Hydration of lipid membranes and the action mechanisms of anesthetics and alcohols

https://doi.org/10.1016/S0009-3084(99)00056-0Get rights and content

Section snippets

Nerve excitation generates heat: thermotropic transition

Nerve excitation is a transition from high temperature (resting) state to low temperature (excited) state of excitable membranes. Nerve excitation generates heat (Abbott et al., 1958, Tasaki and Iwasa, 1981, Ritchie and Keynes, 1985, Tasaki and Byrne, 1987). The heat-flow profile can be superimposed on the action potential on a millisecond scale. It indicates that at least two thermodynamically identifiable states exist for nerve excitation. Because excitation (depolarization) is exothermic and

Are lipid membranes irrelevant to anesthesia mechanisms?

We envision anesthesia as a non-specific cooperative action of hydrophobic molecules on both lipids and proteins. There are no specific receptor sites for anesthetics. The only non-specific drugs similar to anesthetics are disinfectants. Anesthetics and alcohols occupy unusual position in pharmacology. Pharmacology is a discipline that studies the specific effect of a specific drug on a specific receptor on a specific protein. Epinephrine acts only on epinephrine receptors, and does not affect

Temperature and membrane fluidity

Due to the lack of rigorous definition, membrane fluidity invited a number of controversies. The membrane fluidity, induced by anesthetics, can be achieved by temperature elevation of less than 1.0°C. Because an increase of temperature does not induce anesthesia, it is claimed that membrane fluidity is irrelevant to anesthesia mechanisms. Biological activities increase with the elevation of temperature reaching a maximum (optimum temperature) and then decrease. Fluidity increases linearly with

The enthalpic theory of anesthesia

Contrary to the proposed theory, increased anesthetic concentrations at the interaction site at lower temperature (negative ΔH) does not mean an increase of the anesthetic activity at lower temperature. Concentration and activity are not identical. The anesthetic interactions with lipid membranes are usually associated with positive enthalpies (Katz and Diamond, 1974c), whereas negative enthalpies are reported on their interaction with proteins, e.g. bovine serum albumin (Ueda and Yamanaka, 1997

Anesthetic concentration and its activity

Anesthetic potency is expressed by the thermodynamic activity, which is the ratio [anesthetic concentration]/[saturating concentration of the anesthetic] in the reference medium. It is important to distinguish molecular concentration and its activity. Ferguson (1939) proposed ‘thermodynamic activity’ for quantitative analysis of anesthetic potency. The thermodynamic activity is expressed by the anesthetizing gas concentration (minimum alveolar concentration, MAC) divided by the saturated

Anesthetic–water competition: solvent effect

Homopolymers in water change the size according to the solvent property of water. Lipid molecules form bilayer structures only when water is present. Due to its extensive hydrogen bonding capability, water is repulsive to hydrophobic molecules and is a poor solvent. In contrast, water is attractive to hydrophilic molecules and is a good solvent. When amphipathic molecules are dispersed in water, the hydrophobic parts are squeezed out of water, and bilayers and micelles are formed by the balance

Unfreezable bound water and anesthetics: DSC

Anesthetics decrease the subzero freezing peaks of bound water in partially hydrated multilamellar DMPC assemblies. Recent studies on water structure are mainly consisted of computer simulated dynamics of water molecules. Classic theories of water structure, however, provided conceptual models. The significant structure theory (Eyring and Jhon, 1969) assumes two solid-like species and a rotating monomer. In this model, the liquid water is visualized as an equilibrium mixture of a water monomer

Anesthetics increase the transition temperature of multilamellar vesicles with low water content

The thermal transition temperature started to increase when the water content was less than 18 wt.%. Under this condition, addition of anesthetics further increased the transition temperature. This result is a complete reverse of the anesthetic effect on fully hydrated lipid membranes, where anesthetics and alcohols decreased the thermal transition temperature. This result shows that the stability of lipid membranes is a function of water–lipid interaction, and the primary site of action of

Water binding site in reversed micelles: PO, (CH3)3N, and CO stretching

The main water binding site is PO4 moiety. The effect of anesthetics on the water–lipid (DMPC) interaction was studied in a water-in-oil (benzene) reversed micelle (Tsai et al., 1987). The study included clinically used polar agents (halothane, chloroform and enflurane) and apolar carbon tetrachloride CCl4. The OH stretching frequency, representing water, increased from 3369 to a steady 3430 cm−1 when the water:DMPC mole ratio exceeded 18. Nevertheless, the value did not quite reach the

Competition between anesthetics and water: NMR

In reversed micellar solutions, anesthetics released the surface-bound water. NMR results confirm the FTIR data. Yoshida et al. (1984) used water-in-oil reversed micellar system and showed by 1H NMR that anesthetics decreased the structured interfacial water at the membrane surface. The spin-spin relaxation time, T2, of 1H NMR increased when enflurane was added to water-in-oil reversed micelles, showing an increase in the mobility of bound water molecules.

The spin-lattice relaxation time, T1,

Anesthetic binding mode to lipid membranes: langmuir adsorption isotherms

Saturability of anesthetic interaction with protein does not indicate that the anesthetic action site is limited to proteins. The Langmuir adsorption on lipid membranes also saturates. Saturability of anesthetic binding to macromolecules has been an issue in anesthesia mechanisms. When the binding is saturable, presence of specific receptors for anesthetics on proteins is generally implied. With NiCl2-stabilized giant-size (200 mm2) oleylamine (cis-9-octadecanoylamine) black-lipid membrane

1H and 19F NMR chemical shift

NMR showed that anesthetics stayed at the lipid–water interface. From the 1H and 19F NMR chemical shift changes, Yoshida et al. (1989c) established that anesthetics stayed at the lipid–water interfacial area. Anesthetics molecules did not lose contact with bulk water 2H2O.

The depth of halothane penetration into sodium dodecylsulfate micelles was estimated by paramagnetic relaxation of 19F NMR spin-lattice relaxation (Yoshino et al., 1994). The presence of paramagnetic ions increases the

Lipid-coated pore

Anesthetics inhibit ionic flow parallel to the membrane pore in a phospholipid-impregnated pore of polycarbonate filters. Yoshida et al. (1989b) used 10 μm polycarbonate filters with 0.1 μm diameter pore, impregnated with DMPC or DPPC, and measured the effects of temperature and an inhalation anesthetic, halothane, on the surface capacitance and conductance by the impedance dispersion at a frequency range 30 Hz∼1.0 MHz. When the frequency was below 120 Hz, the capacitance showed a peak at the

Anesthetics and transition of lipid membranes

Anesthetics bind to the solid-gel states as well as liquid-crystal states. By analogy to freezing point depression, anesthetic binding to lipid membranes is estimated from the dependence of main transition temperature to anesthetic concentrations (Kaminoh et al., 1988a). However, anesthetics also bind to solid-gel membranes to a significant degree. With radioactive halothane, Simon et al. (1979) reported that the solid-gel membrane binding is about 30% of the liquid-crystal membrane binding.

Anesthesia cut-off

In lipid membranes, the sudden loss of anesthetic potency when the alcohol carbon chain exceeds a certain length (cut-off) is associated with the switch of their affinity from the liquid-crystal to the solid-gel state. In a homologous series of n-alcohols, elongation of the carbon chain is accompanied by an increase in anesthetic potency in parallel with an increase in the oil/water partition coefficients. The anesthetic potency, however, suddenly disappears when the carbon chain exceeds a

First page preview

First page preview
Click to open first page preview

References (70)

  • A. Shibata et al.

    Anesthetic–protein interaction: effects of volatile anesthetics on the secondary structure of poly(l-lysine)

    J. Pharm. Sci.

    (1991)
  • Y. Suezaki et al.

    Statistical mechanics of pressure–anesthetic antagonism on the phase transition of phospholipid membranes: Interfacial water hypothesis

    J. Colloid Interface Sci.

    (1983)
  • Y. Suezaki et al.

    Molecular origin of biphasic response of main phase-transition temperature of phospholipid membranes to long-chain alcohols

    Biochim. Biophys. Acta

    (1985)
  • K. Tamura et al.

    High pressure antagonism of alcohol effects on the main phase-transition temperature of phospholipid membranes: biphasic response

    Biochim. Biophys. Acta

    (1991)
  • I. Tasaki et al.

    Temperature changes associated with nerve excitation: detection by using polyvinylidene fluoride film

    Biochem. Biophys. Res. Commun.

    (1981)
  • I. Tasaki et al.

    Heat production associated with synaptic transmission in the bullfrog spinal cord

    Brain Res.

    (1987)
  • Y.S. Tsai et al.

    Infrared spectra of phospholipid membranes: Interfacial dehydration by volatile anesthetics and phase transition

    Biochim. Biophys. Acta

    (1990)
  • I. Ueda et al.

    Titration calorimetry of anesthetic–protein interaction: Negative enthalpy of binding and anesthetic potency

    Biophys. J.

    (1997)
  • I. Ueda et al.

    Irreversible phase transition of firefly luciferase: contrasting effects of volatile anesthetics and myristic acid

    Biochim. Biophys. Acta

    (1998)
  • I. Ueda et al.

    Is there a specific receptor for anesthetics? Contrary effects of alcohols and fatty acids on phase transition and bioluminescence of firefly luciferase

    Biophys. J.

    (1998)
  • S. Yokono et al.

    Interfacial preference of anesthetic action upon the phase transition of phospholipid bilayers and partition equilibrium of inhalation anesthetics between membrane and deuterium oxide

    Biochim. Biophys. Acta

    (1981)
  • S. Yokono et al.

    400 MHz two-dimensional nuclear Overhauser spectroscopy on anesthetic interaction with lipid bilayer

    Biochim. Biophys. Acta

    (1989)
  • T. Yoshida et al.

    Giant planar lipid bilayer I. Capacitance and its biphasic response to inhalation anesthetics

    J. Colloid Interface Sci.

    (1983)
  • T. Yoshida et al.

    Giant planar lipid bilayer II. Conductance and interfacial effect of inhalation anesthetics

    J. Colloid Interface Sci.

    (1983)
  • T. Yoshida et al.

    A proton nuclear magnetic resonance study on the release of bound water by inhalation anesthetic in water-in-oil emulsion

    Biochim. Biophys. Acta

    (1984)
  • T. Yoshida et al.

    19F-NMR study on micellar solubilization of volatile anesthetic halothane: Dose-related biphasic interaction

    J. Colloid Interface Sci.

    (1988)
  • T. Yoshida et al.

    Saturable and unsaturable binding of a volatile anesthetic enflurane with model lipid vesicle membranes

    Biochim. Biophys. Acta

    (1989)
  • T. Yoshida et al.

    Changes in surface capacitance and conductance parallel to phospholipid membranes associated with phase transition: effects of halothane

    Biochim. Biophys. Acta

    (1989)
  • T. Yoshida et al.

    Molecular orientation of volatile anesthetics at the binding surface: 1H and 19F NMR studies of submolecular affinity

    Biochim. Biophys. Acta

    (1989)
  • T. Yoshida et al.

    Proton flow along lipid bilayer surfaces: effect of halothane on the lateral surface conductance and membrane hydration

    Biochim. Biophys. Acta

    (1990)
  • T. Yoshida et al.

    Negative entropy of halothane binding to protein: 19F-NMR with a novel cell

    Biochim. Biophys. Acta

    (1997)
  • A. Yoshino et al.

    Anesthetic effect on interfacial water: A 2H-NMR study of water-in-oil reversed micelles

    J. Colloid Interface Sci.

    (1989)
  • A. Yoshino et al.

    Surface-oriented saturable binding of halothane with micelles: paramagnetic relaxation of 19F-NMR spin-lattice relaxation rate and gas chromatography studies

    J. Colloid Interface Sci.

    (1994)
  • A. Yoshino et al.

    19F and 1H NMR and NOE study on halothane–micelle interaction: residence location of anesthetic molecules

    J. Colloid Interface Sci.

    (1998)
  • Y. Xu et al.

    Amphiphilic sites for general anesthetic action? Evidence from 129Xe–{1H} intermolecular nuclear Overhauser effects

    Biochim. Biophys. Acta

    (1997)
  • Cited by (47)

    • NIR studies of cholesterol-dependent structural modification of the model lipid bilayer doped with inhalation anesthetics

      2018, Journal of Molecular Structure
      Citation Excerpt :

      From the scores charts the temperature values of the chain-melting phase transitions of all of the studied membranes were determined. The presence of each anesthetic in the lipid bilayer reduced the value of Tm temperature at which an increase in content of gauche rotamers takes place [19,20,42,49,50]. On the basis of the values of loadings, the areas of the greatest spectral changes resulting from an increase in the participation of gauche conformers in the hydrocarbon chains of lipids were identified.

    • Spectral phasor analysis of LAURDAN fluorescence in live A549 lung cells to study the hydration and time evolution of intracellular lamellar body-like structures

      2016, Biochimica et Biophysica Acta - Biomembranes
      Citation Excerpt :

      Under these conditions lipid membranes may display changes in their supramolecular organization, e.g. transitions from lamellar to non-lamellar structures (a phenomenon called lyotropic mesomorphism) [49,50]. The number of water molecules per phospholipid required to stabilize a lamellar lipid structure is approximately 20, of which around 6–7 water molecules are highly coordinated with the lipids (i.e. cannot be crystalized above − 100 °C) [51]. Considering the low hydration state of intracellular LB-like structures in cellulo compared with the secreted ones and the differences observed with a lamellar model system (Fig. 5) we hypothesize the existence of membranous structures other than purely lamellar existing in intracellular these organelles under these conditions.

    View all citing articles on Scopus
    View full text