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 31 毫秒
1.
The synthesis of Leu-enkephalin selectively 17O-enriched in Gly2 and Gly3 is reported. The 17O-nmr chemical shifts of [17O-Gly2, Leu5]- and [17O-Gly3, Leu5]-enkephalins in H2O are almost identical and independent of the pH. Since hydrogen bonding is the dominant factor governing the chemical shifts of the peptide oxygen, it can be concluded that the hydration state of both oxygens is identical and independent of the pH. The 17O chemical shifts of the [17O-Leu5]-enkephalin terminal carboxyl group at pH approximately 1.9 and 5.6 are very different in H2O but very similar in CH3CN/DMSO (4:1) solution. This suggests that the protonation state of the carboxyl group at both pH values in CH3CN/DMSO solution is the same and consequently that Leu-enkephalin exists in the neutral form at pH approximately 5.6. In this organic mixed solvent system both Gly2 and Gly3 oxygen resonances exhibit a significant shift to high frequency by the same extent (delta delta approximately 30 ppm). It is concluded that both peptide oxygens are not hydrogen bonded to an appreciable extent and that no specific 2----5 hydrogen bonding exists to an appreciable extent. This conclusion is in agreement with the energy of activation for molecular rotation, as determined from T1 measurements, which was found to be almost identical for both [17O-Gly2, Leu5]- and [17O-Gly3, Leu5]-enkephalins in CH3CN/DMSO (4:1) mixed solvent.  相似文献   

2.
Molecular dynamics simulations of Leu-enkephalin in water and DMSO.   总被引:2,自引:0,他引:2       下载免费PDF全文
The structure of Leu-enkephalin (L-Enk) and Met-enkephalin (M-Enk) have frequently been studied, in particular by nuclear magnetic resonance spectroscopy. After more than 20 years of research, it was concluded that enkephalins have no preferred structure in aqueous solution, but that they may have in other solvents. We have performed molecular dynamics simulations of zwitterionic L-Enk in water, and zwitterionic as well as neutral L-Enk dimethyl sulfoxide (DMSO). In water the peptide is very flexible, although there seems to be a preference for compact conformations. In DMSO, the peptide forms a clear salt bridge in the zwitterionic form, but has no preferred conformation in the neutral form. This difference in conformation may provide an explanation for measurements in DMSO in which multiple conformations were found to exist. In this paper we introduce a new formulation for a dihedral angle autocorrelation function, and apply it to study side-chain dynamics in L-Enk. We find that the side-chain dynamics of the large Tyr and Phe residues cannot be adequately sampled in 2.0-ns simulations, while this does seem to be possible for the smaller Leu side chain.  相似文献   

3.
Synthetic peptides reproducing both the native domain around the dibasic cleavage site of prosomatostatin, and mutated sequences there of, previously assayed in site-directed mutagenesis experiments, have been studied by CD in different solvent systems, such as water, TFE/H2O, MeCN/H2O and aqueous SDS, in order to ascertain the ability of each solvent to stabilize secondary structural motifs. A combination of deconvolution methods and empirical calculations, that allow subtraction of the contributions due to unordered structures from the spectra, suggests that mainly two distinct families of ordered conformers containing alpha-helix and/or structurally different beta-turns are present in solution, the relative stability of the different conformers depending on the nature of the solvent. The presence of beta-turns is in line with a previous NMR study in DMSO and DMSO/H2O. Comparison of the CD spectra in aqueous SDS of peptides undergoing processing with a sequence not processed in vivo shows that only the latter possesses a stable and detectable alpha-helix population. This observation suggests that the structuration involving beta-turns but no alpha-helix, which was observed by CD both in SDS and organic solvent/H2O mixtures at high water contents, might be of biological significance. The similarity of this structuration to molecular models obtained from NMR data in DMSO and DMSO/H2O is discussed.  相似文献   

4.
The spectroscopic, electrochemical and coreactant electrogenerated chemiluminescence (ECL) properties of Ir(ppy)3 (where ppy = 2‐phenylpyridine) have been obtained in aqueous buffered (KH2PO4), 50 : 50 (v/v) acetonitrile–aqueous buffered (MeCN–KH2PO4) and 30% trifluoroethanol (TFE) solutions. Tri‐n‐propylamine was used as the oxidative‐reductive ECL coreactant. The photoluminescence (PL) efficiency (ϕem) of Ir(ppy)3 in TFE (ϕem ≈ 0.029) was slightly higher than in 50 : 50 MeCN–KH2PO4 (ϕem ≈ 0.0021) and water (ϕem ≈ 0.00016) compared to a Ru(bpy)32+ standard solution in water (Φem ≈ 0.042). PL and ECL emission spectra were nearly identical in all three solvents, with dual emission maxima at 510 and 530 nm. The similarity between the ECL and PL spectra indicate that the same excited state is probably formed in both experiments. ECL efficiencies (ϕecl) in 30% TFE solution (ϕecl = 0.0098) were higher than aqueous solution (ϕecl = 0.00092) system yet lower than a 50% MeCN–KH2PO4 solution (ϕecl = 0.0091). Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

5.
Ascidiacyclamide (ASC), cyclo(-Ile1-Oxz2-d-Val3-Thz4-)2 (Oxz=oxazoline and Thz=thiazole) has a C2-symmetric sequence, and the relationships between its conformation and symmetry have been studied. In a previous study, we performed asymmetric modifications in which an Ile residue was replaced by Gly, Leu or Phe to disturb the symmetry [Doi et al. (1999) Biopolymers49, 459-469]. In this study, the modifications were extended. The Ile1 residue was replaced by Gly, Ala, aminoisobutyric acid (Aib), Val, Leu, Phe or d-Ile, and the d-Val3 residue was replaced by Val. The structures of these analogs were analyzed by X-ray diffraction, 1H NMR and CD techniques. X-Ray diffraction analyses revealed that the [Ala1], [Aib1] and [Phe1]ASC analogs are folded, whereas [Val1]ASC has a square form. These structures are the first examples of folded structures for ASC analogs in the crystal state and are similar to the previously reported structures of [Gly1] and [Phe1]ASC in solution. The resonances of amide NH and Thz CH protons linearly shift with temperature changes; in particular, those of [Aib1], [d-Ile1] and [Val3]ASCs exhibited a large temperature dependence. DMSO titration caused nonlinear shifts of proton resonances for all analogs and largely affected [d-Ile1] and [Val3]ASCs. A similar tendency was observed upon the addition of acetone to peptide solutions. Regarding peptide concentration changes, amide NH and Thz CH protons of [Gly1]ASC showed a relatively large dependence. CD spectra of these analogs indicated approximately two patterns in MeCN solution, which were related to the crystal structures. However, all spectra showed a similar positive Cotton effect in TFE solution, except that of [Val3]ASC. In the cytotoxicity test using P388 cells, [Val1]ASC exhibited the strongest activity, whereas the epimers of ASC ([d-Ile1] and [Val3]ASCs), showed fairly moderate activities.  相似文献   

6.
The octapeptide Lys-Arg-Asn-Lys-Asn-Asn-Ile-Ala (Arg4 in the human sequence) is the C-terminal part of porcine oxyntomodulin, an endogeneous peptide which is a potent inhibitor of stimulated acid secretion. This octapeptide exhibits the whole range of biological activities of the parent hormone. In the present work we report an 1H n.m.r. investigation of the conformational properties of the octapeptides of pig and human sequences in dimethylsulfoxide-d6 (DMSO) solution. The various resonances were assigned on the basis of two-dimensional COSY and NOESY experiments. Other experiments such as (i) temperature and concentration dependence of the amide proton chemical shifts, (ii) effects of ionic strength, (iii) comparison of the spectra with different analogues, were performed. We showed that in DMSO, the conformation of the octapeptide is directly related to the ionisation state of the C-terminus carboxyl group of alanine. In carboxylic state, the peptide adopts an extended conformation, while in the carboxylate state the four last residues (Asn-Asn-Ile-Ala) are involved in a type II beta-turn structure probably locked by a salt bridge between the carboxyl group of Ala8 and the epsilon ammonium group of Lys4 (or the guanidinium group of Arg4). These observations provide an insight into the possible conformational tendencies of this peptide in biological media.  相似文献   

7.
The syntheses of nitrosyl–dimethylsulfoxide–ruthenium(II) complexes with general formula mer-[RuCl3(L)(DMSO)(NO)] (L=DMSO or CD3CN) is reported. The mer-[RuCl3(DMSO)2(NO)] (1) complex was obtained from the reaction of [RuCl3(NO)] with the sulfoxide ligand in acetone. The mer-[RuCl3(CD3CN)(DMSO)(NO)] (2) compound was obtained from mer-[RuCl3(DMSO)2(NO)] maintained in deuterated acetonitrile. These data suggest a slow kinetic reaction due the low lability of the DMSO ligand coordinated to the {RuII–NO+} species. The crystal and molecular structures of (1) and (2) have been determined from X-ray studies. Crystal data: for (1), monoclinic, P21/c, a=8.8340(2) Å, b=12.0230(3) Å, c=13.7064(4) Å, β=114.546(2)°, Z=4, R1=0.0429; for (2), monoclinic, P21/n, a=10.0180(7) Å, b=9.5070(7) Å, c=13.3340(9) Å, β=102.264(4)°, Z=4, R1=0.0472. The spectroscopic characterization of (1), in solid state (infrared spectrum) and in solution (nuclear magnetic resonance and cyclic voltammetry) is also described.  相似文献   

8.
The new iridium(III) complex, imidazolium[trans(DMSO,imidazole)tetrachloroiridate(III)], (I) (DMSO=dimethyl sulfoxide), and the orange form of [(DMSO)(2)H][trans(DMSO)(2)tetrachloroiridate(III)], (II) have been prepared and characterized, both in the solid state and in solution, by X-ray diffraction and by various physicochemical techniques. Single crystal X-ray diffraction studies point out that complex (II) is isomorphous to the ruthenium(III) analogue, [(DMSO)(2)H][trans-RuCl(4)(DMSO)(2)], (III). Crystallographic data are the following: a=16.028(2) A, b=24.699(3) A, c=8.262(1) A, in space group Pbca (Z=8) for (imidazolium)[trans(DMSO,imidazole)tetrachloroiridate(III)], (I); and a=9.189(2) A, b=16.511(4) A, c=14.028(3) A, beta=100.82(2) degrees in space group P2/n (Z=4) for [(DMSO)(2)H][trans(DMSO)(2)tetrachloroiridate(III)], (II). Visible absorption spectra show that both complexes are stable for several days, at pH 7.4, at room temperature. No significant chloride hydrolysis is observed, even at high temperature (70 degrees C), over 24 h. The extreme stability of these iridium(III) complexes within a physiological buffer was further assessed by (1)H NMR; in addition, cyclic voltammetry measurements evidenced a high stability of the oxidation state +3. Preliminary biological studies show that both complexes do not bind appreciably bovine serum albumin nor inhibit significantly the proliferation of representative human tumor cell lines, suggesting that hydrolysis of coordinated chlorides is a crucial feature for the biological properties and the antitumor activity of the parent ruthenium(III) complexes.  相似文献   

9.
Parameters of the EPR signals of monomeric dinitrosyl-iron complexes with 1H-1,2,4-triazole-3-thiol (DNIC-MT), obtained by treating MT+ferrous iron in DMSO solution with gaseous NO, have been compared with those of the crystalline monomeric DNIC-MT with tetrahedral structure. Dissolved DNIC-MT were characterized by the isotropic EPR signal centered at g=2.03 with half-width of 0.7 mT and quintet hyperfine structure when recorded at ambient temperature or the anisotropic EPR signal with g( perpendicular)=2.045, g( parallel)=2.014 from frozen solution at 77 kappa, Cyrillic. DNIC-MT in crystalline state showed the structure-less symmetrical singlet EPR signal centered at g=2.03 and half-width of 1.7 mT at both room and liquid nitrogen temperature. The Lorentz shape of this signal indicates the strong exchange interaction between these complexes in the DNIC-MT crystal. Being dissolved in DMSO the crystalline sample of DNIC-MT demonstrated the EPR signal typical for DNIC-MT, obtained by treating MT+ferrous iron in DMSO solution with gaseous NO. Low spin (S=1/2) d(9) electron configuration of DNIC-MT with tetrahedral structure (formula [(MT-S(.))(2)Fe(-1)(NO(+))(2)](+)) was suggested to be responsible for the signal of DNIC-MT in crystalline state. Dissolving of the crystals of DNIC-MT may result in the change of their spatial and electronic structure, namely, tetrahedral structure of the complexes characterized by low spin d(9) electronic configuration transforms into a plane-square structure with d(7) electronic configuration and low spin S=1/2 state (formula [(MT- S(-))(2)Fe(+)(NO(+))(2)](+)). The latter was suggested to be characteristic of other DNICs with various thiol-containing ligands in the solutions. The proposed mechanism of these DNICs formation from ferrous iron, thiol and NO shows that the process could be accompanied by the ionization of NO molecules to NO(+) and NO(-) ions in the complexes. Detailed analysis of the shape of the EPR signals of these complexes provided additional information about the exchange interaction typical for DNIC-MT in crystals.  相似文献   

10.
Leu-enkephalin, [D-Ala2]Leu-enkephalin, and [D-Ala2]Leu-enkephalinamide (agonists) and [L-Ala2]Leu-enkephalin (inactive analogue) bind to lipid bilayer consisting of phosphatidylcholine and phosphatidylserine. The conformations that these compounds assume, once bound to perdeuterated phospholipid bilayer, have been shown to be unique, as shown by the transferred nuclear Overhauser effect (TRNOE) of 1H NMR spectroscopy. In addition, their location in the bilayer was analyzed by TRNOE in the presence of spin-labeled phospholipids. These analyses showed a clear relationship between the activity and the peptide-membrane interaction. The three active peptides, when bound to membranes, adopt the same conformation, characterized by a type II' beta-turn around Gly3-Phe4 and a gamma-turn around Gly2 (or D-Ala2). The inactive analogue, [L-Ala2]Leu-enkephalin, displayed a completely different TRNOE pattern corresponding to a different conformation in the membrane-bound state. The tyrosine residue of the active compounds is not inserted into the interior of membrane, but it is inserted into the bilayer for the L-Ala2 analogue. According to these results, [L-Ala2]Leu-enkephalin may be explained to be inactive because the mode of binding to the membranes is different from that of active compounds.  相似文献   

11.
To establish an industrially feasible reaction process, the effect of dimethylsulfoxide (DMSO) added to an aqueous solution on the hydrolysis of lipase was investigated using fluorescent substrates. Several lipases from microorganisms were improved in their hydrolysis activities against 4-methylumbelliferyl oleate by DMSO. Variation was found in the effect of DMSO depending on the species of lipase. After the high stability of the lipase from Pseudomonas fluorescens in DMSO solution was confirmed, hydrolysis by this lipase of four acyl-4-methylumbelliferones was studied kinetically at different DMSO concentrations. DMSO added to an aqueous solution increased the Vmax of this lipase for a substrate with strong hydrophobicity, and decreased that value for a substrate with an opposite property. On the other hand, DMSO had a very small effect on Km for each substrate. A fluorometric study suggested that DMSO induced a change of the chemical environment that surrounded tryptophan residues of the lipase. Such conformational change would be one of the causes of the DMSO-induced alteration of its reactive property. These results suggest that the addition of DMSO may be a novel method of 'solvent engineering' of this enzyme.  相似文献   

12.
The endophytic fungus Coniothyrium sp. was isolated from leaves of Quercus robur. Fermentation of this fungus on solid rice medium yielded two new furoic acid derivatives ( 1 and 2 ) and two additional known compounds. The structures of the new compounds were determined by extensive analysis of 1D and 2D nuclear magnetic resonance spectra as well as high-resolution mass spectrometry data. Compound 1 , containing three aromatic chromophores attached by rotatable sigma bonds and a chirality center in benzylic position, was found to be a scalemic mixture with an excess of the (S) enantiomer, the absolute configuration of which was elucidated as by the solution time-dependent density functional theory-electronic circular dichroism approach. The ωB97X/TZVP PCM/MeCN and SOGGA11-X/TZVP SMD/MeCN methods were used for geometry reoptimization to reproduce the solution conformational ensemble. All isolated compounds were tested for their cytotoxicity but proved to be inactive.  相似文献   

13.
A new concept in cryopreservation solution design was developed that focuses on the use of an intracellular-type, hypothermic maintenance medium coupled with additives that inhibit cryopreservation-induced apoptosis. HypoThermosol' (HTS), a hypothermic (4 degrees C) maintenance medium utilized in the long-term storage of cell, tissue, and organ systems, was tested for cryoprotective capability on a renal cell line (Madin-Darby Canine Kidney cells). HTS and HTS derivatives were tested against conventional cell culture medium (Dulbecco's Minimal Essential medium, DME) as the cryoprotectant carrier solution because (1) cells are exposed to an extended state of hypothermia during the freeze-thaw process, and (2) HTS is designed to protect cells exposed to a hypothermic state. Cells separately cryopreserved in either HTS or DME + 5% dimethyl sulfoxide (DMSO) yielded equivalent 24-h postthaw survival (approximately 30%) and 5-d recovery (approximately 90%). Cells cryopreserved in CryoStor CS 5, a HTS derivative containing 5% DMSO, yielded approximately 75% 24-h postthaw survival and recovery to 100% within 3 d. DNA gel electrophoresis was performed to determine the mechanisms of cell death contributing to cryopreservation failure. Cells preserved in DME (DMSO-free) died primarily through necrosis, whereas cells preserved in either DME + 5% DMSO, HTS, or CryoStor CS 5 died through a combination of apoptosis and necrosis. This observation led to the inclusion of an apoptotic inhibitor designed to improve cryopreservation outcome. MDCK cells cryopreserved in CryoStor CS 5 supplemented with an apoptotic inhibitor (Caspase I Inhibitor V), hereafter termed CryoStor CS 5N, resulted in a 24-h postthaw survival and recovery rate exceeding that of any other cryoprotective solution tested (85%). We conclude that: (1) the use of HTS (a dextran-based, intracellular-type solution) without DMSO can yield postthaw viability equivalent to that of standard DMSO-based cryopreservation methods, (2) postthaw viability can be significantly increased through the use of an intracellular-type solution in conjunction with DMSO, (3) the use of HTS allows for cryopreservation to be accomplished with reduced levels of cryoprotectants, and (4) the regulation of apoptosis is essential for the improvement of cryopreservation outcome.  相似文献   

14.
Transport of [tyrosyl-3,5-3H]enkephalin-(5-L-leucine) [( 3H]Leu-enkephalin) across the blood-brain barrier was studied in the adult guinea pig, by means of vascular perfusion of the head in vivo. The unidirectional transfer constant (Kin) estimated from the multiple-time uptake data for [3H]Leu-enkephalin ranged from 3.62 X 10(-3) to 3.63 X 10(-3) ml min-1 g-1 in the parietal cortex, caudate nucleus, and hippocampus. Transport of [3H]Leu-enkephalin was not inhibited by unlabelled L-tyrosine (the N-terminal amino acid) at a concentration as high as 5 mM, or by the inhibitor of aminopeptidase activity bacitracin (2 mM), suggesting that there was no enzymatic degradation of peptide at the blood-brain barrier. By contrast, 2 mM unlabelled Leu-enkephalin strongly inhibited the unidirectional blood-to-brain transport of [3H]Leu-enkephalin by 74-78% in the parietal cortex, caudate nucleus, and hippocampus. The tetrapeptide tyrosyl-glycyl-glycyl-phenylalanine (without the C-terminal leucine of Leu-enkephalin), at a concentration of 5 mM, caused a moderate inhibition ranging from 15 to 29% in the brain regions studied, whereas the tetrapeptide glycyl-glycyl-phenylalanyl-leucine (without the N-terminal tyrosine) at 5 mM was without effect on Leu-enkephalin transport. Unidirectional brain uptake of Leu-enkephalin was not altered in the presence of naloxone at a concentration as high as 3 mM (1 mg/ml), suggesting that there is no binding of Leu-enkephalin to opioid receptors at the blood-brain barrier. It is concluded that there is a specific transport mechanism for Leu-enkephalin at the blood-brain barrier in the guinea pig.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
The line width of the NMR signal of water protons in solutions of native actomyosin and actomyosin denatured by heat, acetone or urea was measured over the temperature range from -10 degrees to below the freezing point. The line widths of the water band which increased exponentially with decreasing temperature were compared with each other and also with those of the corresponding control solution without actomyosin. The line broadening observed for native actomyosin solution on lowering the temperature was significantly smaller than that for heat-denatured actomyosin solution. This difference implies that this signal is sensitive to conformational perturbations of the protein. In addition, the temperature dependence of the line width for heat-, acetone-, or urea-denatured actomyosin solution was similar to that for the corresponding control solution. These phenomena can be interpreted in terms of the state of water associated with the hydrophobic and hydrophilic residues. Similar NMR studies of actomyosin solution containing dimethyl sulfoxide (DMSO) or dimethylformamide (DMF) showed that DMSO and DMF prevent the formation of ice crystals until about -70 degrees, suggesting that the cryoprotective effects of DMSO and DMF are due to the change in the state of water described above. These differences in temperature dependence between the sample and control solutions are well-correlated with the viscosity of the solution. This correlation is useful for elucidation of the mechanism of the protein denaturation.  相似文献   

16.
The effect of organic solvents on carboxypeptidase Y (a serine carboxypeptidase from yeast)-catalyzed hydrolysis of amino acid ester and peptide synthesis from N-acyl amino acid ester and amino acid amide was investigated.

The Km value of ester hydrolysis increased with an increase in the solvent content. Dioxane was the most effective and dimethyl sulfoxide (DMSO) the least, whilst Kcat showed a tendency to increase slightly in N, N-dimethylformamide (DMF) and DMSO. For dioxane and acetonitrile (MeCN) a maximum was observed.

In peptide formation from Fua-Phe-OEt and Gly-NH2, dioxane and MeCN supported high product yield at molar fractions smaller than ca. 0.05 but the yield decreased significantly at higher fractions, although a relatively constant selectivity (ratio of the peptide bond formed to the ester consumed) was maintained. DMSO gave rather low peptide yields and selectivity even at lower molar fractions. DMF showed an intermediate tendency.

An apparent saturation parameter of the amine component was evaluated and the dissociation constant of a complex between acyl-enzyme and amino acid amide (Kn), as well as the rate constant of aminolysis exerted by the amino acid amide bound correctly on the enzyme (Kn), was calculated by initial rate analysis of peptide formation. In contrast to Km values, Kn decreased with increasing concentrations of organic cosolvent. while a suppressive effect was observed (except for DMSO) on the Kn parameter.

Effects of the solvent practically immiscible in water was also studied by use of the enzyme physically “immobilized” on glass beads.  相似文献   

17.
The reaction of the monoalkyl complex trans-[Pt(DMSO)2Cl(CH3)] with a large variety of heterocyclic nitrogen bases L, in chloroform solution, leads to the formation of uncharged complexes of the type [Pt(DMSO)(L)Cl(CH3)], containing four different groups coordinated to the metal center. Only two out of the three different possible isomers were detected in solution. These two trans(C,N) and cis(C,N) species can be unambiguously identified through 1H NMR spectroscopy. For the trans(C,N) isomers, average values of 2JPtH=75±4 Hz and 3JPtH=36±4 Hz have been observed for the coordinated methyl and DMSO ligands, respectively. In the case of the cis(C,N) isomers, these values increase to 2JPtH=83±2 Hz, and decrease to 3JPtH=26±3 Hz due to the mutual exchange of ligands in trans position to CH3 and DMSO. In the case of bulky asymmetric ligands, such as quinoline, 2-quinolinecarboxaldehyde, 2-methylquinoline, 5-aminoquinoline, 2-phenylpyridine and 2-chloropyridine, slow rotation of the hindered group around the Pt---N bond makes the coordinated DMSO ligand prochiral. NMR experiments have shown that the first reaction product is the trans(C,N) isomer as a consequence of the very fast removal of one DMSO ligand by the nitrogen bases from the starting complex trans-[Pt(DMSO)2Cl(CH3)]. This trans kinetic product undergoes a geometrical conversion into the more stable cis(C,N) isomer through the intermediacy of fast exchanging aqua-species. The rate of isomerization and the relative stability of the two isomers depends essentially on the rate of aquation and on the steric congestion imposed by the new L ligand on the metal.  相似文献   

18.
D L Layman  S W Jacob 《Life sciences》1985,37(25):2431-2437
The absorption and excretion of dimethyl sulfoxide (DMSO) were studied in Rhesus monkeys (Macaca mulatta) given daily oral doses of 3 gms DMSO/kg B.W. for 14 days. DMSO and its major metabolite, dimethyl sulfone (DMSO2), were measured in serum, urine and feces by gas-liquid chromatography. DMSO was absorbed rapidly, reached a steady state blood level after 1 day and then was cleared from blood within 72 hrs after ending treatment. Serum DMSO declined in a linear fashion on semilogarithmic coordinates as described by second order kinetics. It had a half-life of 16 hrs. DMSO2 appeared in blood within 2 hrs and reached a steady state concentration after 4 days of treatment. DMSO2 was cleared from blood about 120 hrs after DMSO administration was stopped. Its half-life in blood was calculated to be 38 hrs. Urinary excretion of unmetabolized DMSO and DMSO2 accounted for about 60% and 16%, respectively, of the total ingested dose. Neither DMSO nor DMSO2 was detected in fecal samples. However, when added to fecal samples, DMSO was degraded rapidly. Although dimethyl sulfide (DMS) was not measured, some DMSO was metabolized to this compound because of the particular sweetness of breath of the monkeys. We conclude that the absorption of DMSO by monkeys is similar to that for humans, but that its conversion to DMSO2 and urinary elimination are more rapid in monkeys.  相似文献   

19.
A solid state analysis of the cyclic octapeptide c(-Pro(1)-Pro-Phe-Phe-Ac(6)c-Ile-D-Ala-Val(8)-) (C8-CLA), containing the Pro-Pro-Phe-Phe sequence, followed by the bulky helicogenic C(alpha,alpha)-dialkylated 1-aminocyclohexane-1-carboxylic acid (Ac(6)c) residue and a D-Ala residue in position 7, has been carried out by x-ray diffraction.The crystals, grown from a DMSO solution, are monoclinic, space group P2(1) with a = 13.458(3) A, b = 19. 404(5) A, c = 21.508(4) A, and beta = 90.83(6) degrees, with two independent cyclic molecules in the asymmetric unit, two DMSO molecules, and three water molecules. The structure has been solved using the half and bake procedure by Sheldrick, and refined to final R1 and wR2 indices of 0.0613 and 0.1534 for 9867 reflections with I > 2sigma(I).This cyclic peptide, a deletion analogue of the naturally occurring cyclic nonapeptide cyclolinopeptide A [c(Pro-Pro-Phe-Phe-Leu-Ile-Ile-Leu-Val), CLA] has been designed to study the influence of the ring size reduction on the conformational behavior of CLA and more in general to obtain structural information on asymmetric cyclic octapeptides.The compound exhibits, in the solid state, a "banana-twisted" conformation with a cis peptide bond located between the two proline residues. Five intramolecular H bonds stabilize the structure: one type VIa beta-turn, two consecutive type III/I beta-turns, one gamma-turn, and one C(16) bend.The structure has also been compared with either the solution structure previously reported by us and obtained by nmr and computational analysis, and with solid state structural data reported in the literature on cyclic octapeptides.  相似文献   

20.
Potential methods for cryopreservation of mouse spermatozoa are freeze-drying, desiccation, and suspension in EGTA Tris-HCl buffered solution (ETBS: 50 mM NaCl, 50 mM EGTA, and 10 mM Tris-HCl). To determine the duration that mouse spermatozoa suspended in ETBS-based solutions could retain their normal characteristics without freezing, spermatozoa collected from the cauda epididymis were suspended in ETBS or in ETBS supplemented with the antioxidants, dimethyl sulfoxide (DMSO), or DL-alpha-tocopherol acetate (Vitamin E acetate; VEA) diluted in DMSO, then held at ambient temperature (22-24 degrees C) for up to 9 days. When oocytes were injected with spermatozoa preserved in ETBS alone, activation rates of oocytes and chromosome integrity at the first cleavage metaphase decreased at 1 day (P < 0.001) and 2-4 days (P < 0.01) following treatment. When oocytes were injected with spermatozoa preserved in ETBS supplemented with DMSO or VEA/DMSO, chromosome integrity did not decrease significantly (through 9 days of preservation). Although DMSO maintained sperm chromosome integrity more effectively than VEA/DMSO up to 2-4 days (91 and 67%, normal karyotypes in DMSO and VEA/DMSO, respectively), VEA/DMSO helped to maintain the ability of spermatozoa to activate oocytes, but did not enhance the maintenance of sperm chromosome integrity. These results suggested that deterioration of spermatozoa preserved in ETBS alone was delayed by supplementation with antioxidants.  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