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.
《Small Ruminant Research》2002,43(2):167-178
This study compares cardiopulmonary, hematological, serum biochemical and behavioral effects of sevoflurane, isoflurane or halothane anesthesia in spontaneously breathing, conventionally medicated goats. Six male adult goats were anesthetized repeatedly at 2-week intervals with three anesthetics. Goats were administered atropine (0.1 mg/kg) intramuscularly, and 10 min later, induced to anesthesia by an intravenous infusion of thiopental (mean 14.3 mg/kg). After intubation, goats were anesthetized with halothane, isoflurane or sevoflurane in oxygen and maintained at surgical depth of anesthesia for 3 h. Recovery from anesthesia with sevoflurane was more rapid than that with isoflurane or halothane. Time-related hypercapnia and acidosis were observed during halothane anesthesia, but not observed during sevoflurane or isoflurane anesthesia. Both hypercapnia and acidosis during sevoflurane anesthesia did not differ from isoflurane anesthesia, but were less during halothane anesthesia, especially at prolonged maintenance period. There were no significant differences between anesthetics in respiration and heart rates, arterial pressures, hematological and serum biochemical values. It was concluded that sevoflurane is an effective inhalant for use in goats showing the most rapid recovery from anesthesia, and that cardiopulmonary effects of sevoflurane are similar to isoflurane than halothane.  相似文献   

2.
Partitioning of four modern inhalational anesthetics (halothane, isoflurane, enflurane, and sevoflurane) between the gas phase and nine organic solvents that model different amino acid side-chains and lipid membrane domains was performed in an effort to define which microenvironments present in proteins and lipid bilayers might be favored. Compared to a purely aliphatic environment (hexane), the presence of an aromatic-, alcohol-, thiol- or sulfide group on the solvent improved anesthetic partitioning, by factors of 1.3-5.2 for halothane, 1.7-5.6 for isoflurane, 1.7-7.6 for enflurane, and 1.5-7.3 for sevoflurane. The most favorable solvent for halothane partitioning was ethyl methyl sulfide, a model for methionine. Enflurane and isoflurane partitioned most extensively into methanol, a model for serine, and sevoflurane into ethanol, a model for threonine. Isoflurane also partitioned favorably into ethyl methyl sulfide. The results suggest that volatile general anesthetics interact better with partly polar groups, which are present on amino acids frequently found buried in the hydrophobic core of proteins, compared to purely aliphatic side-chains. Furthermore, if an anesthetic molecule was located in a saturated region of a phospholipid bilayer membrane, there would be an energetically favorable driving force for it to move into several higher dielectric microenvironments present on membrane proteins. The results provide evidence that proteins rather than lipids are the likely targets of volatile general anesthetics in biological membranes.  相似文献   

3.
Molecular genetic analysis of volatile-anesthetic action.   总被引:7,自引:2,他引:5       下载免费PDF全文
The mechanism(s) and site(s) of action of volatile inhaled anesthetics are unknown in spite of the clinical use of these agents for more than 150 years. In the present study, the model eukaryote Saccharomyces cerevisiae was used to investigate the action of anesthetic agents because of its powerful molecular genetics. It was found that growth of yeast cells is inhibited by the five common volatile anesthetics tested (isoflurane, halothane, enflurane, sevoflurane, and methoxyflurane). Growth inhibition by the agents is relatively rapid and reversible. The potency of these compounds as yeast growth inhibitors directly correlates with their lipophilicity as is predicted by the Meyer-Overton relationship, which directly correlates anesthetic potency of agents and their lipophilicity. The effects of isoflurane on yeast cells were characterized in the most detail. Yeast cells survive at least 48 h in a concentration of isoflurane that inhibits colony formation. Mutants resistant to the growth-inhibitory effects of isoflurane are readily selected. The gene identified by one of these mutations, zzz4-1, has been cloned and characterized. The predicted ZZZ4 gene product has extensive homology to phospholipase A2-activating protein, a GO effector protein of mice. Both zzz4-1 and a deletion of ZZZ4 confer resistance to all five of the agents tested, suggesting that signal transduction may be involved in the response of these cells to volatile anesthetics.  相似文献   

4.
Biophysical studies of protein–anesthetic interactions using nuclear magnetic resonance (NMR) spectroscopy are often conducted by the addition of micro amounts of neat inhaled anesthetic which yields much higher than clinically relevant (0.2–0.5 mM) anesthetic concentrations. We report a 19F NMR technique to measure clinically relevant inhaled anesthetic concentrations from saturated aqueous solutions of these anesthetics (halothane, isoflurane, sevoflurane, and desflurane). We use a setup with a 3-mm NMR tube (containing trifluoroacetic acid as standard), coaxially inserted in a 5-mm NMR tube containing anesthetic solution under investigation. All experiments are conducted in a 5-mm NMR probe. We also have provided standard curves for four inhaled anesthetics using NMR technique. The standard curve for each of these anesthetics is helpful in determining the prerequisite amount of aqueous anesthetic solution required to prepare clinically relevant concentrations for protein–anesthetic interaction studies. Parts of the results to be presented at Society for Neuroscience meeting, 2008.  相似文献   

5.
Sawas AH  Pentyala SN  Rebecchi MJ 《Biochemistry》2004,43(39):12675-12685
This study directly examines the enthalpic contributions to binding in aqueous solution of closely related anesthetic haloethers (desflurane, isoflurane, enflurane, and sevoflurane), a haloalkane (halothane), and an intravenous anesthetic (propofol) to bovine and human serum albumin (BSA and HSA) using isothermal titration calorimetry. Binding to serum albumin is exothermic, yielding enthalpies (DeltaH(obs)) of -3 to -6 kcal/mol for BSA with a rank order of apparent equilibrium association constants (K(a) values): desflurane > isoflurane approximately enflurane > halothane >or= sevoflurane, with the differences being largely ascribed to entropic contributions. Competition experiments indicate that volatile anesthetics, at low concentrations, share the same sites in albumin previously identified in crystallographic and photo-cross-linking studies. The magnitude of the observed DeltaH increased linearly with increased reaction temperature, reflecting negative changes in heat capacities (DeltaC(p)). These -DeltaC(p) values significantly exceed those calculated for burial of each anesthetic in a hydrophobic pocket. The enhanced stabilities of the albumin/anesthetic complexes and -DeltaC(p) are consistent with favorable solvent rearrangements that promote binding. This idea is supported by substitution of D(2)O for H(2)O that significantly reduces the favorable binding enthalpy observed for desflurane and isoflurane, with an opposing increase of DeltaS(obs). From these results, we infer that solvent restructuring, resulting from release of water weakly bound to anesthetic and anesthetic-binding sites, is a dominant and favorable contributor to the enthalpy and entropy of binding to proteins.  相似文献   

6.
7.
Exposure of rats to the volatile anesthetics, halothane, enflurane and isoflurane and low FIO2 (0.8%) for two hours results in a transient induction of ODC appearing maximally four hours after exposure. Without the low oxygen accompanying the anesthetic or the low oxygen alone, no significant induction of ODC occurred. The concentration of anesthetic used to produce the ODC induction were 0.5% halothane, 1.5% enflurane and 1.4% isoflurane. Except for halothane, reducing the anesthetic concentration only slightly reduced the effect on ODC levels to control values. Reduction of halothane concentrations to 0.1% was required to reduce the values to control levels. Pretreatment of the animals with either cycloheximide or actinomycin D delayed the onset of ODC induction. The data support the fact that liver damage can occur in the absence of metabolism of the drug.  相似文献   

8.
The halogenated anesthetics halothane, enflurane and isoflurane inhibit the calcium efflux induced by Ruthenium Red in isolated rat liver mitochondria. The extent of the inhibition is higher for enflurane (approximately 50%) than for either isoflurane (approximately 35%) or halothane (approximately 15%), and does not increase significantly between 0.1 and 0.6-1.0 mM anesthetic. Both the mitochondrial respiratory rate and transmembrane electrical potential are unaffected by the halogenated anesthetics concentrations capable to inhibit the efflux of calcium.  相似文献   

9.
The halogenated hydrocarbons, such as halothane, are widely used as anesthetics in clinical practice; however their application is often accompanied with metabolic, cardiovascular and respiratory complications. One of the possible factors for this negative outcome might be the severe toxicity of these agents. In this paper, we investigate in vitro effects of halothane on human lung carcinoma A 549 cells, namely on their cytotoxicity, adhesive properties and metabolic activity. The cytotoxicity response of lung carcinoma A 549 cells to halothane was determined by lactate dehydrogenase (LDH) assay (for cytotoxicity), by detachment assay after adhesion to type IV collagen (for cell adhesive properties) and by surface tension measurements of culture medium (for cell metabolic activity). Regarding the cytotoxicity, the determined maximal non-toxic concentration of halothane on A 549 cells, given here as volume percentages (vol.%) was 0.7 vol.% expressed as aqueous concentration in the culture medium. Direct measurement of the actual halothane concentration in the culture medium showed that 0.7 vol.% corresponds to 1.05 mM and 5.25 aqueous-phase minimum alveolar concentration (MAC). Concentrations equal or higher than 1.4 vol.% (2.1 mM; 10.5 MAC) of halothane provoked complete detachment (cell death), or reduction of initial adhesion to collagen IV in half of the cell population. Surfactant production of A 549 cells, registered up to 48 h after halothane treatment, was inhibited by halothane concentrations as low as 0.6 vol.% (0.9 mM; 4.5 MAC). Our results demonstrate that sub toxic halothane concentrations of 0.6 vol.% inhibits surfactant production; concentrations in the range 0.8-1.4 vol.% induce membrane damages and concentrations equal and higher than 1.4 vol.%--cell death of approximately 50% of the cells.  相似文献   

10.
Anesthetic agents have well-defined pharmacological targets but their effects on energy metabolism in the brain are poorly understood. In this study, we examined the effects of different anesthetics on extracellular lactate and glucose levels in blood, CSF and brain of the mouse. In vivo-microdialysis was used to monitor extracellular energy metabolites in the brain of awake mice and during anesthesia with seven different anesthetic drugs. In separate groups, lactate and glucose concentrations in blood and CSF were measured for each anesthetic. We found that anesthesia with isoflurane caused a large increase of extracellular lactate levels in mouse striatum and hippocampus (300-400%). Pyruvate levels also increased while glucose and glutamate levels were unchanged. This effect was dose-dependent and was mimicked by other gaseous anesthetics such as halothane and sevoflurane but not by intravenous anesthetics. Ketamine/xylazine and chloral hydrate caused 2-fold increases of glucose levels in mouse blood and brain while lactate levels were only moderately increased. Propofol caused a minor increase of extracellular glucose levels while pentobarbital had no effect on either lactate or glucose. Volatile anesthetics also increased lactate levels in blood and CSF by 2-3-fold but had no effect on plasma glucose. Further experiments demonstrated that lactate formation by isoflurane in mouse brain was independent of neuronal impulse flow and did not involve ATP-dependent potassium channels. We conclude that volatile anesthetics, but not intravenous anesthetics, cause a specific, dose-dependent increase in extracellular lactate levels in mouse brain. This effect occurs in the absence of ischemia, is independent of peripheral actions and is reflected in strongly increased CSF lactate levels.  相似文献   

11.
Anesthetic agents have well-defined pharmacological targets but their effects on energy metabolism in the brain are poorly understood. In this study, we examined the effects of different anesthetics on extracellular lactate and glucose levels in blood, CSF and brain of the mouse. In vivo-microdialysis was used to monitor extracellular energy metabolites in the brain of awake mice and during anesthesia with seven different anesthetic drugs. In separate groups, lactate and glucose concentrations in blood and CSF were measured for each anesthetic. We found that anesthesia with isoflurane caused a large increase of extracellular lactate levels in mouse striatum and hippocampus (300–400%). Pyruvate levels also increased while glucose and glutamate levels were unchanged. This effect was dose-dependent and was mimicked by other gaseous anesthetics such as halothane and sevoflurane but not by intravenous anesthetics. Ketamine/xylazine and chloral hydrate caused 2-fold increases of glucose levels in mouse blood and brain while lactate levels were only moderately increased. Propofol caused a minor increase of extracellular glucose levels while pentobarbital had no effect on either lactate or glucose. Volatile anesthetics also increased lactate levels in blood and CSF by 2–3-fold but had no effect on plasma glucose. Further experiments demonstrated that lactate formation by isoflurane in mouse brain was independent of neuronal impulse flow and did not involve ATP-dependent potassium channels. We conclude that volatile anesthetics, but not intravenous anesthetics, cause a specific, dose-dependent increase in extracellular lactate levels in mouse brain. This effect occurs in the absence of ischemia, is independent of peripheral actions and is reflected in strongly increased CSF lactate levels.  相似文献   

12.
To assess the effects of halothane, isoflurane, and sevoflurane on cross bridges in intact cardiac muscle, electrically stimulated (0.25 Hz, 25 degrees C) right ventricular ferret papillary muscles (n = 14) were subjected to sinusoidal load oscillations (37-182 Hz, 0.2-0.5 mN peak to peak) at the instantaneous self-resonant frequency of the muscle-lever system. At resonance, stiffness is proportional to m * omega(2) (where m is equivalent moving mass and omega is angular frequency). Dynamic stiffness was derived by relating total stiffness to values of passive stiffness at each length during shortening and lengthening. Shortening amplitude and dynamic stiffness were decreased by halothane > isoflurane > or = sevoflurane. At equal peak shortening, dynamic stiffness was higher in halothane or isoflurane in high extracellular Ca(2+) concentration than in control. Halothane and isoflurane increased passive stiffness. The decrease in dynamic stiffness and shortening results in part from direct effects of volatile anesthetics at the level of cross bridges. The increase in passive stiffness caused by halothane and isoflurane may reflect an effect on weakly bound cross bridges and/or an effect on passive elastic elements.  相似文献   

13.
The effect of three volatile anesthetics (halothane, enflurane, and isoflurane) on coronary flow and metabolic state of isolated rat hearts was studied. These anesthetics are coronary dilators and their effects are dose dependent. At 2 MAC (minimum alveolar concentration), isoflurane, enflurane, and halothane increase coronary flow by 114 +/- 5.9, 93 +/- 6.1, and 77 +/- 6.4%, respectively (p less than 0.001). At these concentrations, they also have a modest but significant metabolic effect causing a 30% reduction in myocardial ATP and phosphocreatine levels, with no significant modification in ADP and AMP concentrations. Energy charge and lactate/pyruvate ratio were also unaffected by these anesthetics. The vascular and metabolic effects were reversible within 2 and 30 min, respectively. Perfusion of the hearts with a Krebs-Henseleit solution without Pi did not interfere with the vascular and the metabolic effect of the anesthetics; however, in this case, ATP and phosphocreatine concentration did not return to control levels after their discontinuation despite full recovery of the vascular effect. These data suggest that the volatile anesthetics have direct coronary vascular and myocardial metabolic effects and that these effects occur independently.  相似文献   

14.
The effects of halothane, isoflurane, and sevoflurane on elastic stiffness, which reflects the degree of cross-bridge attachment, were studied in intact cardiac muscle. Electrically stimulated (0.25 Hz, 25 degrees C), isometrically twitching right ventricular ferret papillary muscles (n = 15) at optimal length (L(max)) were subjected to sinusoidal length oscillations (40 Hz, 0.25- 0.50% of L(max) peak to peak). The amplitude and phase relationship with the resulting force oscillations was decomposed into elastic and viscous components of total stiffness in real time. Increasing extracellular Ca(2+) concentration in the presence of anesthetics to produce peak force equal to control increased elastic stiffness during relaxation, which suggests a direct effect of halothane and sevoflurane on cross bridges.  相似文献   

15.
We assessed the possible effects of the volatile halogenated anesthetics halothane, enflurane, and isoflurane on Ca(2+) electrode measurements and on the Ca(2+) sensitivity of the bioluminescent protein aequorin. In Ca(2+)-EGTA buffers of different pCa values (7. 870, 6.726, 6.033, 4.974, 4.038, and 2.995) and in serial Ca(2+) dilutions (10(-4), 10(-3), and 10(-2) M), halothane, enflurane, and isoflurane each caused a concentration-dependent and reversible increase in the absolute value of the negative electrode potential. Isoflurane and enflurane had larger effects than halothane. Neither of these anesthetics changed aequorin luminescence at any pCa tested in the range 2-8. There was no potentiation or inactivation of aequorin luminescence over a period of up to 2 h. These results suggest that (1) halothane, enflurane, and isoflurane interfere with Ca(2+) electrode measurements, most likely by changing the physicochemical properties of the membrane; (2) these anesthetics do not inactivate or otherwise modify the characteristics of the reaction of Ca(2+) with aequorin; and (3) these anesthetics do not change the apparent affinity of EGTA for Ca(2+).  相似文献   

16.
Anesthetics are in routine use, yet the mechanisms underlying their function are incompletely understood. Studies in vitro demonstrate that both GABA(A) and NMDA receptors are modulated by anesthetics, but whole animal models have not supported the role of these receptors as sole effectors of general anesthesia. Findings in C. elegans and in children reveal that defects in mitochondrial complex I can cause hypersensitivity to volatile anesthetics. Here, we tested a knockout (KO) mouse with reduced complex I function due to inactivation of the Ndufs4 gene, which encodes one of the subunits of complex I. We tested these KO mice with two volatile and two non-volatile anesthetics. KO and wild-type (WT) mice were anesthetized with isoflurane, halothane, propofol or ketamine at post-natal (PN) days 23 to 27, and tested for loss of response to tail clamp (isoflurane and halothane) or loss of righting reflex (propofol and ketamine). KO mice were 2.5 - to 3-fold more sensitive to isoflurane and halothane than WT mice. KO mice were 2-fold more sensitive to propofol but resistant to ketamine. These changes in anesthetic sensitivity are the largest recorded in a mammal.  相似文献   

17.
BACKGROUND: Desflurane and enflurane have been reported to produce substantial amounts of carbon monoxide (CO) in desiccated sodalime. Isoflurane is said to produce less CO and sevoflurane and halothane should produce no CO at all.The purpose of this study is to measure the maximum amounts of CO production for all modern volatile anesthetics, with completely dry sodalime. We also tried to establish a relationship between CO production and temperature increase inside the sodalime. METHODS: A patient model was simulated using a circle anesthesia system connected to an artificial lung. Completely desiccated sodalime (950 grams) was used in this system. A low flow anesthesia (500 ml/min) was maintained using nitrous oxide with desflurane, enflurane, isoflurane, halothane or sevoflurane. For immediate quantification of CO production a portable gas chromatograph was used. Temperature was measured within the sodalime container. RESULTS: Peak concentrations of CO were very high with desflurane and enflurane (14262 and 10654 ppm respectively). It was lower with isoflurane (2512 ppm). We also measured small concentrations of CO for sevoflurane and halothane. No significant temperature increases were detected with high CO productions. CONCLUSION: All modern volatile anesthetics produce CO in desiccated sodalime. Sodalime temperature increase is a poor predictor of CO production.  相似文献   

18.
Solt K  Johansson JS  Raines DE 《Biochemistry》2006,45(5):1435-1441
Inhaled anesthetics are thought to alter the conformational states of Cys-loop ligand-gated ion channels (LGICs) by binding within discrete cavities that are lined by portions of four alpha-helical transmembrane domains. Because Cys-loop LGICs are complex molecules that are notoriously difficult to express and purify, scaled-down models have been used to better understand the basic molecular mechanisms of anesthetic action. In this study, stopped-flow fluorescence spectroscopy was used to define the kinetics with which inhaled anesthetics interact with (Aalpha(2)-L1M/L38M)(2), a four-alpha-helix bundle protein that was designed to model anesthetic binding sites on Cys-loop LGICs. Stopped-flow fluorescence traces obtained upon mixing (Aalpha(2)-L1M/L38M)(2) with halothane revealed immediate, fast, and slow components of quenching. The immediate component, which occurred within the mixing time of the spectrofluorimeter, was attributed to direct quenching of tryptophan fluorescence upon halothane binding to (Aalpha(2)-L1M/L38M)(2). This was followed by a biexponential fluorescence decay containing fast and slow components, reflecting anesthetic-induced conformational transitions. Fluorescence traces obtained in studies using sevoflurane, isoflurane, and desflurane, which poorly quench tryptophan fluorescence, did not contain the immediate component. However, these anesthetics did produce the fast and slow components, indicating that they also alter the conformation of (Aalpha(2)-L1M/L38M)(2). Cyclopropane, an anesthetic that acts with unusually low potency on Cys-loop LGICs, acted with low apparent potency on (Aalpha(2)-L1M/L38M)(2). These results suggest that four-alpha-helix bundle proteins may be useful models of in vivo sites of action that allow the use of a wide range of techniques to better understand how anesthetic binding leads to changes in protein structure and function.  相似文献   

19.
Ca2+-activated K+ channels in rat glioma C6 cells were investigated using monolayers of these cells in petri dishes. The ion flux through the channels was studied with 86Rb+ after addition of a Ca2+-ionophore to the incubation medium. Both the influx and efflux of 86Rb+ through these Ca2+-activated K+ channels were inhibited by the general anesthetic halothane (at clinical concentrations). Other volatile anesthetics such as isoflurane, enflurane and methoxyflurane also inhibited the Ca2+-activated K+ channels at clinical concentrations. Inhibition of these channels by general anesthetics could have profound effects on signal transmission in the brain.  相似文献   

20.
PC12 cells preloaded with [3H]norepinephrine release this neurotransmitter at a slow rate (basal release). This rate is increased by the addition of phorbol myristate acetate (PMA), but not by a biologically inactive phorbol ester. This effect most likely is mediated by protein kinase C, since desensitization of this kinase abolished the stimulation of the neurotransmitter release by PMA. Unexpectedly, clinical concentrations of the volatile anesthetics halothane, enflurane, isoflurane and methoxyflurane stimulated the PMA evoked neurotransmitter release in good correlation with their anesthetic potency. Since the volatile anesthetics increased the cytoplasmic Ca2+ concentration of the PC12 cells in a dose dependent manner it seems very likely that the effect of the anesthetics on the PMA-evoked neurotransmitter release is mediated by this rise in Ca2+ concentration.  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