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.
A 1H NMR study of urocanic acid dissolved in water/dimethyl sulfoxide mixtures together with potentiometric determinations of its two successive acidities in the same solutions reveal that a proton is transferred from the imidazolium ring to the carboxylate group when the amount of dimethyl sulfoxide is increased. An analysis of the potentiometric data by means of the classical Hammett linear relationship allowed the four microscopic acidity constants to be determined. At 20 degrees C, there are equal fractions of the tautomeric species, i.e. urocanic acid (AH0) and its zwitterion (AH+-) in the mixture containing 34% dimethyl sulfoxide by weight (congruent to 0.1 in mole fraction), while 50% dimethyl sulfoxide by weight (congruent to 0.2 in mole fraction) is necessary to completely bias the equilibrium: AH+- in equilibrium AH0 toward the molecular form AH0.  相似文献   

2.
The rates of hydrolysis of acetyl phosphate in the presence of 0.1 M NaOH and of ATP in the presence of either 1 M HCl or 1 M NaOH were measured at different temperatures and in the presence of different concentrations of the organic solvents dimethyl sulfoxide or ethylene glycol. Under all conditions tested, there was a progressive increase in the rate constant of hydrolysis of both phosphate compounds as the water activity of the medium was decreased by the addition of organic solvents. At 25 degrees C, substitution of 70% of the water of the medium by dimethyl sulfoxide promoted an increase of two orders of magnitude in the rate constant of acetyl phosphate hydrolysis. In the presence of 80% and 90% dimethyl sulfoxide the rate of acetyl phosphate hydrolysis increased by more than two orders of magnitude and was so fast that it could not be measured with the method used. The effect of organic solvents on the rate of ATP hydrolysis was less pronounced than that observed for acetyl phosphate hydrolysis. At 30 degrees C, substitution of 90% of water by an organic solvent promoted a 4-6-fold increase of the rate of ATP hydrolysis. Acceleration of either acetyl phosphate or ATP hydrolysis rates was promoted by a decrease in both activation energies (Ea) and in entropies of activation delta S. The data obtained are discussed with reference to the mechanism of catalysis of enzymes involved in energy transduction such as the Ca2+-ATPase of sarcoplasmic reticulum and the F1-ATPase of mitochondria.  相似文献   

3.
Treatment of amino acids, peptides, and proteins with aqueous solution of dimethyl sulfoxide (Me2SO) and hydrochloric acid (HCl) resulted in the oxidation of methionine to methionine sulfoxide. In addition to methionine, SH groups are also oxidized, but this reaction proceeds after a lag period of 2 h. Other amino acids are not modified by aqueous Me2SO/HCl. The reaction is strongly pH-dependent. Optimal conditions are 1.0 M HCl, 0.1 M Me2SO, at 22 degrees C. The reaction exhibits pseudo-first order kinetics with Kobs = 0.23 +/- 0.015 M-1 min-1 at 22 degrees C. Incubation of methionine sulfoxide with dimethyl sulfide and HCl resulted in the conversion of methionine sulfoxide to methionine. This reaction is fast (t1/2 = 4 min at room temperature) and quantitative at relatively anhydrous condition (i.e. at H2O:concentrated HCl:dimethyl sulfide ratio of 2:20:1). Quantitative conversions of methionine sulfoxide back to methionine are obtained in peptides and proteins as well, with no observable other side reactions in amino acids and proteins. The wide applications of this selective oxidation and reduction of methionine residues are demonstrated and discussed.  相似文献   

4.
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   

5.
Previous work from our laboratory supports an important role for aquaporins (AQPs), a family of water channel proteins, in bile secretion by hepatocytes. To further define the pathways and molecular mechanisms for water movement across hepatocytes, we directly assessed osmotic water permeability (Pf) and activation energy (Ea) in highly purified, rat hepatocytes basolateral membrane vesicles (BLMV) and canalicular membrane (CMV) vesicles by measuring scattered light intensity using stopped-flow spectrophotometry. The time course of scattered light for BLMV and CMV fit well to a single-exponential function. In BLMV, Pf was 108 +/- 4 mum.s-1 (25 degrees C) with an Ea of 7.7 kcal/mol; in CMV, Pf was 86 +/- 5 mum.s-1 (25 degrees C) with an Ea of 8.0 kcal/mol. The AQP blocker, dimethyl sulfoxide, significantly inhibited the Pf of both basolateral (81 +/- 4 mum.s-1; -25%) and canalicular (59 +/- 4 mum.s-1; -30%) membrane vesicles. When CMV were isolated from hepatocytes treated with dibutyryl cAMP, a double-exponential fit was needed, implying two functionally different vesicle populations; one population had Pf and Ea values similar to those of CMV from untreated hepatocytes, but the other population had a very high Pf (655 +/- 135 mum.s-1, 25 degrees C) and very low Ea (2.8 kcal/mol). Dimethyl sulfoxide completely inhibited the high Pf value in this second vesicle population. In contrast, Pf and Ea of BLMV were unaltered by cAMP treatment of hepatocytes. Our results are consistent with the presence of both lipid- and AQP-mediated pathways for basolateral and canalicular water movement across the hepatocyte plasma membrane barrier. Our data also suggest that the hepatocyte canalicular membrane domain is rate-limiting for transcellular water transport and that this domain becomes more permeable to water when hepatocytes are exposed to a choleretic agonist, presumably by insertion of AQP molecules. These data suggest a molecular mechanism for the efficient coupling of osmotically active solutes and water transport during canalicular bile formation.  相似文献   

6.
Zhang L  Li X  Xu X  Zeng F 《Carbohydrate research》2005,340(8):1515-1521
A (1-->3)-beta-D-glucan having (1-->6) branching (L-FV-IB) from Lentinus edodes in water was degraded into seven fractions of different molecular weights by ultrasonic irradiation, and each was further fractionated into three parts, by precipitation from water into acetone at room temperature. The weight-average molecular weight (M(w)), radius of gyration ((z)(1/2)), and intrinsic viscosity ([eta]) of lentinan and its fractions in 0.9% NaCl aqueous solution and dimethyl sulfoxide (Me(2)SO) were determined by size-exclusion chromatography combined with multi-angle laser light scattering (SEC-LLS), LLS, and viscometry. Analysis of M(w), [eta], and (z)(1/2) in terms of known theory for worm-like chains yielded 2240 +/- 100 nm(-1), and 100 +/- 10 nm for molar mass per unit contour length (M(L)), and persistence length (q), respectively, corresponding with theoretical data for triple-helical chains. The [alpha](D) of lentinan in water-Me(2)SO mixtures indicated an order-disorder transition. The results indicated that lentinan exists as a triple helix in 0.9% NaCl aqueous solution and as a single flexible chain in Me(2)SO. Assays in vivo and in vitro against the growth of Sarcoma 180 solid tumor as well as the colorimetric 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) method for lentinan showed that the triple-helix sample exhibited a relatively high inhibition ratio. Interestingly, the triple-helix lentinan with M(w) of 1.49 x 10(6) exhibited the highest antitumor activity in vivo, having an inhibition ratio (xi) of 49.5%, close to that of 5-fluorouracil (xi = 50.5%), whereas the bioactivity (xi = 12.3%) of its single flexible chains almost disappeared. The triple-helix conformation plays an important role in enhancing the antitumor effects of lentinan.  相似文献   

7.
The purpose of this study was to examine whether antioxidants attenuate endotoxin-induced microvascular hyper-permeability for macromolecules in the hamster cheek pouch. Twenty-two adult male Syrian hamsters were anesthetized, and a removable plastic chamber was placed in the cheek pouch to observe and collect suffusate from the microvasculature. Fluorescent-labeled dextran (FITC-D; mol wt 150,000) was injected intravenously, and changes in the number of microvascular leaky sites and microvascular clearance of FITC-D were measured in five groups: saline control (group 1, n = 4), endotoxin (0.1 mg/ml) suffusion for 120 min (group 2, n = 6), endotoxin plus dimethyl sulfoxide (1.0 g/kg iv; group 3, n = 4), endotoxin plus allopurinol (30 mg/kg ip; group 4, n = 4), and endotoxin plus dimethyl sulfoxide and allopurinol (group 5, n = 4). The number of leaky sites and the FITC-D clearance were significantly higher in group 2 [45 +/- 18 (SD) sites/cm2 and 20 +/- 6 X 10(-6) ml/min, respectively; P less than 0.01] than in group 1 (7 +/- 6 sites/cm2 and 7 +/- 5 X 10(-6) ml/min), group 3 (9 +/- 5 sites/cm2 and 8 +/- 2 X 10(-6) ml/min), group 4 (11 +/- 7 sites/cm2 and 9 +/- 4 X 10(-6) ml/min), and group 5 (11 +/- 6 sites/cm2 and 7 +/- 1 x 10(-6) ml/min). The leaky sites appeared predominantly in postcapillary venules. There was a positive and significant correlation between the number of leaky sites and FITC-D clearance.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Golden hamster embryo cells were exposed to 137Cs gamma rays in the presence or absence of dimethyl sulfoxide at both 310 and 77 K. Dimethyl sulfoxide gave significant protection against cell killing at both 310 and 77 K. The extent of radioprotection with 1.28 M dimethyl sulfoxide at 77 K was 85-89% of the lethal effects observed in the absence of dimethyl sulfoxide at 310 K; the dose-modifying factor was 5.7. Dimethyl sulfoxide also exerted protected against gamma-ray-induced DNA single-strand breaks and chromosomal aberrations with a maximum protection of 80-100% at a dimethyl sulfoxide concentration of 1.28 M at 77 K. At 77 K, H atoms, ion holes, and electrons can migrate through frozen cells but OH radicals cannot diffuse. Thus the protective effects of dimethyl sulfoxide against cell killing, chromosomal aberrations, and DNA single-strand breaks at 77 K may be due to the scavenging of H atoms or other ions, rather than OH radicals.  相似文献   

9.
Resonance Raman spectroscopy has been used to define active site structures for oxidized Mo(VI) and reduced Mo(IV) forms of recombinant Rhodobacter sphaeroides biotin sulfoxide reductase expressed in Escherichia coli. On the basis of (18)O/(16)O labeling studies involving water and the alternative substrate dimethyl sulfoxide and the close correspondence to the resonance Raman spectra previously reported for dimethyl sulfoxide reductase (Garton, S. D., Hilton, J., Oku, H., Crouse, B. R., Rajagopalan, K. V., and Johnson, M. K. (1997) J. Am. Chem. Soc. 119, 12906-12916), vibrational modes associated with a terminal oxo ligand and the two molybdopterin dithiolene ligands have been assigned. The results indicate that the enzyme cycles between mono-oxo-Mo(VI) and des-oxo-Mo(IV) forms with both molybdopterin dithiolene ligands remaining coordinated in both redox states. Direct evidence for an oxygen atom transfer mechanism is provided by (18)O/(16)O labeling studies, which show that the terminal oxo group at the molybdenum center is exchangeable with water during redox cycling and originates from the substrate in substrate-oxidized samples. Biotin sulfoxide reductase is not reduced by biotin or the nonphysiological products, dimethyl sulfide and trimethylamine. However, product-induced changes in the Mo=O stretching frequency provide direct evidence for a product-associated mono-oxo-Mo(VI) catalytic intermediate. The results indicate that biotin sulfoxide reductase is thermodynamically tuned to catalyze the reductase reaction, and a detailed catalytic mechanism is proposed.  相似文献   

10.
Biologically active cyclic tetrapeptides, usually found among fungi metabolites, exhibit phytotoxic or cytostatic activities that are likely to be governed by specific conformations adopted in solution. For conformational studies and drug design, there is a strong interest in using fast and reliable methods to determine correctly the conformational population of cyclotetrapeptides. We show here that standard molecular mechanics computational approach gives satisfactory results. The method was validated step by step by experimental data either obtained after synthesis and NMR analysis, or found in the literature. The cyclo(Gly)(4), cyclo(Ala)(4), cyclo(Sar)(4), and cyclo(SarGly)(2) peptides were used to evaluate the prediction of the peptide backbone conformation, and the detailed conformational analysis of tentoxin, a natural phytotoxic cyclotetrapeptide in which N-alkylated peptide bonds alternate with regular secondary ones, was used to validate the computation of conformers proportions. From the knowledge of an initial cyclic primary structure and of the D or L configuration of the amino acids, we show that it is possible to determine the exact orientation of carbonyl groups and to predict the nature of conformers present in solution. The proportion of each conformer can be inferred from a statistical thermodynamics approach by using the potential energy values of each conformer, computed by molecular mechanics methods with the TRIPOS force field, which allowed us to account for the solvent. The solvent contribution was processed by two different methods according to the nature of the interactions: whether through the dielectric constant introduced in the electrostatic potential, when interaction with solute molecules are weak or negligible, or through the computation of free energy of solvation using the algorithm SILVERWARE for solvents explicitly interacting with the solute. When applied to tentoxin, this conformational analysis yielded results in very good agreement with the experimental data reported by Pinet et al. (Biopolymers, 1995, Vol. 36, pp. 135-152), on both the nature of existing conformers and their relative proportions, whatever the nature of the considered solvent.  相似文献   

11.
Biophysical characteristics of the plasma membrane, such as osmotic sensitivity and water and cryoprotectant permeability are important determinants of the function of spermatozoa after cryopreservation. A series of experiments was conducted with rhesus macaque spermatozoa at 23 degrees C to determine their: (1) cell volume and osmotically inactive fraction of the cell volume; (2) permeability coefficients for water and the cryoprotectants dimethyl sulfoxide, glycerol, propylene glycol, and ethylene glycol; (3) tolerance to anisosmotic conditions; and (4) motility after a one step addition and removal of the four cryoprotectants. An electronic particle counter and computer aided semen analysis were used to determine the cell volume and permeability coefficients, and motility, respectively. Rhesus spermatozoa isosmotic cell volume was 27.7+/-3.0 microm3 (mean+/-SEM) with an osmotically inactive cell fraction of 51%. Hydraulic conductivity in the presence of dimethyl sulfoxide, glycerol, propylene glycol, and ethylene glycol was 1.09+/-0.30, 0.912+/-0.27, 1.53+/-0.53, and 1.94+/-0.47 microm/min/atm, respectively. Cryoprotectant permeability was 1.39+/-0.31, 2.21+/-0.32, 3.38+/-0.63, and 6.07+/-1.1 (x10(-3)cm/min), respectively. Rhesus sperm tolerated all hyposmotic exposures. However, greater than 70% motility loss was observed after exposure to solutions of 600 mOsm and higher. A one step addition and removal of all four cryoprotectants did not cause significant motility loss. These data suggest that rhesus sperm are tolerant to hyposmotic conditions, and ethylene glycol may be the most appropriate cryoprotectant for rhesus sperm cryopreservation, as it has the highest permeability coefficient of the tested cryoprotectants.  相似文献   

12.
A Lyophilized preparation of L-2-halo acid dehalogenase was not only stable but also catalytically active in anhydrous dimethyl sulfoxide, toluene, and other organic solvents. 2-Halo acids with long alkyl (C(5)-C(16)) or aromatic (phenyl and benzyl) side chains were inert in water but dehalogenated effectively in anhydrous dimethyl sulfoxide by the lyophilized enzyme. Long chain 2-haloalkanoic acids such as 2-bromohexadecanoic acids were better as substrate than short-chain halo acids (e.g., 2-chloropropanoic acid). The dehalogenation proceed with inversion of C(2) configuration to produce the corresponding (2R)-2-hydroxy acids in anhydrous dimethyl sulfoxide in the same way as found in water.  相似文献   

13.
The sugar trehalose is produced in some organisms that survive dehydration and desiccation, and it preserves the integrity of membranes in model systems exposed to dehydration and freezing. Dimethyl sulfoxide, a solute which permeates membranes, is added to cell suspensions in many protocols for cryopreservation. Using a surface forces apparatus, we measured the very large, short-range repulsion between phosphatidylcholine bilayers in water and in solutions of trehalose, sorbitol, and dimethyl-sulfoxide. To the resolution of the technique, the force-distance curves between bilayers are unchanged by the addition of trehalose or sorbitol in concentrations exceeding 1 kmol · m-3. A relatively small increase in adhesion in the presence of trehalose and sorbitol solutions may be explained by their osmotic effects. The partitioning of trehalose between aqueous solutions and lamellar phases of dioleylphosphatidylcholine was measured gravimetrically. The amount of trehalose that preferentially adsorbs near membrane surfaces is at most small. The presence of dimethyl sulfoxide in water ( 1:2 by volume) makes very little difference to the short-range interaction between deposited bilayers, but it sometimes perturbs them in ways that vary among experiments: free bilayers and/or fusion of the deposited bilayers were each observed in about one-third of the experiments.  相似文献   

14.
Different substances such as dimethyl sulfoxide, tetramethylene sulfoxide, 2-pyrollidone, and the naturally occurring compatible solute betaine enhance PCR amplification of GC-rich DNA templates with high melting temperatures. In particular, cyclic compatible solutes outperform traditional PCR enhancers. We therefore investigated the effects that cyclic naturally occurring ectoine-type compatible solutes and their synthetic derivatives have on melting temperature of double-stranded DNA (dsDNA) and on PCR amplification of different templates. L-ectoine, betaine, and derivatives of L-ectoine decreased, whereas beta-hydroxyectoine increased, the melting temperature of dsDNA. The ability to decrease the melting temperature was greatest for homoectoine, a new synthetic derivative of l-ectoine. Furthermore, compatible solutes, especially homoectoine, enhanced PCR amplification of GC-rich DNA (72.6% GC content; effective range: 0.1-0.5M).  相似文献   

15.
The conformational spaces of five oligomers of tetrahydrofuran-based carbopeptoids in chloroform and dimethyl sulfoxide were investigated through nine molecular dynamics simulations. Prompted by nuclear magnetic resonance experiments that indicated various stable folds for some but not all of these carbopeptoids, their folding behaviour was investigated as a function of stereochemistry, chain length and solvent. The conformational distributions of these molecules were analysed in terms of occurrence of hydrogen bonds, backbone torsional-angle distributions, conformational clustering and solute configurational entropy. While a cis-linkage across the tetrahydrofuran ring favours right-handed helical structures, a trans-linkage results in a larger conformational variability. Intra-solute hydrogen bonding is reduced with increasing chain length and with increasing solvent polarity. Solute configurational entropies confirm the picture obtained: they are smaller for cis- than for trans-linked peptides, for chloroform than for dimethyl sulfoxide as solvent and for shorter peptide chains. The simulations provide an atomic picture of molecular conformational variability that is consistent with the available experimental data.  相似文献   

16.
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   

17.
Dimethyl sulfoxide (2%), hexamethylene bisacetamide (4mM) and butyric acid (2mM) were potent inducers of erythrodifferentiation in Friend erythroleukemia cell lines, 5–18 and C19TK. Hydrocortisone (1μM) markedly inhibited dimethyl sulfoxide induced hemoglobin production in both 5–18 and C19TK cells. d,1-Propranolol (25–50μM) markedly inhibited both dimethyl sulfoxide and hexamethylene bisacetamide induced erythrodifferentiation in 5–18 cells but not in C19TK cells. Addition of either hydrocortisone or propranolol as late as 48 hrs after dimethyl sulfoxide addition still resulted in significant inhibition of hemoglobin synthesis in 5–18 cells. Although the mechanism of action of propranolol is not known, modulation of the β adrenergic receptor is apparently not involved since practolol failed to inhibit either dimethyl sulfoxide or hexamethylene bisacetamide induced erythrodifferentiation in 5–18 cells nor did isoproternol induce hemoglobin synthesis.  相似文献   

18.
Embryonic development of the sea urchin after low-temperature preservation   总被引:1,自引:0,他引:1  
The sea urchin embryos were cooled to -196 degrees by two-step freezing with the use of 1-1.5 M dimethyl sulfoxide as a cryoprotectant. The embryos were equilibrated with the cryoprotectant for 20-30 min at 0 +/- 2 degrees. At -7 degrees ice crystallization was induced and the embryos were cooled to -38-42 degrees at a rate of 6-8 degrees /min. The embryos were then transferred into liquid nitrogen. The embryos were thawed in a water bath at 19 degrees. No less than 90% of the embryos frozen at the stages of blastula, gastrula, or pluteus were capable of recovery and normal development. The length of cryopreservation did not affect the survival of the embryos.  相似文献   

19.
Neutrophils stimulated with phorbol myristate acetate (PMA) in the presence of the spin trap 5,5-dimethyl-1-pyrroline 1-oxide (DMPO), dimethyl sulfoxide, and diethylenetriaminepentaacetic acid (DETAPAC) fail to generate hydroxyl radical (.OH), detected as the methyl spin-trapped adduct of DMPO (2,2,5-trimethyl-1-pyrrolidinyloxyl, DMPO-CH3), unless ferric salts (Fe3+) are also added (Britigan, B. E., Rosen, G. M., Chai, Y., and Cohen, M. S. (1986) J. Biol. Chem. 261, 4426-4431). Even then, .OH formation wanes in spite of ongoing superoxide (O2-.) production. In contrast, ferric salt supplementation of a hypoxanthine/xanthine oxidase O2-. generating system containing DETAPAC produces continual .OH, suggesting that neutrophils limit the formation of this free radical. To evaluate this hypothesis, neutrophil cytoplasts (largely devoid of granules but able to generate O2-.) were stimulated with PMA in the presence of Fe3+, DETAPAC, dimethyl sulfoxide, and DMPO. This resulted in continual production of DMPO-CH3. In the presence of dimethyl sulfoxide, HL-60 (promyelocytic) cells differentiate into cells similar in morphology and O2-. generating capacity to neutrophils. However, their granules lack the iron-binding protein lactoferrin (LF). Ferric salt supplementation of HL-60 cells stimulated with PMA yielded an EPR spectrum similar to cytoplasts. Supernatant obtained following PMA-induced neutrophil degranulation (which releases LF extracellularly) suppressed DMPO-CH3 formation by the hypoxanthine/xanthine oxidase/Fe3+/DETAPAC system. Anti-LF antibody, but not anti-transferrin antibody, prevented stimulated neutrophil supernatant inhibition of hypoxanthine/xanthine oxidase/Fe3+/DETAPAC-mediated .OH formation. Similarly, neutrophils stimulated with PMA in the presence of Fe3+, DETAPAC, and anti-LF antibody (but not anti-transferrin antibody) demonstrated continual formation of .OH. Neutrophil degranulation of LF limits Fe3+-catalyzed .OH formation which in vivo could protect tissue from possible .OH-mediated injury.  相似文献   

20.
Heats of solution for myelin basic protein have been determined using microcalorimetry. All aqueous systems studied yielded negative heats of solution; in contrast, trifluoroethanol produced a small positive heat of solution, while reaction with dimethyl sulfoxide was strikingly exothermic. The heat of interaction for native myelin basic protein with 8 M urea at pH 4.0, 29 degrees C, was found to be -79 +/- 16 kcal/mol. The significance of these results in terms of the protein's structural organization 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