoprotein carrying at least four N-glycans of both the complex and high-mannose type. Purified ARSK turned over p-nitrocatechol and p-nitrophenyl sulfate. This activity was dependent on cysteine 80, which was verified to undergo conversion to formylglycine. Kinetic parameters were similar to those of several lysosomal sulfatases involved in degradation of sulfated glycosaminoglycans. An acidic pH optimum (∼4.6) and colocalization with LAMP1 verified lysosomal functioning of ARSK. Further, it carries mannose 6-phosphate, indicating lysosomal sorting via mannose 6-phosphate receptors. ARSK mRNA expression was found in all tissues tested, suggesting a ubiquitous physiological substrate and a so far non-classified lysosomal storage disorder in the case of ARSK deficiency, as shown before for all other lysosomal sulfatases.  相似文献   

4.
Arylsulfatase from Artemia salina exists in at least two forms (AS I and AS II). The paper presents characterization of the AS II form of the arylsulfatase. The enzyme was able to hydrolyze p-nitrocatechol sulfate (pNCS) as well as ascorbate sulfate. It exhibited maximum activity at temperature of 50 °C and was stable for 2 h at 4-10 °C. Optimum pH shifted from 6.2 at 4 mM pNCS (substrate) to 4.8 at 20 mM pNCS. The enzyme displayed linear kinetics. AS II arylsulfatase exists in two molecular forms (349 and 460 kDa) composed of identical subunits with molecular mass of 53 kDa. Sulfite and phosphate ions were the most potent inhibitors of the enzyme. Cyanide proved to be a weak inhibitor. Sulfate and low concentrations of silver ions had no effect on the enzyme activity. Based on the above results, modifications in the assay for determination of enzyme activity are proposed.  相似文献   

5.
The phototrophic bacterium Rhodobacter capsulatus detoxified p-nitrophenol and 4-nitrocatechol. The bacterium tolerated moderate concentrations of p-nitrophenol (up to 0.5 mM) and degraded it under light at an optimal O2 pressure of 20 kPa. The bacterium did not metabolize the xenobiotic in the dark or under strictly anoxic conditions or high O2 pressure. Bacterial growth with acetate in the presence of p-nitrophenol took place with the simultaneous release of nonstoichiometric amounts of 4-nitrocatechol, which can also be degraded by the bacterium. Crude extracts from R. capsulatus produced 4-nitrocatechol from p-nitrophenol upon the addition of NAD(P)H, although at a very low rate. A constitutive catechol 1,2-dioxygenase activity yielding cis,cis-muconate was also detected in crude extracts of R. capsulatus. Further degradation of 4-nitrocatechol included both nitrite- and CO2-releasing steps since: (1) a strain of R. capsulatus (B10) unable to assimilate nitrate and nitrite released nitrite into the medium when grown with p-nitrophenol or 4-nitrocatechol, and the nitrite concentration was stoichiometric with the 4-nitrocatechol degraded, and (2) cultures of R. capsulatus growing microaerobically produced low amounts of 14CO2 from radiolabeled p-nitrophenol. The radioactivity was also incorporated into cellular compounds from cells grown with uniformly labeled 14C-p-nitrophenol. From these results we concluded that the xenobiotic is used as a carbon source by R. capsulatus, but that only the strain able to assimilate nitrite (E1F1) can use p-nitrophenol as a nitrogen source. Received: 30 December 1996 / Accepted: 3 September 1997  相似文献   

6.
Arylsulfhydrolases A and B from chicken and from bovine liver have been isolated and their reactions with a range of synthetic arylsulfates examined using kinetic methods. Some differences of Michaelis-Menten parameters were observed in comparing the A with the B forms from the two sources at the level of individual substrates. At that level also, interspecies comparisons of A forms and B forms similarly showed differences. However, for none of the four enzymes examined was there consistent correlations of kinetic values with electronic, hydrophobicity, or steric properties of the substrates. The bovine A enzyme displayed the well-documented “anomalous” kinetic behavior at high substrate concentrations; at low concentrations conventional hydrolysis of p-nitrocatechol sulfate occurred, except that there was evidence with this substrate and others of product inhibition. The avian A enzyme reacted normally over all substrate concentrations examined, but again product inhibition occurred. The mammalian but not the avian B enzyme was also clearly subject to product inhibition.  相似文献   

7.
Bacteria that metabolize p-nitrophenol (PNP) oxidize the substrate to 3-ketoadipic acid via either hydroquinone or 1,2,4-trihydroxybenzene (THB); however, initial steps in the pathway for PNP biodegradation via THB are unclear. The product of initial hydroxylation of PNP could be either 4-nitrocatechol or 4-nitroresorcinol. Here we describe the complete pathway for aerobic PNP degradation by Bacillus sphaericus JS905 that was isolated by selective enrichment from an agricultural soil in India. Washed cells of PNP-grown JS905 released nitrite in stoichiometric amounts from PNP and 4-nitrocatechol. Experiments with extracts obtained from PNP-grown cells revealed that the initial reaction is a hydroxylation of PNP to yield 4-nitrocatechol. 4-Nitrocatechol is subsequently oxidized to THB with the concomitant removal of the nitro group as nitrite. The enzyme that catalyzed the two sequential monooxygenations of PNP was partially purified and separated into two components by anion-exchange chromatography and size exclusion chromatography. Both components were required for NADH-dependent oxidative release of nitrite from PNP or 4-nitrocatechol. One of the components was identified as a reductase based on its ability to catalyze the NAD(P)H-dependent reduction of 2,6-dichlorophenolindophenol and nitroblue tetrazolium. Nitrite release from either PNP or 4-nitrocatechol was inhibited by the flavoprotein inhibitor methimazole. Our results indicate that the two monooxygenations of PNP to THB are catalyzed by a single two-component enzyme system comprising a flavoprotein reductase and an oxygenase.  相似文献   

8.
An oxidative pathway for the mineralization of 2,4-dinitrotoluene (2,4-DNT) by Burkholderia sp. strain DNT has been reported previously. We report here the isolation of additional strains with the ability to mineralize 2,4-DNT by the same pathway and the isolation and characterization of bacterial strains that mineralize 2,6-dinitrotoluene (2,6-DNT) by a different pathway. Burkholderia cepacia strain JS850 and Hydrogenophaga palleronii strain JS863 grew on 2,6-DNT as the sole source of carbon and nitrogen. The initial steps in the pathway for degradation of 2,6-DNT were determined by simultaneous induction, enzyme assays, and identification of metabolites through mass spectroscopy and nuclear magnetic resonance. 2,6-DNT was converted to 3-methyl-4-nitrocatechol by a dioxygenation reaction accompanied by the release of nitrite. 3-Methyl-4-nitrocatechol was the substrate for extradiol ring cleavage yielding 2-hydroxy-5-nitro-6-oxohepta-2,4-dienoic acid, which was converted to 2-hydroxy-5-nitropenta-2,4-dienoic acid. 2,4-DNT-degrading strains also converted 2,6-DNT to 3-methyl-4-nitrocatechol but did not metabolize the 3-methyl-4-nitrocatechol. Although 2,6-DNT prevented the degradation of 2,4-DNT by 2,4-DNT-degrading strains, the effect was not the result of inhibition of 2,4-DNT dioxygenase by 2,6-DNT or of 4-methyl-5-nitrocatechol monooxygenase by 3-methyl-4-nitrocatechol.  相似文献   

9.
A continuous-stirred, hydrogen-based, hollow-fiber membrane biofilm reactor (HFMBfR) that was active in nitrate and sulfate reductions was shown to be effective for degradation or detoxification of para-chloronitrobenzene (p-CNB) in water by biotransforming it first to para-chloroaniline (nitro-reduction) and then to aniline (reductive dechlorination) with hydrogen (H2) as an electron donor. A series of short-term experiments examined the effects of nitrate and sulfate on p-CNB bioreduction. The results obtained showed both higher nitrate and sulfate concentration declined the p-CNB bioreduction in the biofilm, and this suggests the competition for H2 caused less H2 available for the p-CNB bioreduction when the H2 demand for the reductions was larger. Denitrification and sulfate reduction intermediates were thought to be potential factors inhibiting the p-CNB bioreduction. Analysis of electron-equivalent fluxes and reaction orders in the biofilm further demonstrated both denitrification and sulfate reduction competed more strongly for H2 availability than p-CNB bioreduction. These findings have significant implications for the HFMBfR used for degrading p-CNB under denitrifying and/or sulfate reducing conditions.  相似文献   

10.
A previously uncharacterized gene in Neosartorya fischeri was cloned and expressed in Escherichia coli. It was found to encode a β-glucosidase (NfBGL1) distinguishable from other BGLs by its high turnover of p-nitrophenyl β-d-glucopyranoside (pNPG). Molecular determinants for the substrate recognition of NfBGL1 were studied through an initial screening of residues by sequence alignment, a second screening by homology modeling and subsequent site-directed mutagenesis to alter individual screened residues. A conserved amino acid, E445, in the substrate binding pocket of wild-type NfBGL1 was identified as an important residue affecting substrate affinity. Replacement of E445 with amino acids other than aspartate significantly decreased the catalytic efficiency (kcat/Km) of NfBGL1 towards pNPG, mainly through decreased binding affinity. This was likely due to the disruption of hydrogen bonding between the substrate and the carboxylate oxygen of the residue at position 445. Density functional theory (DFT) based studies suggested that an acidic amino acid at position 445 raises the substrate affinity of NfBGL1 through hydrogen bonding. The residue E445 is completely conserved indicating that this position can be considered as a crucial determinant for the substrate binding among GHs tested.  相似文献   

11.
A detailed study of the pH dependence of the Michaelis-Menten constants (V and Km) of aryl sulfatase A (EC 3.1.6.1) from rabbit liver indicates that at least two functional groups (pK's ~4.3 and ~7 in the enzyme-substrate complex) participate in the enzymic degradation of substrate. Aryl sulfatase A is inactivated by diethyl pyrocarbonate (ethoxyformic anhydride). The enzyme that has been modified with this reagent can in turn be reactivated by treatment with hydroxylamine. The pH dependence of inactivation reveals a reactive group having a pK of 6.5–7.0. The results indicate that at least one histidine plays an important catalytic role in rabbit liver aryl sulfatase A, consistent with the results of earlier workers who employed diazotized sulfanilic acid. Phosphate ion, a competitive inhibitor, partially protects the enzyme from inactivation by diethyl pyrocarbonate whereas sulfate ion, also a competitive inhibitor, increases the rate of inactivation by diethyl pyrocarbonate. This result is of particular significance in view of the anomalous kinetics of aryl sulfatase A. The kinetic effects of even small amounts of sulfate ion impurities in many commercial sulfate ester substrate preparations is also discussed.  相似文献   

12.
Purified acidic (pI 4.9), neutral (pI 6.9), and basic (pI 8.7) phospholipase A2 from Agkistrodon halys blomhofii showed characteristically different patterns of hemolysis and phospholipid hydrolysis of intact human erthyrocytes. Acidic and neutral enzymes were nonlytic in the early periods of incubations with intact erythrocytes whereas the basic enzyme caused immediate hemolysis (5–8%). Under nonlytic conditions acidic and neutral enzymes hydrolyzed only phosphatidyl choline (PC) (20 and 50%, respectively), whereas basic enzyme hydrolyzed not only PC (60%) but nearly 15% of the phosphatidylethanolamine (PE). Both PC and PE were hydrolyzed significantly when the three phospholipases A2 were incubated individually with erythrocyte lysate or hypotonic ghosts (sealed or unsealed). The order of substrate preference for acidic and neutral enzymes was always PC > PE. On the contrary basic enzyme exhibited the property of substrate specificity reversal. It hydrolyzed PC faster than PE when the membranes were sealed whereas PE hydrolysis was faster than PC hydrolysis in unsealed membranes. Interestingly only the basic enzyme showed activity in the absence of Ca2+ and in the presence of 0.5 mm EDTA. Phospholipase C (Bacillus cereus or Clostridium perfringens) did not show the property of substrate specificity reversal although their ability to hydrolyze PC and PE was different. In general this study demonstrates the unique activity patterns of three physically different pure phospholipases A2 on human erythrocyte membranes which could be of value in selectively modifying membrane phospholipids. In addition it also throws an important light on the fact that results obtained with phospholipases should be interpreted with caution particularly as regards the localization of phospholipids in membranes.  相似文献   

13.
A low-molecular-weight human liver acid phosphatase was purified 2580-fold to homogenity by a procedure involving ammonium sulfate fractionation, acid treatment, and SP-Sephadex ion-exchange chromatography with ion-affinity elution. The purified enzyme contains a single polypeptide chain and has a molecular weight of 14,400 as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The amino acid composition of this enzyme (E) is reported. A pH dependence study using p-nitrophenyl phosphate as a substrate (S) revealed the effect of substrate ionization (pKa 5.2) and the participation of a group in the ES complex having a pKa value of 7.8. The enzyme is readily inactivated by sulfhydryl reagents such as heavy metal ions. Alkylation of the enzyme with iodoacetic acid and iodoacetamide causes complete inactivation of the enzyme and this inactivation is prevented by the presence of phosphate ion. The enzyme is also inactivated by treatment with diethyl pyrocarbonate; protection against this reagent is afforded by phosphate ion. The substrate specificity of this enzyme is unusual for an acid phosphatase. Of the many alkyl and aryl phosphomonoesters tested, the only possibly physiological substrate hydrolyzed by this enzyme was flavin mononucleotide, which exhibits a V which is 3-fold larger at pH 5.0 and 6-fold larger at pH 7.0 than that for p-nitrophenyl phosphate. However, the enzyme also catalyzes the hydrolysis of acetyl phosphate at pH 5.0 with a velocity eight times larger than that reported for an acyl phosphatase from human erythrocytes.  相似文献   

14.
N.N. Rao  V.V. Modi 《Phytochemistry》1976,15(10):1437-1439
Fructose-1,6-diphosphatase (FDPase) from unripe mango was separated into two components by ammonium sulfate fractionation, one active at pH 6 (acidic FDPase) and the other at pH 8.5 (alkaline FDPase). The alkaline component had a lower Km. (0.15 × 10?3 M) than the acidic component (1.7 × 10?3 M) towards the substrate (FDP) and the allosteric inhibitor AMP. It also showed greater heat stability and higher activation in the presence of EDTA as compared to the acidic FDPase. Both components showed a higher activation with Mn2+ ions than with Mg2+ ions.  相似文献   

15.
Formation of Thiosulfate from Sulfite by Desulfovibrio vulgaris   总被引:12,自引:9,他引:3       下载免费PDF全文
Crude extracts of Desulfovibrio vulgaris reduced sulfite to sulfide. Ammonium sulfate fractionation of crude extracts separated a thiosulfate-forming system from sulfite- and thiosulfate-reductase activities. Further purification by sucrose density centrifugation separated the thiosulfate-forming system into two components, both of which were required for the reaction. In addition to these two components, cytochrome c3, ferredoxin, and hydrogenase were required to form thiosulfate from sulfite. By absorption spectra and from the effect of pH and substrate concentration, the ionic species acting as the substrate for thiosulfate-formation was concluded to be bisulfite.  相似文献   

16.
Sedimentation equilibrium studies on arylsulfate sulfohydrolase II (EC 3.1.6.1) from Aspergillus oryzae under nondissociating conditions have resulted in a revised molecular weight of 94,900 ± 7100. Sedimentation equilibrium and gel electrophoresis data collected in the presence of the dissociating agents, urea and sodium dodecylsulfate demonstrate that the native enzyme is composed of two identical subunits as suggested by previous studies employing an irreversible inhibitor.The pH dependencies of the kinetic parameters V and VKm for the enzymic hydrolysis of 4-nitrophenyl sulfate indicate that two groups of pKa 4.7 and 6.0 control the activity of the enzyme. The product inorganic sulfate was shown to be a linear competitive inhibitor of the enzyme at pH 4.0, implying that it is a last released product along the reaction pathway. Inhibition by the phenol product was not observed. Enzymic hydrolysis of 4-nitrophenyl sulfate in 18O enriched water revealed that one atom of solvent oxygen is incorporated per molecule of inorganic sulfate, which is consistent with a mechanism featuring sulfur-oxygen bond cleavage. Evidence is presented based on stopped-flow kinetics, partitioning experiments in the presence of amine nucleophiles, and 18O exchange studies that collectively suggest that the breakdown of a covalent sulfuryl enzyme intermediate probably is not the rate-limiting step along the reaction pathway.The substrate specificity of the enzyme was examined by testing a variety of sulfate and phosphate esters as inhibitors of the hydrolysis of 4-nitrophenyl sulfate. The Cbz-l-Phe-l-Tyrosine-O-sulfate methyl ester serves as a substrate for the enzyme. Apparently substrate activity requires an aromatic sulfate ester whose binding is enhanced by incorporating the aromatic moiety in a hydrophobic matrix.  相似文献   

17.
The p-nitrophenyl phosphatase activity of muscle carbonic anhydrase   总被引:6,自引:0,他引:6  
Carbonic anhydrase III from rabbit muscle, a newly discovered major isoenzyme of carbonic anhydrase, has been found to be also a p-nitrophenyl phosphatase, an activity which is not associated with carbonic anhydrases I and II. The p-nitrophenyl phosphatase activity has been shown to chromatograph with the CO2 hydratase activity; both activities are associated with each of its sulfhydryl oxidation subforms; and both activities follow the same pattern of pH stability. This phosphomonoesterase activity of carbonic anhydrase III has an acidic pH optimum (<5.3); its true substrate appears to be the phosphomonoanion with a Km of 2.8 mm. It is competitively inhibited by the typical acid phosphatase inhibitors phosphate (Ki = 1.22 × 10?3M), arsenate (Ki = 1.17 × 10?3M), and molybdate (Ki = 1.34 × 10?7M), with these inhibitors having no effect on the CO2 hydratase or the p-nitrophenyl acetate esterase activities of carbonic anhydrase III. The p-nitrophenyl acetate esterase activity of carbonic anhydrase III, on the other hand, has the sigmoidal pH profile with an inflection at neutral pH, typical of carbonic anhydrases for all of their substrates, and is inhibitable by acetazolamide (a highly specific carbonic anhydrase inhibitor) to the same degree as the CO2 hydratase activity. The acid phosphatase-like activity of carbonic anhydrase III is slightly inhibited by acetazolamide at acidic pH, and inhibited to nearly the same degree at neutral pH. These data are taken to suggest that the phosphatase activity follows a mechanism different from that of the CO2 hydratase and p-nitrophenyl acetate esterase activities and that there is some overlap of the binding sites.  相似文献   

18.
The kinetic behavior andpH-stability of recombinant human renin was analyzed using a new fluorogenic substrate based on the normal P6-P3′ renin cleavage sequence in human angiotensinogen. The design of this fluorogenic substrate makes possible, for the first time, direct monitoring of the kinetics of proteolytic conversion of prorenin to renin. ThepH-stability profile for renin, measured with the substrate at 25°C, indicated a broad plateau of stability betweenpH 6.0 and 10.0. Analysis of thepH-activity profile of renin for the substrate indicated a minimumK m (~1.8 µM) atpH ~7.4 and a maximumV m betweenpH 7.4 and 8.0. The thermodynamics of the binding of a novel, soluble, peptidomimetic inhibitor to renin indicated it is possible to retain the tight-binding characteristics and enthalpy contributions to binding of larger peptide-derived inhibitors, while reducing inhibitor size and entropic contributions to binding. A novel derivative of the fluorogenic substrate, containing a 3-methyl histidine substitution at the P2 site, was used to test the recent hypothesis that renin functions by virtue of substrate-directed catalysis.  相似文献   

19.
Two neutral β-galactosidase isozymes were purified from human liver. The initial step of purification was removal of the acidic β-galactosidases by adsorption on concanavalin A-Sepharose 4B conjugate. Subsequent purification steps included ammonium sulfate precipitation, diethylaminoethyl cellulose column chromatography, Sephadex G-100 gel filtration, and preparative polyacrylamide-gel isoelectric focusing. The final step of purification was affinity chromatography of the separated isoelectric forms on ?-aminocaproyl-β-d-galactosylamine-Sepharose 4B conjugate. The purified β-galactosidase isozymes had activity toward both β-d-galactoside and β-d-glucoside derivatives of 4-methylumbelliferone and p-nitrophenol with a pH optimum around 6.2. These enzyme forms were also found to possess lactosylceramidase II activity with a pH optimum in the range of 5.4 to 5.6, but not lactosylceramidase I activity and no activity toward galactosylceramide or GM1-ganglioside. The molecular weight was found to be in the range of 37,500–39,500 for the two neutral isozymes and they had similar Km and V values; the more acidic form (designated β-galactosidase N1) was more heat stable than the other form (designated β-galactosidase N2). Antibodies evoked against the N1 and N2 β-galactosidases gave identical precipitin lines retaining enzymatic activity. No cross-reactivity was observed between the neutral and the acidic isozymes when examined with the respective antisera.  相似文献   

20.
Antibody binding of cartilage-specific proteoglycans   总被引:4,自引:0,他引:4  
The spectroscopically observable acid dissociation constant of aspartate aminotransferase (EC 2.6.1.1) varies to different degrees upon the addition of different monovalent anions. These interactions may be described by the minimal scheme
where XEH and XE represent anion complexes with the acidic (EH) and basic (E) forms of the enzyme, respectively. Both graphical and computer procedures were used to determine the three equilibrium constants which describe such a system. The analysis was based upon the effects of salt concentration (X) upon the apparent pKa of the enzyme determined spectrophotometrically. The affinities for anions of the basic enyzme are less than those of the acidic form of the enzyme so that the apparent pKa rises with anion concentration to a limiting value; pK4 of the enzyme anion complex. That different anion-enzyme complexes have different pK4's reflects the fact that the interaction specificities of the basic and acidic enzymes differ. Cacodylate did not appear either to cause significant effects on the chromophoric pK's or to compete with the binding of halide or carboxylate anions which cause a perturbation of the pKa.  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