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.
This study was instituted to investigate in a rat model the effect of topical coadministration of the penetration enhancer oleic acid (10% by volume) and RIMSO-50 (medical grade dimethyl sulfoxide, 50% by volume) on rat skin flap survival. A rectangular abdominal skin flap (2.5 x 3 cm) was surgically elevated over the left abdomen in 40 nude rats. The vein of the flap's neurovascular pedicle was occluded by placement of a microvascular clip, and the flap was resutured with 4-0 Prolene to its adjacent skin. At the end of 8 hours, the distal edge of the flap was reincised to gain access to the clips and the clips were removed. After resuturing of the flap's distal edge to its adjacent skin, the 40 flaps were randomly divided into four groups. Group 1 (control) flaps were treated with 5 g of saline, group 2 (dimethyl sulfoxide) flaps were treated with 2.7 g of dimethyl sulfoxide (50% by volume), group 3 flaps (oleic acid) were topically treated with 0.45 g of oleic acid (10% by volume), and group 4 (dimethyl sulfoxide plus oleic acid) flaps were treated with a mixture of 0.45 g of oleic acid (10% by volume) and 2.7 g of dimethyl sulfoxide (50% by volume) diluted in saline. Each flap was topically treated with 5 ml of drug-soaked gauze for 1 hour immediately after clip removal to attenuate reperfusion injury. Thereafter, drug was applied topically once daily for 4 more days. Digital photographs of each flap were then taken on day 6 and the flaps were then harvested. The percentage of skin survival in each flap was determined by computerized morphometry and planimetry. The mean surviving area of group 3 (oleic acid-treated flaps) was 23.60 +/- 4.19 percent and was statistically higher than that in group 1 (control, saline-treated flaps) at 7.20 +/- 2.56 percent. The mean surviving area of group 2 (dimethyl sulfoxide-treated flaps) at 18.00 +/- 5.23 percent and group 4 (oleic acid- and dimethyl sulfoxide-treated flaps) at 9.90 +/- 3.44 percent did not achieve statistically higher mean surviving areas than controls. A topical solution of oleic acid (10% by volume) caused a statistically significant increase in the survival of rat abdominal skin flaps relative to controls. Dimethyl sulfoxide and the two experimental drugs together did not increase the percentage of flap survival when given as a single 5-ml dose released from a surgical sponge at reperfusion for 1 hour and then daily for a total of 5 days. The reasons for the lack of response are unknown but may have included the technical difficulty of delivering an adequate dose of dimethyl sulfoxide topically and immiscibility between dimethyl sulfoxide and oleic acid. Further studies may be warranted.  相似文献   

2.
The mechanism of sarcoplasmic reticulum (SR) ATPase Mg2+-dependent phosphorylation from Pi was investigated in the presence of 15% v/v dimethyl sulfoxide at pH 6, 20 degrees C, and in the absence of potassium. Measurements of intrinsic fluorescence changes and of 32P-labeled phosphoprotein (*E-P) were in agreement, both at equilibrium and in transient situations. We found that the amount of phosphoenzyme present and its rate of formation depended solely on the concentration of the (Mg X Pi) complex. Up to 6 nmol of phosphate/mg of protein was covalently bound to the enzyme, implying almost complete phosphorylation. Oxygen exchange experiments were also performed in order to allow calculation of the absolute rate constant of *E-P hydrolysis to the noncovalent complex (0.8-1.0 s-1), which differs from the observed rate of enzyme dephosphorylation (0.3-0.5 s-1); in addition, they allowed calculation of the bimolecular rate constant of substrate binding (2-2.4 M-1 s-1). The results demonstrate that in the presence of dimethyl sulfoxide, phosphorylation occurs by the following simple mechanism: relatively slow binding of the neutral substrate (Mg X Pi), with poor affinity, followed by a thermodynamically favorable formation of the covalent bond between phosphate and the possibly hydrophobic active site. The interaction between magnesium and calcium-deprived SR vesicles was studied in the presence of 0-20% v/v dimethyl sulfoxide (or 0-30% v/v glycerol) at pH 7 and 20 degrees C. The presence of either solvent led to the disappearance of the two typical pH-dependent effects we previously characterized for magnesium: loss of the Mg2+-induced spectral shift of tryptophan fluorescence emission and loss of the biphasic pattern displayed by the intrinsic fluorescence rise after addition of calcium to Ca2+-deprived Mg2+-preincubated vesicles. In the absence of solvent, the interaction of magnesium with the calcium-deprived ATPase was also characterized from the point of view of phosphoenzyme formation from ATP or Pi at pH 7 in the absence of potassium: we found that calcium-independent phosphorylation was slower when phosphate was added to SR vesicles preincubated with magnesium that when magnesium was added to vesicles preincubated with phosphate, suggesting that preincubation with magnesium had depleted the phosphate-reactive conformation of the ATPase. A simple reaction scheme for phosphoenzyme formation is described: it implies that the (Mg X Pi) complex is a substrate for this reaction, whereas the Mg2+ itself acts as a pH-dependent, dimethyl sulfoxide sensitive inhibitor of full enzyme phosphorylation.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

3.
The kinetics of tubulin assembly were examined in the absence and presence of dimethyl sulfoxide at 37 degrees C. Inclusion of 1.4 M (10%) dimethyl sulfoxide lowered the critical protein concentration about 8-10-fold, from 9.4 microM in the absence of the organic solvent to 1.1 microM in its presence. This decrease was due solely to an effect on k-, the off rate constant. The on rate constant k+, was essentially unaffected. Another effect of dimethyl sulfoxide was in the nucleation process. The pseudo-first-order rate constant of elongation, kapp (k+[m]), was greatly increased by inclusion of dimethyl sulfoxide. This was due to an increase in the microtubule number concentration, [m]. The microtubules formed in the presence of dimethyl sulfoxide were much shorter than those formed in its absence, accounting for the higher number concentration. The nucleation number, n, was calculated by plots of ln kapp vs. ln c0 or ln t10% vs. ln c0, and the value appeared to be about 4 to 5, although some variability was found. It was shown that a plot of kapp vs. c0 to determine n, is not appropriate because of the inability to distinguish between linear and curved plots in the range of tubulin concentration used in assembly studies.  相似文献   

4.
Information is given concerning two standard buffer solutions suitable as pH references in 30, 40, and 50 mass% dimethyl sulfoxide (DMSO)/H2O mixed solvents at subzero temperatures from -20 to 0 degrees C, with the intention of establishing a pH (designated pH*) scale. The two buffers selected were the ampholytes N,N-bis(2-hydroxyethyl)-2-aminoethane sulfonic acid ("bes") and N-tris(hydroxymethyl)methylglycine ("tricine"), and the reference standard consisted of equal molal quantities of the buffer and its respective sodium salt. The assignment of pH* values was based on measurements of the emf of cells without liquid junction of the type: Pt;H2(g,l atm) /Bes, Na Besate, NaCl / AgCl;Ag and Pt;H2(g,l atm) /Tricine, Na Tricinate, NaCl /AgCl;Ag and the pH* was derived from a determination of K2, the equilibrium constant for the dissociation process (Buffer)+/- in equilibrium with(Buffer)- + H+.  相似文献   

5.
Inhibition of the human erythrocyte calcium pump by dimethyl sulfoxide   总被引:1,自引:0,他引:1  
P.J. Romero   《Cell calcium》1992,13(10):659-667
The action of dimethyl sulfoxide on the human red cell Ca2+ pump was studied in inside-out vesicles. In a high-K+ medium at pH 7.6, the organic solvent inhibited both Ca2+ transport and ATP hydrolysis. Half-maximal effect was obtained with about 2% (v/v). At or below 10% dimethyl sulfoxide, the inhibition was overcome by adding inorganic phosphate or oxalate. In the absence of organic solvent, Ca2+ efflux from Ca(2+)-loaded vesicles consisted of a slow and a fast component whilst in its presence, there appears additionally a leakage component. The size of the latter depended markedly on dimethyl sulfoxide concentration, being about 3% at that level where Ca2+ uptake was half-maximally inhibited. ATP hydrolysis was more sensitive to dimethyl sulfoxide (10%) when free Ca2+ was increased within the millimolar level than when it was raised within the micromolar range. On the other hand, raising Ca2+ with organic solvent greatly stimulated ATP synthesis through ATP-Pi exchange, without reaching saturation. The results suggest that dimethyl sulfoxide blocks the red cell Ca2+ pump by increasing the affinity of the Ca2+ translocating site at the releasing step. They also show that at high concentrations, this solvent increases Ca2+ permeability.  相似文献   

6.
Poly(L -tyrosine) is a random coil in dimethyl sulfoxide. Upon addition of dichloroacetic acid, poly(L -tyrosine) undergoes a conformational transition centered at about 10% dichloroacetic acid. The transition is nearly complete at 20% dichloroacetic acid. Further addition of dichloroacetic acid leads to precipitation of poly(L -tyrosine). We have characterized this transition by optical rotation, viscosity, circular dichroism, and infrared. The optical rotation at 350 nm and the intrinsic viscosity increase sharply to values that are consistent with a transition to the α-helix conformation. The circular dichroism of poly(L -tyrosine) in dimethyl sulfoxide and in dimethyl sulfoxide/dichloroacetic acid (80:20 v/v) agrees with previous reports for random-coil and α-helix conformations, respectively. The infrared spectrum of poly(L -tyrosine) in dimethyl sulfoxide/dichloroacetic acid (80:20 v/v) shows no evidence of β-structure. We conclude that the transition on going from dimethyl sulfoxide to 20% dichloroacetic acid in dimethyl sulfoxide is a coil → α-helix transition. The amide-I band of poly(L -tyrosine) in dimethyl sulfoxide/dichloroacetic acid (80:20) is found to be at 1662 cm?1. It has been suggested that this high frequency may be indicative of a left-handed α-helix. However, this high amide-I frequency is consistent with conformational energy calculations of Scheraga and co-workers. The mechanism of the dichloroacetic acid-induced transition to an α-helix is discussed. Dichloroacetic acid and dimethyl sulfoxide interact strongly and the transition presumably involves a marked decrease in the ability of dimethyl sulfoxide to solvate the peptide backbone and aromatic side chains upon complex formation with dichloroacetic acid.  相似文献   

7.
The cytotoxic and mutagenic effects of X irradiation on a human lymphoblast cell line were examined in the presence of two radioprotective agents which modulate damage to DNA. The cells were treated with X rays alone or in the presence of either dimethyl sulfoxide or cysteamine. Surviving fraction and mutation to trifluorothymidine resistance (tk locus) and to 6-thioguanine resistance (hgprt locus) were measured. Survival was enhanced when the cells were irradiated in the presence of dimethyl sulfoxide; the D0 rose from 58 to 107 rad. However, at both genetic loci the induced mutant fractions were identical in the presence or absence of dimethyl sulfoxide. Survival was enhanced to a greater degree when the cells were irradiated in the presence of cysteamine; the D0 rose from 58 to 200 rad. Cysteamine also protected the cells from X-ray-induced mutation; the frequencies of X-ray-induced mutation at both the tk and hgprt loci were reduced by 50-75%. No protective effects were observed unless dimethyl sulfoxide or cysteamine was present during irradiation. These findings are discussed in terms of the hypothesis that, unlike for cell killing, radiation-induced mutagenesis in human lymphoblast cells is not mediated by the actions of aqueous free radicals, but rather by the direct effects of ionizing radiation.  相似文献   

8.
In order to investigate the mechanism of skeletal muscle relaxation induced by dimethyl sulfoxide, 2-butoxyethanol and dimethyl sulfoxide were examined for their effects on 1) Ca2+ uptake into and efflux from sarcoplasmic reticulum vesicles prepared from rabbit fast skeletal muscle and crayfish tail muscle by the murexide method, 2) ATPase activities of rabbit reticulum vesicles, 3) the isolated phrenic nerve-diaphragm preparation of the rat and 4) crayfish opener muscle preparation. Ca2+ efflux rate from rabbit reticulum vesicles was markedly decreased with increasing concentrations (5-20% v/v) of dimethyl sulfoxide without affecting the maximum Ca2+ uptake by the reticulum. 2-Butoxyethanol showed quite contrary effects. Dimethyl sulfoxide strongly inhibited the activity of basal ATPase rather than of Ca2+-dependent ATPase. 2-Butoxyethanol did not significantly inhibit the activity of basal ATPase, but markedly increased Ca2+-dependent ATPase activity. Antagonisms between dimethyl sulfoxide and caffeine were demonstrated either in contractions of crayfish opener muscles or in the Ca2+ release from crayfish sarcoplasmic reticulum vesicles. These results indicate a possibility that dimethyl sulfoxide reversibly induces skeletal muscle relaxation mainly in the sarcoplasmic reticulum by means of decreasing the rate and the amount of Ca2+ release from the reticulum.  相似文献   

9.
A procedure for fractional determination of soybean sterols is presented. Sterols in lipid extracts were fractionated into four classes, fatty acid esters, the free form, acylated glucosides and non-acylated glucosides, by Florisil column chromatography. Sterol contents in the four classes were determined colorimetrically with ferric chloride-perchloric acid reagent. Before the colorimetry, the fatty acid ester fraction was hydrolyzed with ethanolic KOH, and the sterol was isolated as tomatinide. The free sterol fraction was directly treated with tomatine solution. The tomatinides were dissociated with dimethyl sulfoxide. To avoid the contamination of pigments from the acylated glucoside fraction, the second Florisil column was rinsed with diethyl ether between the elution with the first solvent (0 to 50% diethyl ether in n-heхane) and that with the second solvent (0 to 30% methanol in diethyl ether).  相似文献   

10.
Friend erythroleukemia cells, grown in the presence of dimethyl sulfoxide for 3 d, synthesize unequal amounts of the two chains (alpha and beta) of spectrin with approximately 15-30% more beta than alpha spectrin. When cells were ruptured by nitrogen cavitation, nascent alpha and beta spectrin were found to be associated with a membranous cell fraction and were not detected in the soluble cytoplasmic cell fraction. Nascent membrane-bound spectrin appeared not to be protected by membranes, since it was susceptible to trypsin degradation in the absence of detergent. On fractionation of cells with 1% Triton X-100, more (1.75-fold) nascent spectrin was found in the Triton-soluble fraction than in the Triton-insoluble fraction (cytoskeleton). In the Triton-soluble fraction, there was 55% more nascent beta spectrin than alpha spectrin, while the cytoskeleton contained nearly equal amounts of alpha and beta spectrin. Cells were pulse-labeled with L-[35S]methionine for 2 min and chase incubated for varying periods of time from 15 to 90 min with nonradioactive L-methionine. Radioactive spectrin accumulated in the Triton-soluble fraction for the first 15 min of chase incubation and then dropped by 25% in the next hour. By contrast, the amount of radioactive spectrin in the Triton-insoluble fraction rose gradually for 1 h of the chase period. This indicates that, in Friend erythroleukemia cells, a pool of membrane-bound spectrin containing an excess of the beta polypeptide is used to form the cytoskeletal system which is composed of equal molar amounts of alpha and beta spectrin. The location of spectrin was determined by immunoelectron microscopy. Small amounts of spectrin were detected in cells not treated with dimethyl sulfoxide and in these cells it was located on the surface membrane and within the cytoplasm. On treatment with dimethyl sulfoxide, complex vacuolar structures containing viruses appeared in the cells. In cells treated with dimethyl sulfoxide for 3 d 30% of the spectrin was near the outer membrane and 25% was associated with vacuolar structures, whereas in cells treated for 5 and 7 d the majority of spectrin (57-61%) was located in the vacuolar areas.  相似文献   

11.
Contribution of water to free energy of hydrolysis of pyrophosphate   总被引:2,自引:0,他引:2  
The energy of hydrolysis of phosphate compounds varies depending on whether they are in solution or bound to the catalytic site of enzymes. With the purpose of simulating the conditions at the catalytic site, the observed equilibrium constant for pyrophosphate hydrolysis (Kobsd) was measured in aqueous mixtures of dimethyl sulfoxide, ethylene glycol, or polymers of ethylene glycol. The reaction was catalyzed by yeast inorganic pyrophosphatase at 30 degrees C. All the cosolvents used promoted a decrease of Kobsd. Polymers of ethylene glycol were more effective than dimethyl sulfoxide or ethylene glycol in decreasing Kobsd. The higher the molecular weight of the polymer, the lower the value of Kobsd. A decrease in Kobsd from 346 M (delta G degree obsd = -3.5 kcal mol-1) to 0.1 M (delta G degree obsd = 1.3 kcal mol-1) was observed after the addition of 50% (w/v) poly(ethylene glycol) 8000 to a solution containing 0.9 mM MgCl2 and 1 mM Pi at pH 8.0. The association constants of Pi and pyrophosphate for H+ and Mg2+ were measured in presence of different ethylene glycol concentrations in order to calculate the Keq for hydrolysis of different ionic species of pyrophosphate. A decrease in all the Keq was observed. The results are interpreted according to the concept that the energy of hydrolysis of phosphate compounds depends on the different solvation energies of reactants and products.  相似文献   

12.
The core structures of sodium dodecyl sulfate extracted, pronase digested paired helical filaments of Alzheimer disease were solubilized by heating in dimethyl sulfoxide. Electron microscopy revealed that after heating in dimethyl sulfoxide, intact paired helical filaments were no longer present in the dimethyl sulfoxide soluble fractions or in the insoluble lipofuscin-containing fractions. Enzyme-linked immunosorbent assays of the various fractions with the monospecific antibody A128 to paired helical filaments demonstrated 96% of the immunoreactivity to be in the dimethyl sulfoxide soluble fraction, and only 4% in the dimethyl sulfoxide insoluble fractions. Lyophilization of the dimethyl sulfoxide soluble supernatant and resuspension in water failed to reassociate the paired helical filaments, but did result in an insoluble precipitate. Analysis of the dimethyl sulfoxide solubilized paired helical filament fraction by nuclear magnetic resonance revealed it to be composed of glycolipid in a form that was distinct from similar fractions isolated from normal aged control brains. The aggregation of an altered glycolipid to form paired helical filaments in Alzheimer disease could explain their insolubility.  相似文献   

13.
The hydrolytic cycle of sarcoplasmic reticulum Ca2+-ATPase in the absence of Ca2+ was studied. At pH 6.0, 10 degrees C and in the absence of K+, the enzyme displays a very low velocity of ATP hydrolysis. Addition of up to 15% dimethyl sulfoxide increased this velocity severalfold (from 5-18 nmol of Pi X mg of protein-1 X h-1) and then decreased at higher solvent concentrations. Dimethyl sulfoxide increased both enzyme phosphorylation from ATP and the affinity for this substrate. Maximal levels of 1.0-1.2 nmol of EP X mg of protein-1 and apparent KM for ATP of 5 X 10(-6) M were obtained at a concentration of 30% dimethyl sulfoxide. The same preparation under optimal conditions (pH 7.5, 10 microM CaCl2, 100 mM KCl and no dimethyl sulfoxide at 37 degrees C) displays a velocity of ATP hydrolysis between 8 and 12 X 10(5) nmol of Pi X mg of protein-1 X h-1 while the phosphoenzyme levels varied between 3.5 and 4.0 nmol of EP X mg of protein-1. Enzyme phosphorylation from ATP in the absence of Ca2+ always preceded Pi liberation into the assay media. Two different phosphoenzyme species were formed which were kinetically distinguished by their decomposition rates. The observed steady-state velocity of ATP hydrolysis could be accounted for either by the decay of the fast component or by the simultaneous decomposition of both phosphoenzyme species. The hydrolysis of the phosphoenzyme formed in the absence of Ca2+ was KCl-stimulated and ADP-independent. The rate constant of breakdown was equal to that observed for the phosphoenzyme formed in the presence of Ca2+. It is suggested that the rapidly decaying phosphoenzyme (and possibly both rapidly and slowly decaying species) are intermediates in the reaction cycle of Mg2+-dependent ATP hydrolysis of sarcoplasmic reticulum Ca2+-ATPase and may represent a bypass of Ca2+ activation by dimethyl sulfoxide.  相似文献   

14.
The (Ca2+ + Mg2+)-ATPase of sarcoplasmic reticulum catalyzes the hydrolysis of acetyl phosphate in the presence of Mg2+ and EGTA and is stimulated by Ca2+. The Mg2(+)-dependent hydrolysis of acetyl phosphate measured in the presence of 6 mM acetyl phosphate, 5 mM MgCl2, and 2 mM EGTA is increased 2-fold by 20% dimethyl sulfoxide. This activity is further stimulated 1.6-fold by the addition of 30 mM KCl. In this condition addition of Ca2+ causes no further increase in the rate of hydrolysis and Ca2+ uptake is reduced to a low level. In leaky vesicles, hydrolysis continues to be back-inhibited by Ca2+ in the millimolar range. Unlike ATP, acetyl phosphate does not inhibit phosphorylation by Pi unless dimethyl sulfoxide is present. The presence of dimethyl sulfoxide also makes it possible to detect Pi inhibition of the Mg2(+)-dependent acetyl phosphate hydrolysis. These results suggest that dimethyl sulfoxide stabilizes a Pi-reactive form of the enzyme in a conformation that exhibits comparable affinities for acetyl phosphate and Pi. In this conformation the enzyme is transformed from a Ca2(+)- and Mg2(+)-dependent ATPase into a (K+ + Mg2+)-ATPase.  相似文献   

15.
G Benaim  L de Meis 《FEBS letters》1989,244(2):484-486
In this report it is shown that organic solvents mimic the stimulatory effects of calmodulin and acidic phospholipids on the erythrocyte plasma membrane Ca2+-ATPase. The solvents used were dimethyl sulfoxide (20%, v/v), glycerol (20% v/v), ethylene glycol (20%, v/v) and polyethylene glycol (Mr 6000-8000) (10%, w/v). These solvents increased both the affinity for Ca2+ and the turnover number of the enzyme. The increase in Ca2+ affinity is additive to that achieved with calmodulin. The calcium cooperativity observed in the presence of calmodulin disappears after the addition of dimethyl sulfoxide to the medium. The present data support the proposal that activation of the erythrocyte plasma membrane Ca2+-ATPase is promoted by hydrophobic interactions along the enzyme molecule.  相似文献   

16.
The oxidation of Fe2+ was investigated by electron spin resonance spin trapping techniques with N-t-butyl-alpha-phenylnitrone (PBN) and dimethyl sulfoxide. Under pure oxygen, the spin adduct PBN/.OCH3 was rapidly generated by the addition of Fe2+ (0.2-1.2 mM) into phosphate buffer containing ethylenediaminetetraacetate (EDTA), dimethyl sulfoxide, and PBN at pH 7.4, but it decayed. The decay process of PBN/.OCH3 consists of two components. The fast decay was dependent on Fe2+ concentration. Another was due to destruction of the spin adduct by superoxide anion (.O2-), because superoxide dismutase (SOD) markedly prevented the decay. Catalase decreased the yield of PBN/.OCH3. When EDTA was replaced by diethylenetriaminepentaacetic acid (DTPA), both the generation and decay process of PBN/.OCH3 were slow. SOD and catalase effects were similar to those in EDTA. Fe2+ produced PBN/.OCH3 even in the absence of chelators. We could estimate the kinetic parameters by computer simulation, comparing the Fe2+ oxidation in EDTA with that in DTPA. These results demonstrate that Fe2+ reacts with O2 to generate .O2- and then H2O2, which produces .CH3 by reaction with Fe2+ and dimethyl sulfoxide.(.)OCH3 results from the reaction between .CH3 and O2. The adduct PBN/.OCH3 decays by reaction with Fe2+ and .O2-.  相似文献   

17.
Glutamine-dependent carbamyl phosphate synthetase [EC 2.7.2.9] was purified 1,300-fold from rat ascites hepatoma cells (AH 13) as a multienzyme complex with aspartate transcarbamylase[EC 2.1.3.2] and dihydroorotase[EC 3.5.2.3], using dimethyl sulfoxide, glycerol, and dithiothreitol as stabilizers. The purified complex was essentially homogeneous on agarose-acrylamide composite gel electrophoresis and analytical ultracentrifugation. Its molecular weight was estimated to be about 870,000 by sedimentation equilibrium studies. After alkylation with iodoacetamide or reduction with 0.6% dithiothreitol at 100 degrees, the complex gave a single band on polyacrylamide gel electrophoresis in sodium dodecyl sulfate in a position corresponding to a molecular weight of 210,000. These results indicate that the complex consists of four subunits of similar size.  相似文献   

18.
Purified soluble H+-ATPase from Schizosaccharomyces pombe catalyzes a Pi in equilibrium ATP exchange in the absence of a H+ gradient. When the pH of the assay medium is raised from 5.5 to 8.0 there is a decrease of the ATPase activity and an increase of the rate of Pi in equilibrium ATP exchange. At pH 7.0 the addition of the organic solvent dimethyl sulfoxide (20%, v/v) promotes a decrease of ATPase activity and an increase of the Pi in equilibrium ATP exchange reaction. The effect of the organic solvent on the Pi in equilibrium ATP exchange is related to a decrease of the apparent Km for Pi.  相似文献   

19.
This paper extends a previous study in which a discontinuity in the specific rotation of open chain α-l,4-linked glucopyranosides in the water–dimethyl sulfoxide (H2O–DMSO) system was attributed to a symmetry change about a polymer chain segment. Optical rotation of amylose, cyclohexamylose, methyl β-maltoside, and dextran was measured in the following mixed solvent systems: formamide–dimethyl sulfoxide (F-DMSO), ethylenediamine–dimethyl sulfoxide (E–DMSO), and hexamethylphosphoramide–dimethyl sulfoxide (HMPA–DMSO). Refractive index measurements were used in an attempt to detect hydrogen bonding between solvent components. The specific rotation of amylose corrected for variation in refractive index (CSR), as a function of solvent composition, showed a discontinuity at solvent compositions corresponding to about 1 mole F to 2 moles DMSO and to 1 mole E to at least 8 moles DMSO. A discontinuity in the CSR function of amylose in the H2O-DMSO mixed solvent that occurs at 25°C is not observed at 70°C. The CSR function of methyl-β-maltoside exhibits a discontinuity in solvent composition corresponding to mole ratios between 2F–DMSO and 3F–DMSO. Present results indicate that an amylose chain segment may undergo a symmetry change in solvent compositions corresponding to mole ratios between F–DMSO and F–2DMSO. Our CSR measurements of amylose and model compounds in E–DMSO and HMPA–DMSO do not permit us to distinguish between possible changes in amylose chain segment symmetry and solvent interactions that could affect symmetry properties of the glucopyranose ring.  相似文献   

20.
The soluble hydrogenase of Desulfovibrio vulgaris was purified and some of its properties are described. The molecular weight was determined for the enzyme by sedimentation equilibrium (45,400) and amino acid analysis (44,800). The hydrogenase appears to be a loosely coiled molecule or to have a high axial ratio, which is reflected in an unusually low sedimentation coefficient (2.58S) and a low diffusion coefficient (D 5.85). The molecular weight of the hydrogenase (41,000), as calculated by the Svedberg equation, was in general agreement with the sedimentation equilibrium molecular weight. Amino acid analysis revealed the presence of six halfcystine residues per mole of enzyme and a total of 417 amino acid residues. The specificity of the hydrogenase and its capacity to reduce certain low potential dyes and cytochrome c(3) was studied. Metal analysis of the hydrogenase indicated the presence of 1 mole of ferrous iron per mole of enzyme.  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