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 187 毫秒
1.
The effect of four general anesthetic molecules, i.e., chloroform, halothane, diethyl ether and enflurane, on the properties of a fully hydrated dipalmitoylphosphatidylcholine (DPPC) membrane is studied in detail by long molecular dynamics simulations. Furthermore, to address the problem of pressure reversal, the effect of pressure on the anesthetic containing membranes is also investigated. In order to ensure sufficient equilibration and adequate sampling, the simulations performed have been at least an order of magnitude longer than the studies reported previously in the literature on general anesthetics. The results obtained can help in resolving several long-standing contradictions concerning the effect of anesthetics, some of which were the consequence of too short simulation time used in several previous studies. More importantly, a number of seeming contradictions are found to originate from the fact that different anesthetic molecules affect the membrane structure differently in several respects. In particular, halothane, being able to weakly hydrogen bound to the ester group of the lipid tails, is found to behave in a markedly different way than the other three molecules considered. Besides, we also found that two changes, namely lateral expansion of the membrane and increasing local disorder in the lipid tails next to the anesthetic molecules, are clearly induced by all four anesthetic molecules tested here in the same way, and both of these effects are reverted by the increase in pressure.  相似文献   

2.
We have developed a technique to determine the concentration of volatile anesthetics (halothane, isoflurane and sevoflurane) in blood that is a modification of a method used for volatile anesthetics in Krebs solution. Methylene chloride was the internal standard and chloroform was used to extract the volatile anesthetic from blood. The congealed blood proteins were separated from the chloroform solvent (containing anesthetic) using a two-compartment vial that filtered out the proteinaceous material during centrifuging. Recovery averaged 102%. Linearity was excellent (r = 0.992-0.999) in the 50-600, 50-300 and 50-300 microg/mL range for halothane, isoflurane and sevoflurane, respectively. Intra-day and inter-day precisions were likewise excellent, with relative standard deviations <5.3 and <7.1%, respectively. Accuracy ranged from 0.8 to 9.5% of the estimated theoretical value. Extracted anesthetic in chloroform solvent was stable over 4-5 days, with <3% variability. The time from obtaining the blood sample to determination of the concentration from the chromatographic peak was 15 min or less.  相似文献   

3.
Ion transport from one aqueous phase (W1) to another (W2) across a planar bilayer lipid membrane (BLM) in the presence of inhalation anesthetics was electrochemically investigated. In the absence of inhalation anesthetics in the BLM system, no ion transport current flowed between W1 and W2 across the BLM. When inhalation anesthetics such as halothane, chloroform, diethyl ether and trichloroethylene were added to the two aqueous phases or the BLM, the ion transport current quite clearly appeared. When the ratio of the concentration of KCl or NaCl in W1 to that in W2 was varied, the zero current potential across the BLM was shifted. By considering the magnitude of the potential shift, we concluded that the ion transport current can be predominantly ascribed to the transport of Cl(-) across the BLM. Since the dielectric constants of these anesthetics are larger than that of the inner hydrophobic domain of the BLM, the concentration of hydrophilic electrolyte ions in the BLM increases with the increase in the dielectric constant of the inner hydrophobic domain caused by addition of these anesthetics. These situations lead to an increase in the ion permeability coefficient.  相似文献   

4.
The structural features of general anesthetic binding sites on proteins are being examined using a defined model system consisting of a four-alpha-helix bundle scaffold with a hydrophobic core. Previous work suggested that halothane binding to the four-alpha-helix bundle was improved by (1) introducing a cavity into the hydrophobic core and (2) substituting a methionine side-chain in place of an alpha-helical heptad e position leucine. In this study, the ability of the general anesthetics chloroform and 2,2,2-trichloroethanol to bind to the hydrophobic core of the four-alpha-helix bundle (Aalpha2-L38M)2 is explored. The halogenated alkane chloroform binds with a dissociation constant (Kd) = 1.4 +/- 0.2 mM, whereas 2,2,2-trichloroethanol binds with a Kd = 19.5 +/- 1.2 mM. The affinity of both general anesthetics for the hydrophobic core of the four-alpha-helix bundle approximates their whole animal effective concentration in 50% of test subjects' (EC50) values, as shown previously for halothane. Tryptophan phosphorescence decay rates at 77 K are accelerated by a factor of 4.5 by both bound halothane and chloroform, indicating that the heavy-atom effect is responsible for a portion of the observed fluorescence quenching. Because heavy-atom effects are operative only at short distances, the findings indicate that these general anesthetics are binding in the vicinity of the indole rings of W15 in the hydrophobic core of the four-alpha-helix bundle scaffold. The results indicate that chloroform, halothane and 2,2,2-trichloroethanol may occupy the same sites on protein targets.  相似文献   

5.
The surface activities of six uncharged local anesthetics, dibucaine (DC), bupivacaine (BC), lidocaine (LC), mepivacaine (MC), benzocaine (BzC), and benzyl alcohol (BzOH) were investigated by taking surface tension measurements of their aqueous solutions. The surface densities of the uncharged anesthetics were calculated from the application of thermodynamic equations to the surface tension data. The surface activities for uncharged anesthetics became higher in the order of their hydrophobicities, BzOH相似文献   

6.
Cryogenic oven-trapping (COT) with capillary GC has been successfully applied to analysis of chloroform, dichloromethane, trichloroethylene, diethyl ether, the components of solvent thinner (ethyl acetate, benzene, n-butanol, toluene, and others), xylene isomers, cyanide, ethanol, hexanes, general anesthetics, and styrene in human body fluids. This COT-GC technique was compared with headspace solid-phase microextraction (SPME) coupled with GC for some volatile organic compounds (VOC); for all compounds compared the sensitivity achieved using COT-GC was more than ten times higher than for headspace SPME-GC. The COT-GC method is recommended for widespread use in forensic and environmental toxicology, because it is simple, requires no special GC operations, and yet enables high sensitivity and high resolution.  相似文献   

7.
Mistic (membrane integrating sequence for translation of integral membrane protein constructs) comprises the four-alpha-helix bundle scaffold found in the transmembrane domains of the Cys-loop receptors that are plausible targets for general anesthetics. Nuclear magnetic resonance (NMR) studies of anesthetic halothane interaction with Mistic in dodecyl phosphocholine (DPC) micelles provide an experimental basis for understanding molecular mechanisms of general anesthesia. Halothane was found to interact directly with Mistic, mostly in the interfacial loop regions. Although the presence of halothane had little effect on Mistic structure, (15)N NMR relaxation dispersion measurements revealed that halothane affected Mistic's motion on the microsecond-millisecond time scale. Halothane shifted the equilibrium of chemical exchange in some residues and made the exchange faster or slower in comparison to the original state in the absence of halothane. The motion on the microsecond-millisecond time scale in several residues disappeared in response to the addition of halothane. Most of the residues experiencing halothane-induced dynamics changes also exhibited profound halothane-induced changes in chemical shift, suggesting that dynamics modification of these residues might result from their direct interaction with halothane molecules. Allosteric modulation by halothane also contributed to dynamics changes, as reflected in residues I52 and Y82 where halothane introduction brought about dynamics changes but not chemical shift changes. The study suggests that inhaled general anesthetics could act on proteins via altering protein motion on the microsecond-millisecond time scale, especially motion in the flexible loops that link different alpha helices. The validation of anesthetic effect on protein dynamics that are potentially correlated with protein functions is a critical step in unraveling the mechanisms of anesthetic action on proteins.  相似文献   

8.
Inhalation anesthetics currently in clinical use, such as halothane, methoxyflurane, enflurane, isoflurane, etc., are polar hydrophobic molecules, except nitrous oxide, which is an apolar and weak anesthetic, incapable of inducing surgical stage anesthesia. Experimental data are accumulating that these potent amphipathic inhalation anesthetics preferentially bind membranes and macromolecules on the surface at clinical concentrations. The anesthetic binding to lipid membranes in the low concentration range is characterized by a saturable curve approaching to a limiting value. When the anesthetic concentration s greatly increased above the clinical range, the binding starts to exceed the limiting saturation value. Our model for anesthetic binding to membranes consists of two parts: Langmuir-type adsorption to the membrane surface at the low concentration range and penetration into the hydrophobic core at the high concentration range. The present communication provides a statistical-thermodynamic basis to analyze this twostep interaction. An expression is derived for membrane capacitance as a function of anesthetic concentration, which explains the experimental data well. Binding parameters of anesthetics are estimated according to the theory.This study was supported by NIH grants GM 25716 and GM 26950, and by the Medical Research Service of the Veterans Administration.  相似文献   

9.
The chromatographic retention and membrane activity relationships of local anesthetics were studied to address the possible mechanisms for structure specificity and inflammation-associated decrease of their effects. Five representative drugs (3 mM for each) were reacted with 1,2-dipalmitoyl-sn-glycero-3-phosphocholine liposomes in 25 mM potassium phosphate buffer (pH 5.9-7.9, containing 100 mM NaCl and 0.1 mM EDTA) for 10 min at 37 degrees C and the membrane fluidity changes were analyzed by measuring fluorescence polarization with 1,6-diphenyl-1,3,5-hexatriene. Their capacity factors were determined on octadecyl-, octyl- and phenyl-bonded silica columns with a mobile phase consisting of 25 mM potassium phosphate buffer (pH 5.9-7.9, containing 100 mM NaCl and 0.1 mM EDTA)-methanol (30:70, v/v) at a flow rate of 1.0 ml/min and at a column temperature of 37 degrees C and diode-array detection. Mepivacaine, prilocaine, lidocaine, ropivacaine and bupivacaine fluidized membranes in increasing order of intensity, which agreed with their clinical potency. The relative degree of membrane fluidization correlated with that of retention on an octadecyl stationary phase more significantly than the other phases. Both membrane-fluidizing effects and capacity factors decreased by lowering the reaction and mobile phase pH, being consistent with the hypothesis that anesthetic potency is reduced in inflammation because of tissue acidity. Reversed-phase liquid chromatography appears to be useful for estimating the structure-specific and pH-dependent membrane-fluidizing effects of local anesthetics.  相似文献   

10.
Adsorption of n-pentane, triethylamine, diethyl ether, acetonitrile and chloroform has been investigated on pyrogenic alumina (S=140 m2 g−1). The results of our studies have shown the presence of active sites on the surface of pyrogenic alumina with irreversible adsorption of electron-donating molecules and CHCl3 and the dependence of energetic surface properties on electronic structure of adsorbate, quantity of adsorbed substance and hydration degree of the surface. On the hydrated oxide surface the water molecules screen the active sites of the surface, which resulted in changing of interaction energy of adsorbent-adsorbate and decreasing the region of irreversible adsorption of organic bases and CH-acid. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

11.
The surface tension and pH of aqueous solutions of three hydrochloric acid (HCl) - uncharged anesthetic (mepivacaine (MC), bupibacaine (BC) and dibucaine (DC)) mixtures were measured as a function of total molality and composition of local anesthetic in order to investigate the competitive surface-adsorption of uncharged and charged local anesthetics. The behavior of the surface tension versus total molality and pH versus total molality curves remarkably changed at the composition corresponding to an equimolar mixture. The pH measurements showed that uncharged and charged forms coexisted only at compositions more than the equimolar mixture. The partitioning quantities of respective uncharged and charged anesthetics into the surface-adsorbed film were estimated from their surface densities calculated thermodynamically. The greater quantity of uncharged anesthetics existed in the adsorbed film at the coexisting composition, that is, the uncharged anesthetics adsorbed more preferentially than charged ones. The relative ease with which uncharged anesthetics transferred into the surface-adsorbed film was proportional to the hydrophobicities and well correlated the anesthetic potencies. At compositions in the vicinity of physiological pH (ca. 7.4), the bulk solution is more abundant in charged anesthetics than uncharged ones, whereas the uncharged molecules is conversely more abundant in the surface region. The present results clearly imply that the surface-active molecule of local anesthetic in the physiological pH is the uncharged form and the partitioning is greatly dependent on the hydrophobicity among the anesthetics.  相似文献   

12.
The captive bubble technique in combination with axisymmetric drop shape analysis (ADSA-CB) and with micro gas chromatography is used to study the dynamics of dissolution of different gases and vapors in water in situ. The technique yields the changes in the interfacial tension and bubble volume and surface. As examples, the dissolution of methanol and hexane vapors, inhaled anesthetic vapors, and gases, that is, diethyl ether, chloroform, isoflurane, enflurane, sevoflurane, desflurane, N2O, and xenon, and as nonimmobilizers perfluoropentane and 1,1,2-trichloro-1,2,2-trifluoro-ethane (R113) were investigated. The examination of interfacial tension-time and bubble volume-time functions permits us to distinguish between water-soluble and -insoluble substances, gases, and vapors. Vapors and gases generally differ in terms of the strength of their intermolecular interactions. The main difference between dissolution processes of gases and vapors is that, during the entire process of gas dissolution, no surface tension change occurs. In contrast, during vapor dissolution the surface tension drops immediately and decreases continuously until it reaches the equilibrium surface tension of water at the end of dissolution. The results of this study show that it is possible to discriminate anesthetic vapors from anesthetic gases and nonimmobilizers by comparing their dissolution dynamics. The nonimmobilizers have extremely low or no solubility in water and change the surface tension only negligibly. By use of newly defined molecular dissolution/diffusion coefficients, a simple model for the determination of partition coefficients is developed.  相似文献   

13.
Complex formation equilibria in binary mixtures of chloroform with dipropyl ether (PE), diisopropyl ether (IPE), methyl tert-butyl ether (MBE), tetrahydrofuran (THF). 1,4-dioxane (DOX), acetone (AC), and methyl acetate (MA) have been analyzed in detail using several association models. Vapor-liquid equilibria, excess enthalpy and excess heat capacity data for these mixtures have been correlated using a multiproperty global fitting procedure. The thermodynamic properties for chloroform +PE, +IPE, +MBE, +AC, and +MA are best correlated using the ideal association model while for chloroform +THF and +DOX the best model is an athermal solvation model where the Flory-Huggins expression for the species activity coefficients is considered. The model parameters, i.e., the equilibrium constant, enthalpies and heat capacities of complexation, were found to be reliable, well representing the chloroform-oxygenated solvent H-bonded complexes. A detailed discussion is given on the test proposed by McGlashan and Rastogi to decide whether the solution contains only 11 complexes or 21 complexes as well. The complex formation equilibria in chloroform mixtures is compared to those previously examined for halothane (2-bromo-2-chloro-1,1,1-trifluoroethane) mixed with the same oxygenated solvents. It was found that the H-bonds formed by halothane are stronger than those formed by chloroform.  相似文献   

14.
del Olmo M  Zafra A  Jurado AB  Vilchez JL 《Talanta》2000,50(6):1141-1148
Bisphenol A (BPA) in the presence of phenol is determined using a method based on first-derivative spectrofluorimetry. The proposed method involves a micro liquid–liquid extraction of sodium chloride saturated water samples with diethyl ether followed by direct fluorimetric analysis of extracts. The excitation spectra of both compounds in diethyl ether are recorded between 200 and 290 nm, with the emission wavelength at 306 nm. The first-derivative spectra were calculated, measuring the analytical signal for BPA at 239 nm. The concentration range over which the method was applied was 0.5–10.0 μg·l−1 of BPA with relative standard deviations of 2.9% for a concentration of 4.0 μg·l−1 of BPA. The detection limit was 0.07 μg·l−1. The proposed method was applied satisfactorily to the determination of BPA in synthetic mixtures and water samples from different sources previously spiked with different amounts of these chemicals. Recovery values ranging from 93% to 112% were obtained for water samples.  相似文献   

15.
不同结构破乳剂油水界面扩张粘弹性研究   总被引:13,自引:1,他引:13  
研究了支链破乳剂AE121和直链破乳剂SP169在正癸烷-水界面上的扩张粘弹性质,阐述了两种破乳剂扩张模量随扩张频率和破乳剂浓度的变化规律,考察了两种破乳剂对原油活性组分界面扩张性质的影响,测定了两种破乳剂的水溶液与正癸烷的动态界面张力,并与界面扩张流变性质进行了关联.研究结果表明,两种破乳剂的加入均会大大降低原油活性组分界面膜的扩张模量.较低浓度下直链破乳剂SP169由于吸附能力稍强,降低扩张模量效果较好;而一定浓度以上支链破乳剂AE121由于顶替能力较强,具有一定优势.由于破乳剂本身具有一定的扩张模量,在降低界面扩张模量的效果上,破乳剂的用量并非越大越好.  相似文献   

16.
The mechanism of anesthetic action on membranes is still an open question, regardless of their extensive use in medical practice. It has been proposed that anesthetics may have the effect of promoting pore formation across membranes or at least switching transmembrane channels. In both cases this may be the result of changes in the interfacial curvature of the membrane due to the presence of anesthetic molecules. Aqueous solutions of surfactants display phases that mimic, in a simplified manner, real biological membranes. Therefore, in this study, two nonionic surfactant systems C16E6/H2O in concentrated solution and C10E3/H2O in dilute solution have been used as model membranes for the investigation of the effects of six common anesthetics (halothane, sodium thiopental, lidocaine base form and hydrochloride, prilocaine hydrochloride, and ketamine hydrochloride). Both binary surfactant-water systems exhibit phase transitions from the lamellar phase, Lalpha, that has zero spontaneous curvature and zero monolayer curvature to phases with more local interfacial curvature. These are the random mesh phase, Mh1(0), which consists of lamellae pierced by water-filled pores with local areas of positive interfacial curvature and the sponge phase, L3, that consists of the lamellar phase with interlamellae attachments, often referred to as a "melted" cubic phase, possessing negative monolayer curvature. Small-angle X-ray scattering and 2H NMR experiments upon the C16E6/2H2O system and optical observations of the C10E3/H2O system showed that all anesthetics employed in this study cause a shift in the Mh1(0) to Lalpha phase transition temperature and in the Lalpha to L3 transition temperature, respectively. All of the anesthetics studied bind to the interfacial region of the surfactant systems. Two types of behavior were observed on anesthetic addition: type I anesthetics, which decreased interfacial curvature, and type II, which increased it. However, at physiological pH both types of anesthetics decreased interfacial curvature.  相似文献   

17.
The interfacial tension of the binary two-phase mixture of water and ethylene glycol isobutyl ether (EIB) was measured as a function of temperature in the vicinity of the lower critical solution temperatureT c under atmospheric pressure. The interfacial tension decreased with decreasing temperature and approached zero atT c . The thermodynamic quantities of interface formation were evaluated by applying equations developed previously to the interfacial tension vs temperature curves. The results were compared with those of the water and diethylene glycol diethyl ether system examined previously, and the effect of the molecular structure of the ether molecule on its interfacial behavior was discussed. It was suggested that the hydration of the ethylene oxide groups of ether molecules was an important factor in the interface formation as well as in the mixing of component molecules of the systems investigated here.  相似文献   

18.
Studies of the reaction of tin dichloride and of metallic tin with hydrochloric acid gas in diethyl ether are described. On the basis of spectroscopic and analytical data it is concluded that the product is not to be regarded as a tin-hydrogen bonded analog of chloroform, viz. trichlorostannane(IV), and the formation of an equilibrium mixture of hydrogen trichlorostannate(II) and dihydrogen tetrachlorostannate(II) is tentatively suggested. The complexes slowly decompose as a result of diethyl ether cleavage, yielding ethanol and ethyl chloride, together with traces of ethyltin trichloride.The mechanism of addition of the complexes to methyl acrylate is discussed in terms of a 1,4-addition involving initial protonation of oxygen. Reaction of the resulting β-carbomethoxyethyltin trichloride with tin or with zinc gives rise to mixtures of bis(β-carbomethoxyethyl)tin dichloride and tris(β-carbomethoxyethyl)tin chloride Similar reactions were observed for the first time with methyltin trichloride, the reaction with zinc being extremely fast at room temperature.  相似文献   

19.
The reaction of trimethylaluminum and dry, high-surface-area (500 m2/g) silica gel in a mixed slurry was studied using multinuclear, solid-state NMR spectroscopy. The products of the initial reaction were characterized, and their progress through subsequent washing with diethyl ether and reactions with measured amounts of water was followed. The quantitative distribution of different chemical forms of carbon deposited on the silica surface by the initial reaction was measured. The products of the initial reaction are dominated by methyl species of the types Al(CH3)n (with Si-O-Al linkages), Si-O-CH3, and (Si-O)4-nSi(CH3)n; aluminum is seen to exist predominantly as a five-coordinate species. Subsequent treatment with diethyl ether fails to remove any surface species, but instead the ether becomes strongly associated with the surface and highly resistant to removal. Stepwise additions of water hydrolyze the Al-CH3 and Si-O-CH3 moieties, leading to conversion of five-coordinate aluminum to four- and six-coordinate aluminum, and affect the partial release of the surface-associated diethyl ether; Si-CH3 moieties remain. The effect of aromatic and saturated solvents on the initial reaction was examined and found to cause a small but significant change in the distribution of products. Structures of aluminum-centered species on the silica surface consistent with the spectroscopic data are proposed.  相似文献   

20.
Reactions of 2-benzylidenemalononitrile and 2-nitro-3-phenylacrylonitrile with aryl azides in diethyl ether at room temperature gave mixtures of regioisomeric 1(3)-aryl-5-phenyl-4,5-dihydro-1(3)H-1,2,3-triazole-4,4-dicarbonitriles and 1-aryl-5(4)-phenyl-1H-1,2,3-triazole-5(4)-carbonitriles, respectively. 2-Benzylidenemalononitrile reacted with the same arylazides on heating in boiling chloroform to produce 1-aryl-2-phenylaziridine-2,2-dicarbonitriles.  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