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 546 毫秒
1.
We have previously demonstrated that a solid-state buffer could be successfully used to control the ionization state of subtilisin Carlsberg cross-linked microcrystals (CLECs) suspended in supercritical ethane (sc-ethane) in the presence of acid–base active species such as salt hydrates and zeolite molecular sieves. Here we studied the effect of six zwitterionic proton/sodium (pH–pNa) solid-state acid–base buffers on the catalytic activity of subtilisin CLECs in sc-ethane at high and low water activity (aW). CLECs were strongly activated by increasing aW. At high aW, and despite the high hydrolysis rates, transesterification activities were still about one order of magnitude higher than those observed at lower aW. This is in contradiction with what was previously reported in the absence of acid–base control and supports the hypothesis that the poor catalytic performance of subtilisin CLECs at high aW observed in those studies was due to the inhibitory effect of the hydrolytic by-product, rather than to the competition of water with propanol for the acyl-enzyme intermediate. Although the catalytic activity of subtilisin showed a general positive correlation with the aqueous pKa of the acid–base buffers tested here, our results also show that as expected, the acid–base behavior of the buffers in nonaqueous media is more complex than what can be predicted from aqueous-based parameters alone. This work further confirms the usefulness of solid-state acid–base buffers in supercritical biocatalysis but highlights the need for further research on the topic.  相似文献   

2.
The mechanism of the vanadate (V(v))-dependent oxidation of NADH was different in phosphate buffers and in phosphate-free media. In phosphate-free media (aqueous medium or HEPES buffer) the vanadyl (V(v)) generated by the direct V(v)-dependent oxidation of NADH formed a complex with V(v). In phosphate buffers V(v) autoxidized instead of forming a complex with V(v). The generated superoxide radical (O2) initiated, in turn, a high-rate free radical chain oxidation of NADH. Phosphate did not stimulate the V(v)-dependent NADH oxidation catalyzed by O2-generating systems. Monovanadate proved to be a stronger catalyzer of NADH oxidation as compared to polyvanadate.  相似文献   

3.
The SH2 domain containing inositol 5-phosphatase 2 (SHIP2) catalyzes the dephosphorylation of phosphatidylinositol 3,4,5-trisphosphate (PtdIns(3,4,5)P3) to phosphatidylinositol 3,4-bisphosphate (PtdIns(3,4)P2) and participates in the insulin signalling pathway in vivo. In a comparative study of SHIP2 and the phosphatase and tensin homologue deleted on chromosome 10 (PTEN), we found that their lipid phosphatase activity was influenced by the presence of vesicles of phosphatidylserine (PtdSer). SHIP2 PtdIns(3,4,5)P3 5-phosphatase activity was greatly stimulated in the presence of vesicles of PtdSer. This effect appears to be specific for di-C8 and di-C16 fatty acids of PtdIns(3,4,5)P3 as substrate. It was not observed with inositol 1,3,4,5-tetrakisphosphate (Ins(1,3,4,5)P4) another in vitro substrate of SHIP2, nor with Type I Ins(1,4,5)P3/Ins(1,3,4,5)P4 5-phosphatase activity, an enzyme which acts on soluble inositol phosphates. Vesicles of phosphatidylcholine (PtdCho) stimulated only twofold PtdIns(3,4,5)P3 5-phosphatase activity of SHIP2. Both a minimal catalytic construct and the full length SHIP2 were sensitive to the stimulation by PtdSer. In contrast, PtdIns(3,4,5)P3 5-phosphatase activity of the Skeletal muscle and Kidney enriched Inositol Phosphatase (SKIP), another member of the mammaliam Type II phosphoinositide 5-phosphatases, was not sensitive to PtdSer. Our enzymatic data establish a specificity in the control of SHIP2 lipid phosphatase activity with PtdIns(3,4,5)P3 as substrate which is depending on the fatty acid composition of the substrate.  相似文献   

4.
The effect of MRS broth on the stability of hydrogen peroxide (H2O2) has been studied. Known concentrations (1–100 μg ml−1) of H2O2 were prepared in distilled water, phosphate buffer (pH 7·0) and MRS broth (pH 6·2 and 3·9). H2O2 was very stable in aqueous and buffer solutions but it was rapidly degraded in MRS broth (pH 3·9). The presence of H2O2 in MRS broth (pH 6·2) could not be detected.  相似文献   

5.
Acidification of the starchy endosperm by the aleurone layer following germination has been established; however, the physiological and metabolic responses of this tissue to external pH have been incompletely investigated. In this investigation, isolated wheat (Triticum aestivum) aleurone layers were incubated in different solutions at initial pH values of 3, 4 and 6 in the absence of phytohormones. After 24 h of incubation, the initial pH of all malate and succinate buffers shifted towards a value close to 4.2. These results suggest the existence of a pH-stating mechanism, instead of the simple acidification process reported previously. The rise of initial pH 3 by aleurone layers was accompanied by a high net uptake of external malic- or succinic acid. In contrast, incubation in glycyl-glycine buffer (a supposedly non-permeating cation at pH 3) partially prevented that pH rise in a pH-3 solution. The 14C-malate taken up from media at pH 3 was mostly broken down to CO2, indicating that an effective metabolic control of the intracellular malate level was operating. At pH 6, an uptake of 14C-malate and 14CO2 production occurred as well, but at slower rates than at pH 3. When buffer concentration was increased, at initial pH values of 3 or 6, a higher uptake or secretion of malic acid, respectively, was carried out by the aleurone layers. The pH of these buffers varied less than that of dilute ones, but always showed a tendency toward a pH near 4. These results suggest that a balance between secretion and uptake of malic acid, accompanied by the corresponding biosynthesis or degradation, is the basis of this pH-stating mechanism.  相似文献   

6.
Lipid peroxidation (LPO) of polyunsaturated fatty acids (PUFAs) is suspected to be involved in the generation of chronic diseases. A model reaction for LPO is the air oxidation of PUFAs initiated by Fe2+ and ascorbic acid. In the course of such model reactions glycolaldehyde (GLA) was detected as main aldehydic product. Since it is difficult to explain the generation of GLA by oxidation of PUFAs, it was suspected that GLA might be derived by oxidation of ascorbic acid. This assumption was verified by treatment of ascorbic acid with Fe2+.

Produced aldehydic compounds were trapped by addition of pentafluorobenzylhydroxylamine hydrochloride (PFBHA-HCl), trimethylsilylated and finally identified by gas chromatography/mass spectrometry (GC/MS). Oxidation of ascorbic acid with O2 in presence of iron ions produced not only glycolaldehyde (GLA), but also glyceraldehyde (GA), dihydroxyacetone (DA) and formaldehyde. Glyoxal (GO) and malondialdehyde (MDA) were detected as trace compounds.

The yield of the aldehydic compounds was increased by addition of lipid hydroperoxides (LOOH) or H2O2. The buffer influenced the reaction considerably: Iron ions react with Tris buffer by producing dihydroxyace-tone (DA). Since ascorbic acid is present in biological systems and Fe2+ ions are obviously generated by cell damaging processes, the production of GLA and other aldehydic components might add to the damaging effects of LPO.

Glucose suffers also oxidation to short-chain aldehydic compounds in aqueous solution, but this reaction requires addition of equimolar amounts of Fe2+ together with equimolar amounts of H2O2 or 13-hydroperoxy-9-cis-11-trans-octadecadienoic acid (13-HPODE). Therefore this reaction, also influenced by the buffer system, seems to be not of biological relevance.  相似文献   

7.
A cloned bacterial enzyme for nerve agent decontamination   总被引:2,自引:0,他引:2  
Organophosphorus acid (OPA) anhydrolases offer considerable potential for safe, non-corrosive decontamination of chemical nerve agents. The Alteromonas sp. strain JD6.5 gene encoding an OPA anhydrolase (designated as OPAA-2), which hydrolyzes a wide variety of nerve agents, has been cloned in Escherichia coli. Employing agent-analog diisopropyl fluorophosphate (DFP) as a substrate, the effects of buffers, pH, temperature, and various protein stabilizing agents on OPAA-2 activity were studied. Ammonium carbonate, which is innocuous and inexpensive, proved to be a superior buffer for enzyme activity. Compared with enzyme assayed under standard conditions, enzyme activity with ammonium carbonate was six-fold greater. To evaluate effects of storage and reconstitution on enzyme activity, the cloned enzyme was lyophilized, rehydrated, and then assessed by measuring activity against DFP. Whereas almost 100% of the hydrolytic activity was recovered with enzyme reconstituted in (NH4)2CO3-buffered distilled water or chlorinated drinking water, approximately 20% of the activity was recovered with ocean water. Enzyme stability in blast-containment foam or fire-fighting foam was also demonstrated by high activity in (NH4)2CO3-buffered distilled water or drinking water. These findings suggest the potential of a foam-based enzyme system for field decontamination of chemical nerve agents.  相似文献   

8.
The stability of papain was studied in aqueous-organic mixtures by means of residual proteolytic activity along with various spectroscopic analyses (fluorescence and ATR-FTIR combined with isotopic exchange with D2O). The investigated systems contained 1 or 10% (v/v) of an aqueous buffered solution (pH 8.0) in acetonitrile (ACN), methanol (MeOH) or dimethyl formamide (DMF). The results evidenced that papain retained almost all its catalytic activity after 24 h of incubation in the presence of ACN, and a more compact conformation of the enzyme was detected. Papain suffered an important loss of enzymatic activity (ca. 80%) after 24 h incubation in MeOH although, no global conformational change and minor secondary structure rearrangements were detected. This observation suggests that somehow the active site region was altered. On the other hand, papain suffered a complete inactivation when exposed to those media containing DMF. Fluorescence analyses revealed that an irreversible conformational change took place after 24 h incubation, and a moderate increase in β-sheet and β-turn structures was the most relevant finding when secondary structure was analyzed. The evidences demonstrated that the organic solvents induce a more rigid and compact structure of papain regardless of the organic:buffer ratio investigated. In turn, these modifications affect the active catalytic site in the particular case of MeOH and DMF. These findings were in agreement with the thermo-stability of the enzyme performed after heating at 353 K in all the studied media, that is the presence of ACN did not substantially affect the secondary structure of papain. Nevertheless, the α-helix domain demonstrated to be less thermally stable than the β-sheet domain, turning into aggregated structures after heating, especially in the presence of MeOH and DMF.  相似文献   

9.
Intracellular pH (pHi) is an important modulator of cardiac function. Because it is readily influenced by metabolic processes, pHi is controlled physiologically. Classical models of intracellular pH regulation comprise acid/base transport proteins expressed in the sarcolemma, acting in concert with intracellular buffers. These two processes are coupled via a diffusive movement of protons. Because intracellular H+ buffering is high, Hi+-diffusion occurs through a passive shuttling on intrinsic mobile buffers such as acetylated carnosine, anserine and homocarnosine: low molecular weight imidazole compounds. This mechanism is assisted by carbonic buffer, a system regulated biochemically by the enzyme carbonic anhydrase. Hi+-mobility via the buffer shuttles is low, and this can result in significant pHi non-uniformity under conditions of high proton flux across the sarcolemma or within the cell. Spatial regulation of pHi is complemented by passive H+ permeation between cells through gap junctions. This permeation is also mediated via protonated buffers. The control of pHi is therefore dependent on carrier molecules that spatially shuttle protons within and between cells. In this review, we consider the physiological regulation of Hi+-mobility and permeation, and its relevance to pHi-control in normal and pathophysiological states such as myocardial ischaemia, a clinical condition associated with severe intracellular acidosis.  相似文献   

10.
The importance of the 2′-hydroxyl and 2-amino groups of guanosine residues for the catalytic efficiency of a hammerhead ribozyme has been investigated. The three guanosines in the central core of a hammerhead ribozyme were replaced by deoxyinosine, inosine, and deoxyguanosine, and ribozymes containing these analogues were chemically synthesized. Most of the modified ribozymes are drastically descreased in their cleavage efficiency. However. deletion of the 2-amino group at G8 (replacement with inosine, deoxyguanosine, deoxyinosine) caused little alteration in the catalytic activity relative to that obtained with the unmodified ribozyme. Whereas, deletion of the 2′-amino group at G12 and G5 (replacement with inosine, deoxyinosine, and deoxyguanosine) resulted in ribozymes with drastic decrease in the catalytic activity relative to that obtained with the unmodified ribozyme. In contrast, two uridine residues, U7 and U4, in the ribozyne sequence were replaced by deoxyuridine (dU). The dU4 complex resulted in a decrease in the catalytic rate, with relative cleavage activity that ws about half that observed for the native complex. By comparison, the dU7 complex exhibited a relative cleavage activity within 3.3-fold of that observed with native ribozyme/substrate complex. This result suggests that the 2′-hydroxyl group at U 7 is not essential for activity.

The importance of the 2′-hydroxyl, and 2-amino groups of guanosine residues for the catalytic efficiency of a hammerhead roibozyme has been investigated. Most of the modified rybozymes are drastically decreased in their cleavage efficiency. However, deletion of the 2-amino group at G8 or deletion of the 2′-hydroxyl group at G12 caused little alteration in the catalytic activity relative to that obtained with the unmodified ribozyme. In contrast, two uridine residues, U7 and U4, in the ribozyme sequence were replaced by deoxyuridine (dU). The U4 complex resulted in a decrease in the catalytic rate, with relative cleavage activity that was about half that observed for the native complex.  相似文献   


11.
Abstract: The pH optimum of native adrenal medulla tyrosine hydroxylase activity is shifted from 5.8 to 6.4 by polyanions (heparin, dextran sulphate), salts (NaCl, Na2SO4) and the anionic buffer 2-( N -morpholino)ethanesulphonic acid (MES). Simultaneously, the activity at the optimal pH is increased. Kinetic studies have shown that this activation is associated with a decrease of the apparent K m of the enzyme for the cofactor 6,7-dimethyltetrahydropterin (DMPH4) and an increase in the V max for tyrosine and DMPH4. The K m for the tyrosine remained unchanged. These data have been interpreted in terms of the polyelectrolyte theory. The adsorption of tyrosine hydroxylase on various affinity gels containing heparin, dextran sulphate or unsulphated polymer dextran as ligands indicate that the activation of the enzyme is mediated by electrostatic interactions with the anionic species. The site of electrostatic interaction possesses some specificity since the binding constants are higher for heparin or dextran sulphate than for NaCl or MES buffer. Moreover, 3-( N -morpholino)propanesulphonic acid (MOPS) a slightly structurally different buffer inhibits the enzyme activity whereas N -(2-acetamido)-2-amino-ethanesulphonic acid (ACES) has no effect. A limited proteolytic digestion which preserves the enzymatic activity, destroys the effects of the anions. The isoelectric point and the molecular parameters of tyrosine hydroxylase are markedly altered after limited digestion. It is therefore suggested that the interaction between the hydroxylase and anionic compounds occurs on a part of the protein which is different from the active site and which is lost by proteolysis. This portion of the protein might be involved in regulation of native tyrosine hydroxylase.  相似文献   

12.
The ability to reduce the peroxidase (myeloglobin/H2O2)-generated ABTS•+ [2,2'-azinobis-(3-ethylbenzthiazoline-6-sulfonic acid) radical cation] has been used to rank the antioxidant activity of various agents including dietary flavonoids and chalcones. Surprisingly, we found that in the presence of catalytic concentrations of the phenol B-ring containing flavonoids, apigenin, naringenin and the chalcone phloretin, the formation of the ABTS•+ was initially increased. The enhanced formation of the ABTS•+ was attributed to the peroxidase/H2O2 mediated generation of polyphenolic phenoxyl radicals that were able to co-oxidize ABTS. The relative ABTS•+ generating ability of these dietary polyphenolics correlated with their ability to co-oxidize NADH to the NAD* radical with the resultant generation of superoxide. This pro-oxidant activity was not observed for either luteolin or eriodyctiol, which are B-ring catecholic analogues of apigenin and naringenin, respectively, suggesting that these antioxidants are incapable of the transition metal-independent generation of reactive oxygen species. This pro-oxidant activity of the polyphenolics therefore needs to be taken into account when quantifying antioxidant activity.  相似文献   

13.
We have utilized a commercially available, computer-driven stopped-flow spectrophotometer to rapidly measure the self-dismutation or catalyzed decay of superoxide in aqueous buffers. In the self-dismutation assay, a dimethyl sulfoxide solution of superoxide is mixed in less than 2 ms with an aqueous buffer. The decay of superoxide is monitored directly by its absorbance at 245 nm and the data is processed by computer. By careful purification of the water and the use of metal-free buffers, a decay of superoxide that fits second-order kinetics is obtained without using metal ion chelators in the buffer. The second-order rate constant for superoxide decreased with increasing pH and decreased by a factor of 3.3 by using D2O in place of H2O in the buffer. The rapid mixing time makes it possible to determine rate constants for active superoxide dismutase catalysts at a pH as low as 7. A first-order decay of superoxide is obtained when the aqueous buffer contains bovine Cu/Zn superoxide dismutase or aquo copper(II), which are known catalysts of superoxide dismutation. The rate of superoxide decay was established to be first-order in catalyst. The catalytic rate constant for bovine Cu/Zn superoxide dismutase was determined to be 2.3 x 10(9) M-1 s-1 in H2O and D2O-based buffers and was independent of pH over the range 7-9. Aquo copper(II) gave a catalytic rate constant of 1.2 x 10(8) M-1 s-1, but was ineffective in the presence of EDTA. The catalytic rate constants obtained by stopped-flow kinetics are in excellent agreement with studies carried out by the direct method of pulse radiolysis.  相似文献   

14.
The initial rate and enantioselectivity of enzymatic asymmetric hydrolysis of amino acid esters were examined in methylimidazolium-based ionic liquids with anions including tetrafluoroborate, chloride, bromide and bisulfate and in typical organic solvents. Papain displayed much higher enantioselectivity but lower activity in phosphate buffer solution of 1-butyl-3-methylimidazolium tetrafluoroborate BMIM·BF4 than in other media tested (i.e. E=100, V 0=0.21 mM min-1 in BMIM·BF4, E=2, V 0=0.43 mM min-1 in phosphate buffer, E=14-92, V 0=0.22-0.25 mM min-1 in organic solvents for D,L-phenylglycine methyl ester). The influence of BMIM·BF4 on enzyme activity and enantioselectivity also varied with the substrate and the enzyme used. All of the enzymes assayed showed no activity or low enantioselectivity in the ILs with anions including chloride, bromide and bisulfate.  相似文献   

15.
Recently, we reported a new high-activity biocatalyst for use in organic media termed protein-coated microcrystals (PCMC) (Kreiner et al. [2001] Chem Commun 12:1096-1097). These novel particles consist of water-soluble micron-sized crystalline particles coated with the given biocatalyst(s) and are prepared in a one-step rapid dehydration process. In this study we extended the choice of immobilisation matrix from a simple inorganic salt, K(2)SO(4), to other compounds, both inorganic and zwitterionic, that act as solid-state buffers for biocatalysis in organic media. The catalytic activity of serine proteases subtilisin Carlsberg (SC) and alpha-chymotrypsin (CT) were significantly increased when coated onto the surface of solid-state buffers, as measured in acetonitrile/1wt% H(2)O. SC-PCMC with both organic and inorganic buffer carriers (Na-AMPSO, Na(2)CO(3), and NaHCO(3)) showed a 3-fold greater activity than that observed when using the unbuffered system (PCMC-SC/K(2)SO(4)). In comparison with freeze-dried preparations, this represents an approximately 3,000-fold increase in catalytic activity. Importantly, there is no improvement in catalytic activity upon external addition of any of the solid-state buffers to the reaction mixture. When acting in a solid-state buffer capacity, good buffering capacity was observed with SC-PCMC (3 wt% protein loading) prepared from a 1:1 mixture of AMPSO and AMPSO-Na. Alternatively, increasing the amount of solid-state buffer in the system allows improvement of the buffering. This can be achieved either by decreasing the protein loading of the SC/Na-AMPSO-PCMC or by addition of further external solid-state buffer to the reaction mixture. The catalytic activity of lipase-PCMC prepared from solid-state buffers was found less responsive to immobilisation.  相似文献   

16.
Micellar catalysis of polyphenol oxidase in AOT/cyclohexane   总被引:4,自引:0,他引:4  
The catalytic behaviour of mushroom polyphenol oxidase has been studied in dioctylsulphosuccinate (AOT)/cyclohexane reverse micelles. The steady-state conditions were accomplished up to 20 min and 17 μg protein in the assay towards 4-methylcatechol and no loss of specific activity was observed relative to aqueous medium. The pH activity profile of the enzyme was kept in reverse micelles as in water, showing a plateau between 5 and 6.5. The stability of polyphenol oxidase to pH was also studied and about 20% inactivation was found in reverse micelles relative to aqueous medium at neutral pHs. Moreover there was a decrease of stability at acidic pHs. The optimum Wo obtained was 20 and the enzyme was nearly independent of the surfactant concentration at constant Wo.

Kinetic studies of polyphenol oxidase towards several substrates showed that the substrate inhibition by p-cresol and 4-methylcatechol observed in buffer was not kept in AOT/cyclohexane reverse micelles. Moreover, the Km increased and the catalytic efficiency (V/Km) of the enzyme decreased as the hydrophobicity of substrates was increased.  相似文献   


17.
The chromatographic behaviour of abscisic acid (ABA), indole-3-acetic acid (IAA), phenylacetic acid (PAA), and gibberellins A1, A4, A8, A9, A13 and A20 on columns of Sephadex LH-20 and insoluble poly- N -vinylpyrrolidone (PVP) eluted with buffers of different pH values is described. PVP shows considerable batch differences that must be carefully checked. Chromatography of acidic ethyl acetate-soluble fractions of Scots pine ( Pinus sylvestris L.) extracts at pH 4.5 resulted in great losses of phytohormones, due to poor solubility of the extracts. If the extracts were applied to the column dissolved in buffer of pH 7.5, subsequent elution at pH 4.5 was possible with only small losses. The performance of the chromatographic column was strongly affected by the application volume. A combined PVP/Sephadex LH-20 column eluted at pH 4.5 allows remarkable purification of pine and spruce ( Picea abies (L.) Karst.) extracts, collection of IAA in a fraction that can be directly analyzed by e.g. the indolo-α-pyrone method, and collection of another fraction containing ABA, PAA and probably most of the known C19-gibberellins; whereas the C20-gibberellin A13 is eluted later (with IAA).  相似文献   

18.
The reactivity of horseradish peroxidase (HRP) with water insoluble phenolic compounds has been studied in 1-butyl-3-methylimidazolium tetrafluoroborate ([BMIM][BF4])/water mixtures. The enzyme retained some catalytic activity up to 90% ionic liquid in water at 25 °C only at pH values higher than 9.0. Activity steadily decreased towards neutral and acidic conditions, as judged by 4-aminoantypirin/phenol activity tests. Inhibition of horseradish peroxidase under neutral acidic condition was due to the binding of fluoride anions released from tetrafluoroborate anion to the heme iron as demonstrated by the sharp UV–visible absorption transition diagnostic of the conversion from a five coordinated to a six coordinated high spin ferric heme iron. Thus, reactions with water insoluble phenols were carried out under alkaline reaction conditions and 75% [BMIM][BF4]/water mixture. Under these conditions, the distribution of the reaction products was much narrower with respect to that observed in aqueous buffers or water/dimethylformamide or water/dimethylsulfoxide mixtures, and polymeric species other than dimers were not observed. Technical scale preparations of a novel 4-phenylphenol ortho dimer [2,2′-bi-(4-phenylphenol)] with a high yield of the desired product were obtained.  相似文献   

19.
Hybrid membrane particles from two mutants of Escherichia coli K12, Bv4 and KI1, defective in oxidative phosphorylation, have been prepared, in which ATP-driven membrane energization is restored.

A soluble factor of mutant KI1 was found to have properties similar to parental crude coupling factor, ATPase (EC 3.6.1.3). Membrane particles of this mutant could not be reconstituted by parental coupling factor. Either parental coupling factor, or the soluble factor of mutant KI1 could reconstitute both respiration-driven and ATP-driven energization to membrane particles of mutant BV4 or to parental particles depleted of ATPase. Mutant BV4 was found to be devoid of coupling factor activity, while retaining the ability to hydrolyze ATP. Both mutants possess an ATPase with an altered binding to the membrane.

Mutant KI1 is impaired in respiration-driven amino acid transport, in contrast to mutant BV4.

The three major subunits of parental Escherichia coli ATPase have been isolated and antibodies have been prepared against these subunits. Antibodies against the largestsubunit ( component) or against the intact catalytic subunits ( + β components) inhibit both ATP-Pi exchange in the parent organism as well as ATP hydrolytic activity in parent and mutants. Antibodies against the two other subunits (β or γ components) also inhibit these two reactions, but were found to be less effective. Mutant NI44, which lacks ATPase activity, shows no precipitin lines with anti-, anti-β, anti-γ, or anti-( + β) preparations. In contrast, mutants BV4 and KI1, exhibit cross-reactivity with all of the antisera.  相似文献   


20.
Acetylene was reduced by zinc amalgam in the presence of three synthetic polynuclear complexes: {[Mg2Mo8O22(OMe)6(MeOH)4]−2·[Mg(MeOH)6]2+}6MeOH (I), (Bu4N)2[Fe4S4(SPh)4] (II), [Me4N][VFe3S4Cl3(DMF)3]·2DMF (III) and the iron-molybdenum cofactor of nitrogenase Azotobacter vinelandii MoFe7(S2−)9·homocitrate, FeMo-co (IV). Thiophenol was found to greatly facilitate the reaction in the presence of complexes I, II, IV. The reaction is catalytic and for I and IV proceeds at the amalgam surface. Thiophenol seems to increase the adsorption of the complexes, serving as an electron bridge to transfer electrons to the catalyst. In the case of II a homogeneous reduction of the substrate occurs presumably after the cluster reduction at the surface and with III the catalytic reduction proceeds only under the action of sodium amalgam; no thiophenol cocatalytic action is observed. Relevance to N2 enzymatic reduction is discussed.  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