chapter 25 - inhaled anesthetics: mechanisms of action...inhaled anesthetics: mechanisms . of...

29
614 Chapter 25 Inhaled Anesthetics: Mechanisms of Action MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread clinical use of general anesthetics, our current understanding of their molecular, cellular, and network mechanisms is incomplete. This critical gap in the pharmacology of one of medicine’s most important drug classes not only impedes rational use of available anesthetics but also hinders the development of newer anesthetics that might selectively achieve the desirable end points of anesthesia with fewer adverse cardiovascu- lar, respiratory, and possibly neuropathologic side effects. Although major progress has been made in understand- ing the pharmacology of the intravenous anesthetics by molecular genetic approaches (see Chapter 30), the actions of the inhaled anesthetics at the molecular and cellular levels are more enigmatic. It is still not possible to trace precisely the sequence of events that leads from inhaled anesthetic–target interactions, through ascending levels of biologic complexity, to the various behavioral effects that characterize the composite state of clinical anesthesia in humans. Nevertheless, investigations con- tinue to reveal fundamental principles of action and have led to a framework for understanding anesthetic effects at different organizational levels. The focus of this chapter is on the mechanisms involved in the principal therapeutic effects (anesthesia) and on the side effects of the inhaled anesthetics (Fig. 25-1), a chemically and pharmacologi- cally diverse group that includes the potent halogenated K EY P OINTS Anesthesia consists of separable and independent components or physiologic substates, each of which involves distinct, but possibly overlapping, mechanisms at different sites in the central nervous system. The potency of general anesthetics correlates with their solubility in oil, indicating the importance of interactions with hydrophobic targets. General anesthetics act by binding directly to amphiphilic cavities in proteins. Binding sites can be identified by a combination of site-directed mutagenesis and high-resolution structural analysis of anesthetic binding. Mutations rendering putative effector proteins insensitive to inhaled anesthetics have been constructed and expressed in mice but have not generated breakthroughs analogous to the success of this strategy with intravenous anesthetics. The effects of inhaled anesthetics cannot be explained by a single molecular mechanism. Rather, multiple targets contribute to the component actions comprising the anesthetic effects of each anesthetic. However, these effects do converge on a limited number of states underlying the behavioral effects. The immobilizing effect of inhaled anesthetics involves actions in the spinal cord, whereas sedation/hypnosis and amnesia involve supraspinal mechanisms that interact with endogenous memory, sleep, and consciousness pathways and networks. Volatile inhaled anesthetics enhance inhibitory synaptic transmission postsynaptically by potentiating ligand-gated ion channels activated by γ-aminobutyric acid (GABA) and glycine, extrasynaptically by enhancing GABA receptors and leak currents, and presynaptically by enhancing basal GABA release. Inhaled anesthetics suppress excitatory synaptic transmission presynaptically by reducing glutamate release (volatile anesthetics) and postsynaptically by inhibiting excitatory ionotropic receptors activated by glutamate (gaseous, and to some extent volatile, anesthetics). No comprehensive theory of anesthesia yet describes the sequence of events leading from the interaction between an anesthetic molecule and its targets to the behavioral effects. Downloaded from ClinicalKey.com at Buddhist Tzu Chi General Hospital JC September 17, 2016. For personal use only. No other uses without permission. Copyright ©2016. Elsevier Inc. All rights reserved.

Upload: others

Post on 12-Oct-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

C h a p t e r 2 5

Inhaled Anesthetics: Mechanisms of ActionMISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR.

K e y P o i n t s

• Anesthesia consists of separable and independent components or physiologic substates, each of which involves distinct, but possibly overlapping, mechanisms at different sites in the central nervous system.

• The potency of general anesthetics correlates with their solubility in oil, indicating the importance of interactions with hydrophobic targets.

• General anesthetics act by binding directly to amphiphilic cavities in proteins. Binding sites can be identified by a combination of site-directed mutagenesis and high-resolution structural analysis of anesthetic binding. Mutations rendering putative effector proteins insensitive to inhaled anesthetics have been constructed and expressed in mice but have not generated breakthroughs analogous to the success of this strategy with intravenous anesthetics.

• The effects of inhaled anesthetics cannot be explained by a single molecular mechanism. Rather, multiple targets contribute to the component actions comprising the anesthetic effects of each anesthetic. However, these effects do converge on a limited number of states underlying the behavioral effects.

• The immobilizing effect of inhaled anesthetics involves actions in the spinal cord, whereas sedation/hypnosis and amnesia involve supraspinal mechanisms that interact with endogenous memory, sleep, and consciousness pathways and networks.

• Volatile inhaled anesthetics enhance inhibitory synaptic transmission postsynaptically by potentiating ligand-gated ion channels activated by γ-aminobutyric acid (GABA) and glycine, extrasynaptically by enhancing GABA receptors and leak currents, and presynaptically by enhancing basal GABA release.

• Inhaled anesthetics suppress excitatory synaptic transmission presynaptically by reducing glutamate release (volatile anesthetics) and postsynaptically by inhibiting excitatory ionotropic receptors activated by glutamate (gaseous, and to some extent volatile, anesthetics).

• No comprehensive theory of anesthesia yet describes the sequence of events leading from the interaction between an anesthetic molecule and its targets to the behavioral effects.

614

Despite the widespread clinical use of general anesthetics, our current understanding of their molecular, cellular, and network mechanisms is incomplete. This critical gap in the pharmacology of one of medicine’s most important drug classes not only impedes rational use of available anesthetics but also hinders the development of newer anesthetics that might selectively achieve the desirable end points of anesthesia with fewer adverse cardiovascu-lar, respiratory, and possibly neuropathologic side effects. Although major progress has been made in understand-ing the pharmacology of the intravenous anesthetics by molecular genetic approaches (see Chapter 30), the actions of the inhaled anesthetics at the molecular and

Downloaded from ClinicalKey.com at BuddhiFor personal use only. No other uses without permi

cellular levels are more enigmatic. It is still not possible to trace precisely the sequence of events that leads from inhaled anesthetic–target interactions, through ascending levels of biologic complexity, to the various behavioral effects that characterize the composite state of clinical anesthesia in humans. Nevertheless, investigations con-tinue to reveal fundamental principles of action and have led to a framework for understanding anesthetic effects at different organizational levels. The focus of this chapter is on the mechanisms involved in the principal therapeutic effects (anesthesia) and on the side effects of the inhaled anesthetics (Fig. 25-1), a chemically and pharmacologi-cally diverse group that includes the potent halogenated

st Tzu Chi General Hospital JC September 17, 2016.ssion. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 2: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

Chapter 25: Inhaled Anesthetics: Mechanisms of Action 615

Figure 25-1. Structure of representative general anes-thetics and a nonimmobilizer (F6) shown as space-filling models.

Halothane Isoflurane Sevoflurane

Nitrous oxide Xenon

Desflurane 1-chloro-1,2,2-trifluorocyclobutane

(F3)

1,2-dichlorohexafluorocyclobutane(F6)

TABLE 25-1 OVERVIEW OF CANDIDATE SITES OF ANESTHETIC ACTION

Site Effect Targets

Proteins Amphiphilic binding sites Conformational flexibility, ligand binding

Ion channels, receptors, signaling proteins

Action potential Nervous system Small reduction in amplitude Na+ channelsCardiovascular system Reduced amplitude, duration Ca2+ channels, K+ channels

Synaptic transmissionInhibitory Presynaptic terminal Enhanced transmitter release ?

Postsynaptic receptors Enhanced transmitter effects Glycine, GABAA receptorsExcitatory Presynaptic terminal Reduced transmitter release Na+ channels, K2P channels

Postsynaptic receptors Reduced transmitter effects NMDA receptors, nicotinic acetylcholine receptors

Neuronal networks Neuronal circuitNeuronal integration

Altered long-term potentiation (LTP)/long-term depression (LTD)

Altered rhythmicity, coherence

Synaptic plasticityHCN channels, K2P channels,

extrasynaptic GABAA receptors, etc.Central nervous

systemNeocortex, hippocampus,

amygdalaSedation, amnesia Slow 1-Slow 4, δ-, α-, θ-, γ-rhythms,

cross-frequency couplingDiencephalon (thalamus),

brainstem (reticular formation)

Unconsciousness γ-band transfer entropy?, cross-frequency coupling? cortical integration capability?

Spinal cord Immobility Thalamic deafferentation?Nocifensive reflex

Cardiovascular system

Myocardium Negative inotropy Excitation-contraction coupling

Conduction system Dysrhythmias Action potentialVasculature Vasodilation Direct and indirect vasoregulation

HCN, Hyperpolarization-activated cyclic nucleotide; NMDA, N-methyl-d-aspartate.

ether (isoflurane, sevoflurane, desflurane, enflurane) and alkane (halothane) volatile anesthetics and the gaseous anesthetics (nitrous oxide and xenon). This critical sum-mary of the current state of knowledge begins with an historical overview and a review of the behavioral end points of anesthesia. We then trace, where possible, inhaled anesthetic effects through ascending levels of organization from molecules, cells, circuits, networks, and organs to mammalian behavior; an overview is provided in Table 25-1. We also briefly address studies of anesthetic effects in model organisms, with anesthetic end points

Downloaded from ClinicalKey.com at Buddhist Tzu

For personal use only. No other uses without permission

being identified that bear unclear relationships to those in mammals.1

HISTORY

THE UNIFIED PARADIGM OF ANESTHETIC THEORIES

The first monograph reporting experimental work on anesthetic mechanisms, proposing a soon-to-be

Chi General Hospital JC September 17, 2016.

. Copyright ©2016. Elsevier Inc. All rights reserved.
Page 3: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

tsion. Copyright ©2016. Elsevier Inc. All rights reserved.

PART III: Anesthetic Pharmacology616

discredited lipid-elution theory of anesthetic action, was published only 6 months after Morton’s Ether Dome dem-onstration. For decades thereafter, the phenomenon of anesthesia puzzled, inspired, and awed those who tried to understand it. A most influential paradigm of anesthetic action formulated by Claude Bernard in the 1870s posited that anesthesia was a “unified” phenomenon: a unitary mechanism applicable to all forms of life. Although the anesthetized state could be brought about by a variety of agents, its essence was the same in all living creatures. In fact, Bernard thought that life itself was defined by sus-ceptibility to anesthesia. Bernard also proposed a more specific theory of anesthesia, coagulation of protoplasm, which competed with a number of coexisting theories entertained by the scientific community. In a major work published in 1919, Hans Winterstein summarized the perplexing diversity of anesthetic theories by listing more than 600 references, the majority to original labora-tory work—a convincing testimony to the interest of the scientific world in this phenomenon. Of note, the work of Meyer and Overton at the end of the nineteenth cen-tury had only a limited effect on the trajectory of research until the 1960s.1 Only then was the striking simplicity of the Meyer-Overton correlation (Fig. 25-2, A) of anes-thetic potency with solubility in olive oil interpreted by the majority of researchers as an indicator that lipids are likely the anesthetic target. This interpretation focused attention on anesthetic effects on the bulk physical prop-erties of cell membranes, which were known at that time to consist primarily of lipid molecules. Such nonspecific or “lipoid-based” anesthetic theories dominated the field from the 1960s to the 1980s.

MINIMUM ALVEOLAR CONCENTRATION: A BRIDGE BETWEEN PAST AND PRESENT

The potencies of inhaled anesthetics for immobilization were established in the classic studies in the 1960s of Eger and colleagues,2,3 who defined the minimum alveolar con-centration (MAC) as the inhaled anesthetic atmospheric pressure required to prevent movement in response to a defined noxious stimulus in 50% of subjects. The concept of MAC evolved within a unitary paradigm of anesthetic action and reflected the priorities of clinical practice. As a result, prevention of movement (immobility) became a universal yardstick for anesthetic effects, presumed to occur in the brain. Moreover, the simple elegance of the relationship between MAC and lipid solubility (see Fig. 25-2, A) graphically illustrated Meyer and Overton’s con-clusion that “All chemically indifferent substances that are soluble in fat are anesthetics … their relative potency as anesthetics will depend on their affinity to fat on the one hand and water on the other hand, that is, on the fat/water partition coefficient.”1 This was interpreted as favoring lipids as the primary targets of anesthetics and a single nonspecific theory to explain anesthesia. The appeal of a single unified mechanism to explain anes-thesia was (and remains) intellectually appealing. This focused the bulk of research efforts on delineating how anesthetic interactions with lipid membranes might lead to the behavioral changes observed under anesthesia, the nonspecific lipoid theory.

Downloaded from ClinicalKey.com at BuddhisFor personal use only. No other uses without permis

6

5

4

3

2

1

0

–1–2 –1 0 1 2 3 4 5

6

5

4

3

2

1

0

–1–1 0 1 2 3 4 5 6

A

Methanol

Methanol

AcetoneEthanol

2-ButanoneParaldehyde

n-Propanol

n-Butanol

n-Hexanol HalothaneMethoxyflurane

n-Pentane

n-Hexane

n-Decanol

n-Nonanoln-Octanol

Benzyl AlcoholChloroform

Diethyl ether

Acetone

n-Propanol

n-Pentanol

n-Heptanol

n-Pentane

n-Hexanen-Decanol

n-Nonanoln-Octanol

n-HexanolHalothane

Methoxyflurane

Benzyl Alcohol

Log 1

0 (P

oten

cy fo

r ge

nera

l ane

sthe

sia)

Log10 (Lipid bilayer/water partition coefficient)

n-Butanol

Ethanol

B

Log 1

0 (P

oten

cy fo

r ge

nera

l ane

sthe

sia)

Log10 (Potency for luciferase inhibition)

Figure 25-2. General anesthetics act by binding directly to pro-teins. A, The iconic Meyer-Overton correlation (ca. 1900) between anesthetic potency and the lipid-water partition coefficient was ini-tially interpreted as evidence that lipids of nerve membranes were the principal anesthetic target sites. B, Progress in the twentieth century showed that general anesthetic potencies correlate equally well with their ability to inhibit activity of the soluble firefly enzyme luciferase, which is not a physiologically relevant anesthetic target itself but serves as a lipid-free model protein for anesthetic binding. Inset, The crystal structure of luciferase with bound anesthetic (red). (Reprinted with permission from Franks NP, Lieb WR: Molecular and cellular mecha-nisms of general anesthesia, Nature 367:607-614, 1994.)

Tzu Chi General Hospital JC September 17, 2016.

Page 4: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

C

Because inhaled anesthetic concentrations reflect con-centrations in the tissues after equilibration, which is most rapidly achieved for well-perfused organs such as the brain and heart, MAC is analogous to the plasma con-centration for 50% effect (EC50) for intravenous anesthet-ics. In clinical applications, MAC is usually expressed as volume percent (vol%), which varies considerably with temperature and atmospheric pressure caused by changes in aqueous solubility, whereas the equivalent liquid-phase molar concentration is temperature- and pressure- independent.4 The MAC concept provided researchers and clinicians with a universal standard to measure a defined anesthetic end point (immobility), making meaningful comparisons of experimental results possible and accelerating clinical and laboratory research into anesthetic mechanisms. Today, a more nuanced under-standing of MAC takes into account the structural and functional diversity of the physiologic targets for the dif-ferent components of the anesthetic state.

SHIFT FROM LIPID- TO PROTEIN-CENTERED MECHANISMS

Lipid-centered mechanisms of anesthesia prevailed in the 2 decades after definition of the MAC concept. Alternative targets were repeatedly proposed but largely neglected by the scientific mainstream. Experimental inconsistencies of lipid targets,5,6 as well as evidence compatible with proteins as primary sites of action,7,8 remained largely unnoticed. A shift from lipid- to protein-centered mecha-nisms occurred in the 1980s, owing largely to the land-mark discoveries by Franks and Lieb,9,10 who in a classic series of publications demonstrated that protein targets were also compatible with the Meyer-Overton correlation (see Fig. 25-2, B)—a proof of concept that, within a cou-ple of years, redirected the bulk of research efforts toward proteins. As a corollary of this reorientation, evidence against lipid-based theories was recognized. Examples include the cutoff in anesthetic potency in homologous series of long-chain anesthetic alcohols, and the identifi-cation of hydrophobic drugs that do not obey the Meyer-Overton correlation.1,11 The enantiomeric selectivity of several anesthetics further strengthened the case for specific binding sites on proteins because stereoselectiv-ity is difficult to reconcile with lipid targets.12 Today, there is widespread (but not universal) acceptance of the notion that critical signaling proteins (e.g., ion channels or ligand-gated receptors) are the relevant molecular targets of anesthetic action, even though the submo-lecular mechanisms of their modulation by anesthetics are debated. The exact identity of proteins contributing to specific anesthetic end points continues to be sought, with research addressing not only the “where” (target) but also the “how” (process) of anesthetic mechanisms.

DIVERSITY OF POTENTIAL ANESTHETIC TARGETS

At high concentrations in vitro, most inhaled anesthetics affect the functions of multiple proteins, many of which can be plausibly connected to the components of the anesthetic state or anesthetic side effects. However, when

Downloaded from ClinicalKey.com at Buddhist Tzu For personal use only. No other uses without permission. C

hapter 25: Inhaled Anesthetics: Mechanisms of Action 617

considering a specific anesthetic end point, anesthetics are effective in vivo over quite a narrow concentration range. This makes the concentration at which a relevant anesthetic effect is observed a critical consideration for deciding potential relevance. The mechanistic relevance of small effects observed in vitro at relevant concentra-tions is less clear; that is, what effect is too small to be considered relevant to anesthesia?13-15 Whether anesthe-sia results from the sum of minor perturbations at mul-tiple sites that are amplified in cascades across multiple levels of integration to produce the macroscopic effect, or from strong effects on a small number of targets should be resolved as results of reductionist molecular studies are integrated into more complex molecular and cellular net-works, and as genetic studies are extended to encompass inhaled anesthetics. The picture that is emerging includes multiple cellular and molecular targets in distinct brain regions that are involved in both the desired and adverse effects of general anesthetics.

ANESTHESIA: A COMPOSITE NEUROPHARMACOLOGIC STATE

Along with progress in identifying the molecular mecha-nisms of anesthesia, our understanding of the nature of the anesthetic state has evolved. Whereas a drug-induced comalike state of general anesthesia can be induced by inhaled anesthetics administered at appropriate concen-trations (approximately 1.3 times MAC, equivalent to the EC95 of a volatile anesthetic), the use of such high concentrations can lead to short-, and possibly long-term, side effects. It is now clear that anesthesia consists of separable and at least partially independent compo-nents or substates, each of which involves distinct, but possibly overlapping, mechanisms in different regions of the central nervous system (CNS) and with variations in relative potencies between specific agents.16 Immobiliza-tion, the core measure of MAC, is mediated largely at the level of the spinal cord by inhaled anesthetics17,18 but not by barbiturates.19 On the other hand, the spinal cord is unlikely to be the major site of such phenomena as amne-sia, sedation, and unconsciousness, which rather are produced by anesthetic effects on cerebral cortical func-tion (Fig. 25-3). A functional separation between amne-sia and sedation has been demonstrated for intravenous anesthetics,20 and it seems likely that this will apply to inhaled anesthetics as well. The state commonly referred to as “unconsciousness” is in itself heterogeneous, with evidence for distinct states of unresponsiveness and unconsciousness.21 These and similar findings have led to the concept that general anesthesia consists of multiple independent components that can be resolved experi-mentally and clinically.

In principle, each component of anesthesia can be preferentially induced in a concentration- and agent-specific manner using individual cellular/molecular pathways in various regions of the CNS. For example, injections of pentobarbital into discrete sites in the meso-pontine tegmentum induce a comatose state,22 whereas sedation induced by systemic administration of propofol can be reversed by microinjections of γ-aminobutyric acid (GABA)A receptor antagonists into the tuberomammillary

Chi General Hospital JC September 17, 2016.opyright ©2016. Elsevier Inc. All rights reserved.

Page 5: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology618

nucleus, a sleep-regulating nucleus in the hypothala-mus.23 Thus, general anesthetics produce separate iden-tifiable anesthetic substates via agent-specific actions at discrete anatomic sites in the CNS through different molecular targets. An important consequence of this complexity is that MAC, which is based exclusively on a motor response, might not proportionately reflect other components of anesthesia. Although this heterogeneity of anesthetic actions complicates a mechanistic under-standing, it does open the possibility of developing sub-state-specific drugs.

INTEGRATED EFFECTS ON CENTRAL NERVOUS SYSTEM FUNCTION

IMMOBILITY

Electroencephalography as a monitor of brain activity has been applied both to the study of anesthetic mecha-nisms and as a monitor of the anesthetic state (also see Chapter 44). Failure to find a correlation between quan-titative electroencephalographic activity and immobility in response to noxious stimulation led to the somewhat radical (at the time) hypothesis that immobility was not a cerebral cortex–mediated phenomenon.24 Experimental demonstration that volatile anesthetics act on the spinal cord to suppress movement17,18 supported this hypoth-esis and was a major factor leading to the contemporary separation of anesthetic substates, of which immobility

0

50

100

0.750.500.25 1.0

Res

pond

ers

(%)

MAC fraction

Am

nesiaU

nconsciousness

Imm

obility

Figure 25-3. Multiple behavioral end points and sites of action underlie inhaled anesthetic action. Amnesia, the most sensitive anes-thetic end point, probably involves the hippocampus, amygdala, mediotemporal lobe, and other cortical structures. Unconsciousness likely involves the cerebral cortex, thalamus, and reticular forma-tion. Sedation and hypnosis (loss of responsiveness) are part of the consciousness-unconsciousness continuum and are not shown. Immo-bility occurs by anesthetic action in the spinal cord, although supra-spinal effects (dotted arrow) are likely important for some anesthetics. Anesthetic action in the spinal cord blunts ascending impulses arising from noxious stimulation and might indirectly contribute to anes-thetic-induced unconsciousness and amnesia (dashed arrow). Cardio-vascular responses occur at even greater MAC fractions (not shown). (Courtesy Joseph Antognini, University of California, Davis.)

Downloaded from ClinicalKey.com at BuddFor personal use only. No other uses without perm

requires the highest drug concentrations (see Fig. 25-3). Taking advantage of the unusual blood supply of goat CNS, which allows separate perfusion of brain and spinal cord, Antognini and associates17,25 showed that immobi-lity involves anesthetic effects at the spinal level because selective delivery of isoflurane or halothane only to the brain required 2.5-fold to 4-fold higher concentrations. At the same time, experiments by Rampil and co-workers18 that used surgical separation of the forebrain and midbrain from the spinal cord in rats led to the conclusion that immobilization involves primarily suppression of the nocifensive withdrawal reflex arc at the level of the spinal cord (Fig. 25-4).

In the 20 years since the identification of the spi-nal cord as the site of anesthetic-induced immobil-ity, research has centered on pharmacologic, genetic, and complex network approaches. The pharmacologic approach to identify receptor-level contributions to iso-flurane-induced immobility (isoflurane being the stan-dard potent ether for experimental purposes) yielded the surprising insight that actions at GABAA receptors appear to be unimportant.26 Perhaps less surprisingly, inhibition of central nicotinic acetylcholine receptors also plays no role in immobilization.27 A role for voltage-gated sodium (Na+) channels was suggested by the finding that intra-thecal administration of a selective inhibitor of Na+ chan-nels potentiates anesthetic immobility (reduces MAC), whereas a Na+ channel activator does the opposite.28 Anesthetic-resistant transgenic mice confirmed that GABAA receptors containing α1- or α3-subunits do not contribute to the immobilizing action of isoflurane.29,30 In contrast, mutant mice lacking the TASK-1, TASK-3, or TREK-1 K2P channels have higher MAC values for volatile anesthetics but not intravenous anesthetics,31-33 indicat-ing a role for these channels, possibly by a presynaptic mechanism.34

Cerebralcortex

A

B

Noxiousstimulation

Spinothalamictract

Figure 25-4. Inhaled anesthetics produce immobility at the spi-nal level. A, Decerebration by removal of the forebrain rostral to the black line does not alter the MAC of isoflurane in rats, indicat-ing that volatile anesthetic immobilization does not depend on the cerebral cortex.17,18 B, Anesthetics suppress the nocifensive with-drawal reflex response to noxious stimulation transmitted to the dorsal horn by sensory nerves at the spinal level. Current efforts are focused on identifying the molecular, cellular, and anatomic substrates for this effect.

hist Tzu Chi General Hospital JC September 17, 2016.ission. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 6: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

C

Work with physiologic preparations that attempt to preserve parts of the complex spinal cord circuitry sug-gests that anesthetic inhibition of afferent (noxious sen-sory) input to the dorsal horn plays a subordinate role to the suppression of the efferent (motor) output from the ventral horn, although this may vary by specific agent. This motor output is coordinated by neuronal networks organized in so-called central pattern generators that con-trol the activity of cholinergic motoneurons.35 Not unlike understanding the anesthetic effects on higher cognitive function, the key to understanding immobility will likely lie in resolving the effect of anesthetics on integrated spi-nal network activity.

UNCONSCIOUSNESS

Consciousness, as a quality of the mind, is easy to rec-ognize but difficult to define (see Chapters 13 and 14). It has been described as “what abandons us every night when we fall into dreamless sleep and returns the next morning when we wake up or when we dream.” This intuitive description applies equally as well to general anesthesia as it does to sleep; however, it does not pro-vide a concrete definition that is useful for scientific studies. Consciousness consists of inner qualitative subjective states of sentience or awareness.36 A recent suggestion that seeks to provide a quantitative measure of consciousness, and that may serve as a definition, is that consciousness is the capacity of a system to inte-grate information.37 Loss of consciousness (or hypnosis) is a hallmark of the onset of anesthesia, and anesthetic drugs are being used as tools for understanding the neural correlates of consciousness (NCC).38 However, what is commonly referred to as unconsciousness under anesthesia might be better described as unresponsiveness, which could obscure states of self- and environmental awareness lacking explicit memory traces.21

Although it is a relatively new field, particularly within the field of anesthesiology, the “science of con-sciousness” has generated considerable interest and has led to a number of specific hypotheses that are experi-mentally tractable. The “thalamic theory” of anesthesia proposes a somatosensory deafferentation by anesthetic action in the thalamus as a mechanism of unconscious-ness.39 In support of this hypothesis, isoflurane hyper-polarizes and shunts thalamic neurons,40 an action consistent with impaired thalamic transfer of sensory information.41 Functional human brain imaging shows preferential suppression of thalamic activity by some, but not all, anesthetics, and this has led to the “tha-lamic switch” hypothesis.39 However, loss of conscious-ness occurs within a very narrow range of anesthetic concentrations,42 typically up to 0.5 MAC,43,44 whereas quantifiable effects in the thalamus occur above this concentration range and are typically incremental (dim-merlike) rather than sudden (switchlike). A comprehen-sive theory of anesthetic-induced unconsciousness must be more inclusive than a simple block of information transfer through the thalamus to be compatible with existing evidence45,46 and must also explain the suppres-sion of endogenous cortical activity that is generated without external stimulation.

Downloaded from ClinicalKey.com at Buddhist Tzu For personal use only. No other uses without permission. C

hapter 25: Inhaled Anesthetics: Mechanisms of Action 619

Contemporary neuroscience has replaced the Car-tesian view of an anatomically discrete brain struc-ture as the center of consciousness with the concept that consciousness requires integration of information between multiple brain regions across large-scale cere-bral networks.45,47 The rich connectivity of the cerebral cortex and its hierarchical organization are especially suited to enable high levels of information integration in the human brain. Some brain areas present a “rich-club” organization (i.e., highly connected nodes tend to be preferentially connected to other highly con-nected nodes), which has been suggested to be optimal for information integration.48,49 These hubs might be promising targets for the hypnotic action of general anesthetic drugs.

Anesthetics might act by interfering with the opera-tional synchronicity and coherence of these networks. Consequent disruption of cortical functional and effec-tive connectivity has been observed during natural slow wave sleep50 and midazolam-induced loss of respon-siveness.51 This breakdown of cortical connectivity, rather than pharmacologic deafferentation from the environment, could underlie loss of consciousness.45 Unconsciousness would then be characterized not by the absence but by the fragmentation of cortical pro-cessing. Although the mechanism of “binding” (i.e., creating the unity of perception) is uncertain, syn-chronicity of neuronal activity in the 40- to 90-Hz range across functionally connected cortical areas (commonly referred to as 40 Hz- or γ-rhythm) is a viable candidate. Animal52,53 and human42 data implicate activity in the γ-band throughout the cortex as a network-level target of general anesthetics. Anesthetic actions on cortical information processing probably consist not merely of suppression of responses but of reduced complexity and variability reflected counterintuitively in increased reli-ability and precision of evoked responses.54,55

LEARNING AND MEMORY

Anterograde amnesia, one of the core desirable anesthetic end points, is achieved at lower anesthetic concentra-tions (∼0.25 MAC) than those required for unconscious-ness (∼0.5 MAC). Perhaps the closest analogue in rodents to explicit memory in humans is medial temporal lobe–dependent learning of temporal and spatial sequences known as hippocampus-dependent spatial learning. This can be tested by a variety of experimental paradigms, includ-ing fear conditioning to context (Fig. 25-5). Other learn-ing paradigms, such as fear conditioning to tone, are by contrast independent of the hippocampus. Isoflurane and the nonimmobilizer F6 both inhibit hippocampus-dependent learning at about one-half the concentra-tion necessary for disrupting hippocampus-independent learning.56 Similarly, anesthetic concentrations that inhibit explicit memory in humans (memory that can be explicitly recalled as opposed to motor learning, clas-sical conditioning, and so on) are similarly lower than concentrations that impair implicit memory (not subject to willful recollection).46 Taken together, these findings implicate effects on function of the medial temporal lobe, including the hippocampus, in the suppression of explicit

Chi General Hospital JC September 17, 2016.opyright ©2016. Elsevier Inc. All rights reserved.

Page 7: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology620

Figure 25-5. Differential sensitivity of different types of learning to anesthetics and nonimmobilizers. Freezing in anticipation of a noxious stimulus is a measure of learning in rats; less freezing indicates less learning. Left, The learning protocol involves pre-equilibration of rats in the equilibration chamber to isoflurane or the nonimmobilizer F6 at the desired concentration before placement into the training chamber. For test-ing of memory to context, the training and test chamber are identical. For testing of memory to tone, training and testing take place in differ-ent chambers. Right, Hippocampus-dependent learning (fear conditioning to context, closed symbols) is inhibited by isoflurane (circles) at lower concentrations than hippocampus-independent learning (fear conditioning to tone, purple squares). This differential sensitivity is mirrored by the nonimmobilizer F6 (blue circles and blue squares for context and tone, respectively). (Left panel adapted with permission from Eger EI 2nd, et al: Isoflurane antagonizes the capacity of flurothyl or 1,2-dichlorohexafluorocyclobutane to impair fear conditioning to context and tone, Anesth Analg 96: 1010-1018, 2003; right panel data points reconstructed from Dutton RC, et al: Short-term memory resists the depressant effect of the nonimmobilizer 1-2-dichlorohexafluorocyclobutane (2N) more than long-term memory, Anesth Analg 94:631-639, 2002, and Dutton RC, et al: The concentration of isoflurane required to suppress learning depends on the type of learning, Anesthesiology 94:514-519, 2001.)

0

100

70

60

40

20

0.40.20.0 0.6 0.8 1.0

Free

zing

(%

)

MAC fraction

F6 contextF6 toneIsoflurane contextIsoflurane tone

Equilibrationchamber

Training and contexttesting chamber

Tonetesting chamber

Speaker

Speaker

Shockgrid

memory by anesthetics. Effects on other structures, such as the amygdala, may be relevant to anesthetic impair-ment of implicit or other types of memory.57

Comparison of the amnesic potency of five inhaled anesthetics using an inhibitory avoidance paradigm revealed that nitrous oxide was the most and halothane the least potent amnestic (expressed in MAC fractions), with the halogenated ethers of intermediate potency.58 Because inhaled anesthetics affect multiple cellular tar-gets at amnesic concentrations, it is difficult to attri-bute amnesia to specific cellular mechanisms. It is also unclear whether inhibition of learning and memory by drugs known to have different receptor affinities shares common mechanisms at some level of integration. Comparison with more selective drugs provides some insight. Yet, θ-rhythms (4-12/Hz) are clearly important for hippocampus-dependent learning and memory.59 Benzodiazepines60 and cannabinoids61 slow and suppress hippocampal θ-rhythms in proportion to their ability to impair hippocampus-dependent learning (also see Chap-ters 13 and 30). Isoflurane and the nonimmobilizer F6 have comparable effects on θ-rhythms at amnesic con-centrations while having different receptor-level profiles and opposite effects on sedation.62 Thus alterations in neuronal synchrony provide a common network-level substrate for memory impairment. The synchronization between amygdalar and hippocampal θ-rhythms that occurs during fear memory retrieval indicates that this

Downloaded from ClinicalKey.com at BuddhFor personal use only. No other uses without perm

principle might also apply to other forms of memory and their impairment by anesthetics.63 As with other com-ponents of the anesthetic state, the precise mechanisms of memory impairment by anesthetics, and of memory itself, remain to be fully elucidated.

SEDATION

Sedation (defined as a decrease in activity, alertness, arousal, and/or vigilance), which is on a behavioral continuum leading to hypnosis, is achieved at anes-thetic doses similar to those that produce amnesia (<0.5 MAC). There is no clear mechanistic or clinical separa-tion between sedation and hypnosis. By contrast, even though sedation can be difficult to separate from amne-sia, for intravenous anesthetics (see Chapter 30) there may be separate but overlapping substrates for these two end points.20,64 The mechanisms involved in these behavioral effects are likely to resemble those of less pro-miscuous drugs, for which genetic approaches have been informative. An amino acid knock-in mutation (H101R) in mice that renders the α1 GABAA receptor subunit insensitive to modulation by benzodiazepines produces resistance to the sedative and amnesic effects of benzo-diazepines while maintaining other behavioral effects, among them anxiolysis.65 The α1 subunit is abundantly expressed in the CNS, mainly in the cortical areas and thalamus. Volatile anesthetics have qualitatively similar

ist Tzu Chi General Hospital JC September 17, 2016.ission. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 8: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

C

effects on α1-containing GABAA receptors (but also those containing other subunits) at low concentrations. The observation that the nonimmobilizer F6, which is devoid of sedative properties,62 is amnesic66 but does not mod-ulate benzodiazepine-sensitive α1-containing GABAA receptors,67,68 is compatible with a role for α1-containing receptors in volatile anesthetic-induced sedation, because few other targets are affected at purely sedative concen-trations. Possible targets for the sedative effects of the gaseous anesthetics nitrous oxide and xenon, which do not affect GABAA receptors, include N-methyl-d-aspartate (NMDA) receptor antagonism69 and K2P channel activa-tion.70 Consistent with this distinct pharmacologic pro-file, nitrous oxide has strikingly different effects from benzodiazepines in behavioral tests aimed at evaluating sedation in mice.71

Recognizing that there may be more than a super-ficial similarity between natural sleep and anesthetic-induced sedation and hypnosis, the effects of some anesthetics apparently share natural sleep mechanisms by directly activating discrete sleep-promoting nuclei in the hypothalamus.23 Electroencephalographic pat-terns during natural slow-wave sleep and anesthesia show similarities,72 and recovery from sleep depriva-tion can occur under propofol73 and inhalational anes-thesia,74 supporting this concept (also see Chapter 14). Anesthetic effects on other cortical75 and subcortical structures21 may also contribute to anesthetic-induced sedation and hypnosis.

ANESTHETIC NEUROTOXICITY AND NEUROPROTECTION

Early Postnatal NeurotoxicityGeneral anesthetics can cause persistent neurologic effects beyond their classic reversible production of anes-thesia. Yet its clinical significance remains unknown.76-78 Since the discovery that postnatal exposure of the devel-oping rodent brain to high doses of commonly used anesthetic drugs, in isolation or in combination, for a period of several hours, can induce apoptotic cell death with potentially long-term functional consequences,78 the neonate rodent model has been extensively studied and discussed. Similar effects have been reproduced with all commonly used general anesthetic drugs. How the results translate quantitatively and qualitatively from short-lived altricial (such as rodents) to long-lived, more precocial (such as Homo sapiens) species remains to be defined. Interestingly, postexposure environmental fac-tors appear to play a substantial role in the expression of neurotoxicity.79

Ischemic NeuroprotectionPharmacologic neuroprotection from cerebral ischemia and reperfusion is a rapidly evolving area of research that is characterized by great potential, but comparable disappointments, because results from promising ani-mal studies have failed to translate into clinical benefits. Despite considerable research, clinical evidence for isch-emic brain protection by inhaled anesthe tics remains controversial.80 Ischemic neuronal injury results from

Downloaded from ClinicalKey.com at Buddhist Tzu CFor personal use only. No other uses without permission. C

hapter 25: Inhaled Anesthetics: Mechanisms of Action 621

early excitotoxic cell death mediated by excessive release of excitatory amino acid transmitters such as glu-tamate and oxidative stress caused by reperfusion injury together with delayed cell death as a result of apopto-sis.81,82 Volatile anesthetics (e.g., isoflurane) and xenon exhibit early neuroprotection in animal models that is sustained only for mild insults. This is consistent with a beneficial effect on excitotoxicity but minimal effects on delayed cell death caused by apoptosis. A therapeu-tic strategy to extend this early neuroprotection is sug-gested by the finding that isoflurane combined with caspase inhibitors to prevent apoptosis results in more sustained neuroprotection.83 In contrast, xenon appears to have an intrinsic antiapoptotic action that contrib-utes to its neuroprotective properties, consistent with its distinct molecular mechanisms.84 Interestingly, “pre-conditioning” with ketamine appears to protect against ketamine-induced neurotoxicity.85 In addition, volatile anesthetics probably protect by suppression of brain energy requirements owing to inhibition of excitatory transmission and potentiation of inhibitory recep-tors and ion channels.81 The potential for beneficial effects from anesthetic-induced preconditioning has also received considerable recent attention. Similar to the protective effects of anesthetics administered before cardiac ischemia (anesthetic preconditioning; see later), protective effects have also been observed in animal models of focal cerebral ischemia (see also Chapters 67, 69, and 70).86

Postoperative Cognitive EffectsAdverse postoperative cognitive effects persisting beyond a duration explainable by pharmacokinetic factors have been attributed to anesthetics since the nineteenth century. Three clinical entities must be distinguished: delirium, dementia, and postoperative cognitive dys-function (POCD). Contrary to delirium and dementia, POCD is not a clinical diagnosis but a result of compari-son of pre- and postoperative scores on neuropsycho-logical test batteries among patients undergoing surgery and matched control populations that did not undergo surgery. This is a highly complex experimental para-digm that is prone to artifacts. For example, as a result of these analyses, postoperative cognitive improvement coexists with POCD,87 and whether either or both of these entities exists as pathologic entities and are clini-cally relevant remains an open question. The situation is fundamentally different with respect to delirium and neurodegeneration, which can be diagnosed by estab-lished criteria and standardized tools.

Although direct effects of inhalational anesthet-ics have been linked to memory deficits persisting for days in young adult mice, specifically via interaction with the α5 subunit of the GABAA receptor,88 postan-esthesia changes in exploratory behaviour are depen-dent on interaction with central cholinergic signaling pathways.89 Accumulating evidence, however, points to immune- and/or inflammation-mediated changes triggered by surgical trauma and anesthesia as poten-tial mechanisms for delirium-like, short- or intermedi-ate-term postoperative cognitive dysfunction (also see Chapter 99).90,91

hi General Hospital JC September 17, 2016.opyright ©2016. Elsevier Inc. All rights reserved.

Page 9: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology622

Experimental evidence obtained using mice rendered genetically susceptible to Alzheimer disease–like neuro-degeneration does not support a role of anesthetic agents per se in promoting neurodegenerative disease.92 It pro-vides experimental support for observational studies in humans that cognitive decline after surgery is caused by neither accelerated neurodegeneration nor anesthetic agents.93

Rather than an anesthetic-mediated effect, a determin-ing role for the underlying disease trajectory (as opposed to the surgical/anesthetic intervention) in determining the long-term cognitive course of most patients is likely.94 Short-term effects on cognition appear to be related to physiologic perturbations resulting from invasive inter-ventions with either aggravating and/or mitigating roles of anesthetics.

INTEGRATED EFFECTS ON THE CARDIOVASCULAR AND RESPIRATORY SYSTEMS

Mechanisms of Cardiovascular EffectsThe classic cardiovascular effects of the volatile anes-thetics are usually undesired deleterious side effects that limit their safe use in the critically ill, but recent stud-ies have shown that there are also direct cardioprotec-tive effects.95 All volatile anesthetics produce agent- and dose-dependent reductions in myocardial contractil-ity, systemic vascular resistance, and cardiac preload, with consequent reduction in mean arterial pressure. However, significant differences do exist in the relative potencies of these effects between anesthetics (see Chapter 28).96 Volatile anesthetics depress contractility by reduc-ing Ca2+ availability and/or Ca2+ sensitivity of the con-tractile apparatus.97 The major targets responsible for the negative inotropic effects of volatile anesthetics are cardiac Ca2+ channels, sarcoplasmic Ca2+ handling, and the contractile apparatus. Inhibition of depolarization-induced increases in myoplasmic Ca2+ concentration occur primarily via inhibition of cardiac L-type voltage-gated Ca2+ currents and shortening of action potential duration.98 Volatile anesthetics also inhibit the sarco-endoplasmic Ca2+ adenosine triphosphatase (SERCA), whereas halothane, but not isoflurane or sevoflurane, opens sarcoplasmic reticulum (SR) Ca2+ release chan-nels (ryanodine receptors), depleting SR Ca2+ content and reducing excitation-evoked SR Ca2+ release.97 The resulting negative inotropic effect of this reduced Ca2+ availability is enhanced by reductions in the Ca2+ sen-sitivity of myofibrils. In contrast, xenon has no effect on ventricular contractility, conduction, or major cation currents.98,99 consistent with its lack of significant car-diovascular effects. Nitrous oxide causes mild reductions in ventricular function99 by poorly characterized effects on Ca2+ availability. This is usually accompanied by sympathetic nervous system stimulation that increases vascular resistance and counteracts the myocardial depression.100,101

Volatile anesthetics can produce vasodilation at clinical concentrations.102 (Also see Chapter 28.) The vascular effects of volatile anesthetics are multifactorial

Downloaded from ClinicalKey.com at BuddhFor personal use only. No other uses without perm

and tissue specific, and the precise cellular mechanisms are poorly understood.103,104 Peripheral vasodilation is mediated by both direct endothelium-independent vasodilating effects on vascular smooth muscle cells and indirect effects involving the sympathetic nervous system and the vascular endothelium. The mechanisms of these effects involve agent-specific effects on presyn-aptic norepinephrine release, inhibition of smooth mus-cle Ca2+ influx via L-type Ca2+ channels, activation of hyperpolarizing adenosine triphosphate (ATP)-sensitive K+ (KATP) and K-Ca channels, and endothelium-dependent factors, including nitric oxide production.105 As with the CNS, cardiovascular function depends on the integrated function of multiple ion channels, many of which are expressed in both of these excitable tissues. Volatile anesthetics have agent-specific effects on heart rate and induction of arrhythmias caused by actions on cardiac ion channels. Because multiple cardiac ion channels are sensitive to volatile anesthetics at clinical concentrations, and because most manipulations of car-diac ion channel function are potentially proarrhythmic, it is difficult to link anesthetic arrhythmogenicity to actions on specific channels.106 Electrophysiologic studies indicate that cardiac L-type Ca2+ channels, which are critical to the plateau phase of the cardiac action potential and electromechanical coupling, are inhibited by volatile anesthetics, leading to shorten-ing of the refractory period. Multiple voltage-gated K+ channels are also inhibited and might predispose to arrhythmias by delaying repolarization. Alternatively, inhaled anesthetics can protect the heart against isch-emia and reperfusion injury, possibly involving anti-oxidant, antiinflammatory, and/or preconditioning mechanisms.107,108 Volatile anesthetics109 and xenon110 can mimic the strong cardioprotective effects produced by ischemic preconditioning (termed anesthetic pre-conditioning by analogy) through activation of vari-ous cytoprotective G-protein–coupled receptors and protein kinases, including protein kinase C (PKC), mitogen-activated protein kinases (MAP kinases), extracellular signal–regulated kinases (ERK), Akt (pro-tein kinase B), and tyrosine kinases.95,111 Although not entirely clarified, sarcolemmal and putative mitochon-drial KATP channel activation, involving PKC activa-tion and increased formation of free radicals and nitric oxide, are the likely end effectors of anesthetic precon-ditioning in the heart.

Mechanisms of Respiratory EffectsInhaled anesthetics also have important effects on the respiratory system. All volatile anesthetics cause signifi-cant respiratory depression at concentrations necessary for surgical anesthesia. The peripheral chemo-reflexes and upper airway patency are particularly sensitive to subanesthetic concentrations of volatile anesthetics.112 The mechanisms involved in these potentially serious effects involve depression of central respiratory networks mediated by depression of excitatory and facilitation of inhibitory transmission. The precise molecular targets responsible for the exquisite sensitivity of these networks to low concentrations of volatile anesthetics remain to be elucidated (see Chapters 19 and 27).

ist Tzu Chi General Hospital JC September 17, 2016.ission. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 10: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

IDENTIFICATION OF MOLECULAR SITES OF ANESTHETIC ACTION

CRITERIA FOR IDENTIFYING SITES RELEVANT TO ANESTHESIA

Specific criteria have been proposed to evaluate the rel-evance of the many potential molecular targets of anes-thetics.113 These criteria include:

1. Reversible alteration of target function at clinically rel-evant concentrations. This criterion requires comparable in vivo and in vitro sensitivity and depends on the anesthetic end point under consideration. For exam-ple, targets involved in immobility must be sensitive to anesthetics near MAC, whereas targets mediating amnesia must be affected at a fraction of MAC. Recent evidence for persistent effects of inhaled anesthetics demonstrable in the absence of continued anesthetic exposure is challenging the notion of reversibility for certain effects.

2. Expression of the target in appropriate anatomic locations to mediate the specific anesthetic end point. For example, immobilization by inhaled agents appears to involve primarily actions in the spinal cord independent of actions in the brain.

3. Concordant stereoselectivity of anesthetic effects in vivo and on the target in vitro. Without a specific pharmacologic antagonist of anesthesia, correlation between the ste-reoselective actions of general anesthetics in vivo and in vitro is a useful test of pharmacologic relevance of putative molecular targets. Stereoselectivity data cor-relating in vivo potency and in vitro receptor actions implicate GABAA receptors as a target for the anesthetic actions of etomidate, pentobarbital, neurosteroid anes-thetics, and possibly isoflurane.

4. Appropriate sensitivity or insensitivity to model anes-thetic and nonanesthetic compounds. Anesthetic halo-genated cyclobutanes together with structural analogs that do not produce anesthesia at concentrations pre-dicted to be anesthetic by the Meyer-Overton corre-lation (nonimmobilizers) can be used to discriminate relevant volatile anesthetic targets in vitro. For example, the anesthetic F3 (1-chloro-1,2,2-trifluorocyclobutane), but not the structurally similar nonanesthetic F6 (1,2-dichlorohexafluorocyclobutane), affect GABAA, glycine, AMPA, kainite, and 5-HT3 receptors, and Na+ channels, consistent with possible roles in immobility, whereas both F3 and F6 affect neuronal nicotinic, M1 muscarinic, 5-HT2C, and mGluR5 receptors, indicat-ing that these targets are not involved in immobility. F6 is interesting in that it lacks sedative and immobiliz-ing effects but does possess amnesic effects, hence the more accurate term nonimmobilizer, making it a useful pharmacologic tool for discriminating targets for these actions.

5. Predictable effects of genetic manipulations targeted to putative molecular targets. The effects of targeted dele-tion of specific molecules implicated as anesthetic tar-gets (knockout mutations) or genetic engineering to introduce specific mutations that modify anesthetic sensitivity (knock-in mutations) in model organisms

Downloaded from ClinicalKey.com at Buddhist TzuFor personal use only. No other uses without permission.

Chapter 25: Inhaled Anesthetics: Mechanisms of Action 623

provide powerful approaches to test the roles of putative molecular targets of anesthetic action. This approach has been particularly successful in implicat-ing specific GABAA receptor subtypes in the effects of the GABAergic intravenous anesthetics propofol and etomidate, where single amino acid substitutions in specific receptor subtypes eliminates anesthetic effects both in vitro and in vivo.114 Targeted mutations of putative anesthetic targets provide a bridge between in vitro observations and whole-animal experiments essential for demonstrating anesthetic end points. The existence of multiple targets and redundancy among ion channel subtypes make this a more challenging experimental approach for inhaled anesthetics com-pared with intravenous anesthetics (discussed later).

PHYSICOCHEMICAL PROPERTIES OF ANESTHETIC BINDING SITES

A convergence of x-ray crystallography, molecular model-ing, and structure-function studies indicate that inhaled anesthetics bind in hydrophobic cavities formed within proteins.14,115 The lipophilic (or hydrophobic) nature of these binding sites explains their adherence to the Meyer-Overton correlation. An element of amphiphilicity (pos-sessing both polar and nonpolar characteristics) is also required for effective interaction with these cavities, as indicated by improvements in the Meyer-Overton corre-lation with more amphipathic solvents (possessing both hydrophobic and hydrophilic properties).

FROM MODEL PROTEINS TO RECEPTORS

Identifying inhaled anesthetic binding sites on plausible target proteins is difficult because of their low-affinity interactions, the paucity of atomic resolution structures of pharmacologically relevant target proteins, and the lack of specific antagonists. Consequently, most anes-thetic binding sites have been identified in well-char-acterized model proteins for which three-dimensional atomic resolution structures are available, but which are not themselves relevant to anesthesia, such as luciferase and albumin.14,115 These studies indicate that anesthetics bind in pockets with both nonpolar and polar noncova-lent chemical interactions. Binding involves weak hydro-gen bond interactions with polar amino acid residues and water molecules, nonpolar van der Waals interactions, and a polarizing effect of the amphiphilic binding cavity on the relatively hydrophobic anesthetic molecules. Inter-nal cavities are important for the conformational flexibil-ity required for ion channel gating and ligand-induced signal transduction of receptor proteins. Occupation of a critical volume within these cavities by anesthetics pro-vides a plausible mechanism for alteration of receptor and ion channel function by selective stabilization of a particular confirmation (e.g., an open or inactivated state of an ion channel). Anesthetics also gain binding energy from the entropy generated by displacing bound water from these relatively promiscuous binding sites. Studies of glycine, GABAA, and NMDA receptors provide con-vincing evidence for the existence of anesthetic binding sites in critical neuronal signaling proteins.14 Amino acid

Chi General Hospital JC September 17, 2016.Copyright ©2016. Elsevier Inc. All rights reserved.

Page 11: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology624

Figure 25-6. X-ray structures of propofol and desflurane bound to a pentameric ligand-gated ion channel. A, Cartoon of membrane-plane view of the bacterial homologue of mammalian pentameric ligand-gated ion channels (Gloebacter violaceus [GLIC]) with a bound general anesthetic molecule. B, Molecular surface of the general anesthetic intrasubunit cavities (yellow) and neighboring intersubunit cavities (pink) of the whole pentameric channel. (Modified from Nury H, et al: X-ray structure of general anaesthetics bound to a pentameric ligand-gated ion channel, Nature

Desflurane

A B

Intersubunitcavity

Generalanestheticcavity

LinkingtunnelPropofol

Mem

bran

e

residues critical for volatile anesthetic actions, and, by inference, binding, have been identified in the α-subunit of the GABAA receptor.14

Structural studies using the more accessible prokaryotic homologues of eukaryotic ion channels have provided a powerful tool for the study of anesthetic binding sites in biologically plausible proteins. For example, both propo-fol and desflurane have been co-crystalized with GLIC, a bacterial homologue of eukaryotic inhibitory ligand-gated ion channels (glycine and GABAA receptors). Both anesthetics bind a common preexisting site in the upper part of the transmembrane domain between the trans-membrane segments of a single subunit (Fig. 25-6).116 Molecular modeling based on structurally homologous proteins has also been used to identify putative anes-thetic binding sites in the transmembrane domains of vertebrate GABAA and glycine receptors (Fig. 25-7). These models suggest that different drugs may either bind in different orientations within a single amphiphilic cavity or occupy different cavities within the protein, causing similar functional effects. Refinement of these molecu-lar models will continue to provide new insights in the molecular basis for general anesthetic action that can be experimentally tested. For example, potential sites of interaction of xenon and isoflurane with the NMDA receptor have also been identified using this approach. One site, which can contain up to three xenon atoms or one molecule of isoflurane, overlaps the known binding site for the co-agonist glycine in the NR1 subunit.117 This suggests that two chemically dissimilar inhaled anesthet-ics inhibit NMDA receptors by direct competitive inhibi-tion of co-agonist binding.

Molecular Targets of Inhaled AnestheticsIon channels have emerged as the most promising molec-ular targets for inhaled anesthetics. Neurotransmitter-gated ion channels, in particular GABAA, glycine, and NMDA-type glutamate receptors, are leading candidates owing to their appropriate CNS distributions, essential physiologic roles in inhibitory and excitatory synap-tic transmission, and sensitivities to clinically relevant

469:428-433, 2011.)

Downloaded from ClinicalKey.com at BuddFor personal use only. No other uses without perm

concentrations of anesthetics.16,118 Other ion channels that are sensitive to inhaled anesthetics include the hyperpolarization-activated cyclic nucleotide (HCN)-gated family of channels that give rise to pacemaker currents118 and regulate dendritic excitability, two-pore domain (K2P) “leak” K+ channels that maintain resting membrane potential in many cells,119 and voltage-gated Na+ and Ca2+ channels.118

Inhaled anesthetics can be divided into two classes based on their distinct pharmacologic properties. The first class is the potent inhaled (volatile) anesthetics, which exhibit positive modulation of GABAA receptors and also produce significant, anesthesia-compatible effects on a number of other receptors/channels, including enhancement of inhibitory glycine receptors, inhibition of excitatory NMDA and neuronal nicotinic acetylcholine receptors, activation of K2P channels, and inhibition of presynaptic Na+ channels. Intravenous anesthetics such as propofol and etomidate represent more potent and specific positive modulators of GABAA receptors. The second class is the gaseous inhaled anesthetics, which include cyclopropane, nitrous oxide, and xenon. These anesthetics are inactive at GABAA receptors but block NMDA receptors and activate certain K2P channels at clinical concentrations.

LIGAND-GATED ION CHANNELS

Potentiation of Inhibitory GABAA and Glycine ReceptorsThe ether anesthetics (including isoflurane, sevoflurane, and desflurane), the alkane anesthetic halothane, most intravenous anesthetics (including propofol, etomidate, barbiturates), and the neurosteroid anesthetics enhance GABAA and glycine receptor (GlyR) function. GABAA and GlyRs are members of the same cys-loop ligand-gated ion channel superfamily that also includes the cation-permeable nicotinic acetylcholine and 5HT3 receptors. GABAA receptors are the principal transmitter-gated Cl– channels in the neocortex and allocortex, whereas GlyRs

hist Tzu Chi General Hospital JC September 17, 2016.ission. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 12: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

Chapter 25: Inhaled Anesthetics: Mechanisms of Action 625

Figure 25-7. Putative anesthetic binding sites on GABAA receptors identified by molecular modeling. A, Molecular model of the mouse GABAA receptor built using homology modeling techniques with computational chemistry optimizations and molecular dockings. The amino acid back-bone is displayed in ribbon format and outlined by the transparent solvent-accessible molecular surface. Each of the five subunits is colored uniquely. The GABA binding site is noted in the extracellular domain, whereas the putative anesthetic binding pocket (ABP) for potentiation is noted in the outer third of the transmembrane domain between the α and β subunits. Two binding sites are shown but only one has computation-ally docked desflurane. B, Cross-section at the dotted line level in A showing orientation of subunits with pentameric symmetry about a central ion pore. C, Magnified region of intersubunit anesthetic binding site derived from region in B showing relevant amino acid positions (in space-filling format) interacting with desflurane (in ball-and-stick format at the same scale). (Courtesy the Bertaccini laboratory, Stanford University.)

A

B

C

Tran

smem

bran

edo

mai

n

Cl–

ASN 265

Desflurane

MET 286 LEU 231

Ext

race

llula

rdo

mai

n

GA

BA

ABP

u

fulfill this function in the spinal cord, with some overlap in the diencephalon and brainstem. Activated receptors conduct chloride ions, driving the membrane potential toward the Cl– equilibrium potential. Both receptors are inhibitory because the Cl– equilibrium potential is usu-ally more negative than the normal resting potential. Channel opening also reduces membrane resistance and “shunts” excitatory responses. Most functional GABAA and GlyRs are heteropentamers, typically consisting of three different GABAA subunits (e.g., two α, two β, and one γ or ∂) or two different GlyR subunits (three α and two β).120 The subunit composition of GABAA receptors determines their physiologic and pharmacologic proper-ties and varies between and within brain areas, as well as between different compartments of individual neu-rons. Examples are the preferential expression of the α5-subunit in the dendritic field of the hippocampal CA1 area (a region important for memory formation), of the α4-subunit in the thalamus, and of the α6-subunit in the cerebellum. Presence of a γ-subunit is required for benzo-diazepine modulation of GABAA receptors and can also influence modulation by inhaled anesthetics. Although the molecular mechanisms of receptor modulation by inhaled anesthetics are not clear, these receptors have been key to our understanding of anesthetic-receptor interactions. Using chimeric receptor constructs between

Downloaded from ClinicalKey.com at Buddhist Tz

For personal use only. No other uses without permission.

anesthetic-sensitive GABAA and insensitive GlyR sub-units, specific amino acid residues in transmembrane domains 2 and 3 critical to the action of inhaled anesthet-ics have been identified.121 This laid the groundwork for the construction of anesthetic-resistant GABAA receptors and the generation of transgenic mice with altered anes-thetic sensitivity (discussed later).

The related cation-permeable 5-hydroxytryptamine (serotonin)-3 (5HT3) receptors are similarly potentiated by volatile anesthetics.122 5HT3 receptors are involved with autonomic reflexes and also probably contribute to the emetogenic properties of volatile anesthetics (see Chapter 97).

Inhibition of Excitatory Acetylcholine and Glutamate ReceptorsNeuronal nicotinic acetylcholine receptors (nnAChRs), like the other members of the cys-loop family, are het-eropentameric ligand-gated ion channels, but are cation selective. They are composed of α- and β-subunits, but functional homomeric receptors can be formed by cer-tain α-subunits. In the CNS, nnAChRs are localized pri-marily presynaptically.123 Homomeric α7-receptors have high permeability to Ca2+ that can exceed that of NMDA receptors.123 In contrast to GABAA and GlyRs, nnAChRs pass cations when activated and therefore depolarize

Chi General Hospital JC September 17, 2016.

Copyright ©2016. Elsevier Inc. All rights reserved.
Page 13: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology626

the membrane potential. Receptors containing α4β2 sub-units are very sensitive to block by isoflurane and propo-fol.124,125 Relevance of nnAChR block to immobilization, sedation, and unconsciousness by inhaled anesthetics is unlikely because nnAChRs are also blocked by nonimmo-bilizers, although it is possible that they do contribute to amnesia.

NMDA receptors are a major postsynaptic receptor subtype of inotropic receptors for glutamate, the prin-cipal excitatory neurotransmitter in the mammalian CNS.126 Typical NMDA receptors, defined pharmacologi-cally by their selective activation by the exogenous agonist NMDA, are heteromers consisting of an obligatory GluN1 subunit and modulatory GluN2 subunits. Channel open-ing requires glutamate (or another synthetic agonist such as NMDA) binding to the GluN2 subunit while the endog-enous co-agonist glycine binds to the GluN1 subunit. NMDA receptors also require membrane depolarization to relieve voltage-dependent block by Mg2+. Depolariza-tion is typically provided by binding of glutamate to non-NMDA glutamate receptors (discussed later). Because of this requirement for both presynaptic transmitter release and postsynaptic depolarization, synaptic NMDA recep-tors function as coincidence detectors, and this charac-teristic is thought to be central to their role in learning and memory. NMDA receptors are also involved in the development of chronic pain, perhaps because of mech-anisms similar to those underlying synaptic plasticity, and in ischemia-induced excitotoxicity by virtue of their capacity to allow entry of the ubiquitous intracellular sig-nal Ca2+. The nonhalogenated inhaled anesthetics xenon, nitrous oxide, and cyclopropane have minimal effects on GABAA receptors but depress excitatory glutamatergic syn-aptic transmission postsynaptically via NMDA glutamate receptor blockade (see Fig. 25-8).70,127 Volatile anesthetics can also inhibit isolated NMDA receptors at higher con-centrations.128 Along with presynaptic inhibition of gluta-mate release, this might contribute to their depression of NMDA receptor–mediated excitatory transmission.

A second class of inotropic glutamate receptors includes the non-NMDA receptors, which are subdivided into α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) and kainate receptors based again on their sensitivities to selective exogenous agonists.126 Inhaled anesthetics only weakly inhibit AMPA receptors, hence this action is unlikely to be important in their actions.129 Interestingly, kainate receptors are enhanced by inhaled anesthetics, but this is unlikely to be involved in immo-bility because MAC is not altered in mice deficient in the GluR6 receptor subunit.130 Most evidence suggests that the principal mechanism for depression of glutamatergic transmission by volatile anesthetics is presynaptic, with minor contributions from postsynaptic receptor block-ade131-133 (see “Cellular Mechanisms”).

VOLTAGE-GATED AND OTHER ION CHANNELS

Inhibition of Excitatory Na+ ChannelsVoltage-gated Na+ channels are critical to axonal con-duction, synaptic integration, and neuronal excitability.

Downloaded from ClinicalKey.com at BuddhiFor personal use only. No other uses without permi

In contrast to findings in invertebrate giant axons,134 axonal conduction in small (0.1 to 0.2 μm) unmyelin-ated hippocampal axons is depressed by volatile anes-thetics,135,136 and small reductions in preterminal action potential amplitude significantly depress transmitter release and hence postsynaptic responses at a mam-malian synapse.137 Heterologously expressed mam-malian voltage-gated Na+ channels are sensitive to clinically relevant concentrations of volatile anesthet-ics. The Na+ channel family consists of nine homolo-gous pore-forming α-subunits with distinct cellular and subcellular distributions.138 Isoflurane and other vola-tile anesthetics inhibit the major mammalian Na+ channel isoforms, including neuronal (Nav1.2), skeletal muscle (Nav1.4), cardiac (Nav1.5), and peripheral (Nav1.8) isoforms.139 Volatile anesthetics, but not nonimmobiliz-ers, also inhibit native neuronal and nerve terminal Na+ channels,140-143 lending support to the notion that Na+

100 200 300 400 500

0

−2

−4

−6

−8

−10

0A

20015010050

0

−4

−8

−12

0B

Time (msec)

Inhi

bito

ry p

osts

ynap

tic c

urre

nt (

nA)

+Xenon

+Xenon

+Xenon

20s

Control

Control

GABA

200

pA

Time (msec)

Exc

itato

ry p

osts

ynap

tic c

urre

nt (

nA)

Figure 25-8. The actions of xenon on inhibitory GABAergic and excit-atory glutamatergic synapses in cultured rat hippocampal neurons. Xenon (3.4 mM, ˜1 MAC) has no significant effect on the inhibitory postsynaptic current (A) but depresses the excitatory glutamatergic synaptic current, almost exclusively the slow NMDA receptor-medi-ated component of the current (B). In contrast, the principal effects of ˜1 MAC isoflurane are a prolongation in the inhibitory current decay and a reduction in excitatory current peak height, with little change in time course (not shown, see Fig. 25-10). (Reprinted in modified form by permission from de Sousa SLM, et al: Contrasting synaptic activity of the inhalational general anesthetics isoflurane and xenon, Anesthesiol-ogy 92:1055-1066, 2000.)

st Tzu Chi General Hospital JC September 17, 2016.ssion. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 14: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

C

channel blockade contributes to the depression of synap-tic neurotransmitter release.143 In contrast, xenon has no detectable effect on Na+, Ca2+, or K+ channels in isolated cardiomyocytes.92 The demonstration that NaChBac, a prokaryotic homologue of voltage-gated Na+ channels, is also inhibited by volatile anesthetics opens the way for structure-function studies of these channels.141

Inhibition of Ca2+ ChannelsMultiple cellular functions depend on the tightly con-trolled concentration of intracellular free Ca2+ ([Ca2+]i), which is determined by the integrated activities of volt-age-gated Ca2+ channels, capacitative Ca2+ channels, plasma membrane and sarcoplasmic/endoplasmic reticu-lum Ca2+-adenosine triphosphatases (pumps), Na+/Ca2+ exchangers, and mitochondrial Ca2+ sequestration. Alter-ation of any of these mechanisms by anesthetics can affect the many cellular processes regulated by the second mes-senger actions of Ca2+, including synaptic transmission, gene expression, cytotoxicity, and muscle excitation–con-traction coupling. Excitable cells translate their electrical activity into action by Ca2+ fluxes mediated primarily by voltage-gated Ca2+ channels in the plasma membrane. Distinct Ca2+ channel subtypes are expressed in various cells and tissues and are classified pharmacologically and functionally by the degree of depolarization required to gate the channel, such as low voltage–activated (LVA; T-type) and high voltage–activated (HVA; L-, N-, R-, and P/Q-type) channels. More recently, the molecular iden-tity of their pore-forming α-subunits has been used for classification.144 There is convincing evidence that vola-tile anesthetics inhibit certain Ca2+ channel isoforms but not others.145

Inhibition of presynaptic voltage-gated Ca2+ chan-nels coupled to transmitter release has been proposed as a mechanism by which volatile anesthetics reduce excit-atory transmission.146 Indeed, N-type (Cav2.2) and P-type (Cav2.1) channels, which mediate Ca2+ entry coupled to neurotransmitter release, are modestly sensitive to volatile anesthetics,147,148 but not in all neuron types,149 suggesting the importance of auxiliary subunits, post-translational modification, or other potential modulators of anesthetic sensitivity. A modest contribution of R-type Ca2+ channels (Cav2.3) to anesthesia is suggested by their sensitivity to volatile anesthetics and a small increase in MAC produced by genetic deletion in mice.150 T-type Ca2+ channels are particularly sensitive to volatile anesthet-ics151 and nitrous oxide.152 However, mutant mice lacking a major neuronal T-type Ca2+ channel isoform (Cav3.1) have normal volatile anesthetic sensitivity, although the onset of anesthesia is delayed.153 Thus the role that inhi-bition of these or other Ca2+ channels play in the CNS effects of inhaled anesthetics is unclear.

By contrast, a role for Ca2+ channel inhibition in the negative inotropic effects of volatile anesthetics, promi-nent at higher doses, is well established. The force of myocardial contraction is determined by the magnitude of cytosolic Ca2+ increase after electrical excitation, the responsiveness of the contractile proteins to Ca2+, and sarcomere length. Negative inotropic effects of volatile anesthetics are mediated by reductions in Ca2+ availabil-ity, Ca2+ sensitivity of the contractile proteins, and rate of

Downloaded from ClinicalKey.com at Buddhist Tzu For personal use only. No other uses without permission. C

hapter 25: Inhaled Anesthetics: Mechanisms of Action 627

cytosolic Ca2+ clearance. Volatile anesthetics reduce the Ca2+ transient and shorten action potential duration in cardiomyocytes primarily by inhibiting L-type (Cav1.2) Ca2+ currents, resulting in a negative inotropic effect and arrhythmogenicity.97,106,154 In contrast, xenon does not depress myocardial function or inhibit L-type Ca2+, Na+, or K+ currents in isolated cardiomyocytes.98,99 Inhibition of trans-sarcolemmal Ca2+ influx through cardiac L-type Ca2+ channels plays a major role in the negative inotropic effects of volatile anesthetics—greatest for halothane—along with contributions from effects on myofilament Ca2+ sensitivity and sarcolemmal Ca2+ release.106,155

In contrast to voltage-gated Ca2+ channels that regu-late the influx of extracellular Ca2+, intracellular Ca2+ channels regulate Ca2+ release from intracellular stores, particularly the endoplasmic reticulum (ER) and sarco-plasmic reticulum (SR). These include 1,4,5-trisphosphate receptors (IP3Rs), regulated by the second messenger IP3, and ryanodine receptors (RyRs), which mediate the rapid release of intracellular Ca2+ critical to excitation-contrac-tion coupling in muscle. Volatile anesthetic–induced Ca2+ leak occurs via both IP3R and RyR channels, which leads to depletion of intracellular Ca2+ stores from the SR and ER. This blunts changes in intracellular Ca2+ in response to stimulation and contributes to the smooth muscle–relaxing properties of volatile anesthetics that underlie bronchodilation and vasodilation.156 Malignant hyperthermia susceptibility is a pharmacogenetic disor-der that manifests as a potentially fatal hypermetabolic crisis triggered by volatile anesthetics, particularly halo-thane. It is often associated with mutations in RyR1 and the physically associated L-type Ca2+ channel (Cav1.1) that functions as the voltage sensor.157 Volatile anesthet-ics activate the abnormal RyRs, resulting in uncontrolled intracellular Ca2+ release, muscle contraction, and meta-bolic activity158 (see Chapter 43).

K+ Channels and HCN ChannelsPotassium (K+) channels are an extremely diverse ion channel family noted for their varied modes of activa-tion. They regulate electrical excitability, muscle contrac-tility, and neurotransmitter release. They are important in determining input resistance and in driving repolar-ization, and thus they determine excitability and action potential duration. Given the large diversity in K+ chan-nel structure, function, and anesthetic sensitivity, it is not surprising that there is considerable diversity in their sensitivity and response to inhaled anesthetics159: from relatively insensitive (voltage-gated K+ channels Kv1.1, Kv3)160 to sensitive (some members of the 2-pore domain K+ channels [K2P] family), resulting in either inhibition, activation, or no effect on K+ currents.

Volatile anesthetic activation of certain “leak” K+ channels was first observed in the snail Lymnaea,161 although the molecular identity of the affected ion chan-nels was unknown. Activation of K2P channels by volatile and gaseous anesthetics, including xenon, nitrous oxide, and cyclopropane, was subsequently observed in mam-mals.162 Increased K+ conductance can hyperpolarize neurons, reducing responsiveness to excitatory synaptic input and possibly altering network synchrony. Targeted deletion of the TASK-1, TASK-3, and TREK-1 K2P channels

Chi General Hospital JC September 17, 2016.opyright ©2016. Elsevier Inc. All rights reserved.

Page 15: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology628

in mice reduces sensitivity to immobilization by volatile anesthetics in an agent-specific manner, implicating these channels as contributory anesthetic targets in vivo.32-34 Other members of this large family of K+ channels are also sensitive to volatile anesthetics.163

The recognition that inherited channelopathies are arrhythmogenic and constitute an important contribu-tor to sudden cardiac death,164 particularly in young children,165 highlights the importance of analyzing anesthetic modulation of cardiac ion channels. Recom-binant hERG (human ether-a-go-go related) channels are moderately inhibited by halothane, and their depres-sion likely contributes to arrhythmogenic effects of vola-tile anesthetics106,166; they are also involved in acquired (drug-induced) and inherited long QT syndrome. Car-diac inward-rectifying (KIR), voltage-gated (Kv), and Ca2+-activated K+ channels are generally relatively insensitive to clinical concentrations of volatile anesthetics and xenon.98,106,167 In contrast, there is considerable evidence that volatile anesthetics and xenon activate cardiac mito-chondrial and sarcolemmal KATP channels,107 an effect that plays a critical role in anesthetic preconditioning. Direct electrophysiologic effects of anesthetics with pre-conditioning properties have been demonstrated on both mitochondrial and sarcolemmal KATP channels, although the precise mechanisms remain to be clarified.

Volatile anesthetics also inhibit HCN pacemaker channels, reducing the rate of rise of pacemaker poten-tials and the bursting frequency of certain neurons that show autorhythmicity. They decrease the Ih conduc-tance in neurons168 and modulate recombinant HCN1 and HCN2 channel isoforms at clinically relevant concentrations.169 Because HCN channels contribute to resting membrane potential; control action potential fir-ing, dendritic integration, neuronal automaticity, and temporal summation; and determine periodicity and synchronization of oscillations in many neuronal net-works,170 the anesthetic modulation of these channels could play an important role in anesthetic effects on neu-ronal integrative functions.

INTRACELLULAR SIGNALING MECHANISMS

Cell signaling mechanisms are critical to all phases of organ function and have been attractive targets for explaining the broad effects of general anesthetics. Anes-thetics have poorly understood actions on intracellular cell signaling pathways, which include processes down-stream from cell surface receptors and ion channels, such as effects of second messengers, protein phosphorylation pathways, and other regulatory mechanisms.171

G Protein–Coupled ReceptorsA variety of signals, including hormones, neurotrans-mitters, cytokines, pheromones, odorants, and photons, produce their intracellular actions by interactions with metabotropic receptors that activate heterotrimeric gua-nine nucleotide–binding proteins (G proteins). In con-trast to the inotropic receptors that directly couple to ion-selective channels, G proteins act as indirect molecu-lar switches to relay information from activated plasma membrane receptors to appropriate intracellular targets.

Downloaded from ClinicalKey.com at BuddhisFor personal use only. No other uses without permis

Heterotrimeric G proteins consist of a large α-subunit and a smaller β/γ-subunit dimer, each expressed as multiple isoforms with distinct properties and downstream targets. G proteins regulate a plethora of downstream effectors to control the levels of cytosolic second messengers such as Ca2+, cyclic adenosine monophosphate, and inositol tri-phosphate. These in turn regulate effector proteins such as ion channels and enzymes, either directly or via sec-ond messenger–regulated protein phosphorylation path-ways. Drugs that act through G protein–coupled receptors (GPCRs), such as agonists for μ-opioid and α2-adrenergic receptors, can affect anesthetic sensitivity (reduce MAC). Inhaled anesthetics can also directly affect signaling via GPCRs.172 For example, volatile anesthetics activate mul-tiple olfactory GPCRs in the rat in vivo in a receptor- and agent-selective manner.173 Analogous effects on related GPCRs more relevant to critical anesthetic end points are possible but remain to be demonstrated. The observa-tion that both volatile anesthetics and nonimmobilizers inhibit mGluR5 glutamate receptors, 5-HT2A serotonin receptors, and muscarinic acetylcholine receptors sug-gests that these GPCR effects do not contribute to anes-thetic immobilization.174-176

Protein PhosphorylationPhosphorylation of proteins on specific serine, threo-nine, or tyrosine hydroxyl groups, a posttranslational modification involved in the regulation of many anes-thetic-sensitive receptors and ion channels, is pivotal to synaptic plasticity (e.g., long-term potentiation [LTP]). Phosphorylation is controlled by the balance of activ-ity between protein kinases and phosphatases, several of which are plausible anesthetic targets. The protein kinase C (PKC) family of multifunctional protein kinases is acti-vated by the lipid-signaling molecule diacylglycerol and is involved in the regulation of many ion channels and receptors. Halothane177 and sevoflurane178 enhance the activity of some PKC isoforms and stimulate phosphory-lation of specific PKC substrates. Structural studies have identified a potential binding site in the diacylglycerol binding domain of PKCδ, consistent with the ability of certain anesthetics to mimic this natural regulator by binding to the activating site.179 A specific role for a direct pharmacologically relevant effect mediated by anesthetic activation of PKC or of any other kinase has yet to be demonstrated. Intrathecal injection of isoform-specific inhibitors of PKC does not affect sensitivity to halothane in vivo.180 Knock-out mice lacking the PKCγ isoform show normal sensitivity to halothane and desflurane while iso-flurane MAC was increased,181 suggesting that PKC is not critical to volatile anesthetic immobilization.

An important role for effects of volatile anesthetics and xenon on cell signaling mechanisms has been dis-covered for anesthetic-induced preconditioning in the heart (see Chapter 28) and brain against ischemic dam-age.83,87,81,85,109 Anesthetic-induced and ischemic cardiac preconditioning share critical signaling mechanisms, including activation of multiple GPCRs (e.g., adenosine, opioid, adrenergic) and protein kinases (e.g., src kinase, PKCδ, PKCɛ, Akt, mitogen-activated protein kinases [MAPKs]) and their downstream targets, particularly sar-colemmal and/or mitochondrial KATP channels, possibly

t Tzu Chi General Hospital JC September 17, 2016.sion. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 16: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

C

initiated by changes in reactive oxygen species as the crit-ical second messenger.95 Volatile anesthetics and xenon share cardioprotective and neuroprotective effects involv-ing these signaling pathways.109

The effects of anesthetics on the phosphorylation of individual residues in specific substrates can be studied using phosphorylation state–specific antibodies that are able to detect the phosphorylated forms of kinase sub-strates. A comparison of the effects of three mechanis-tically diverse anesthetics (isoflurane, propofol, and ketamine) on critical intracellular protein phosphory-lation signaling pathways that are known to integrate multiple second messenger systems reveals both shared and agent-specific actions in vivo.182 All three anesthet-ics reduce phosphorylation of activating sites on NMDA and AMPA glutamate receptors and of the downstream extracellular signal-regulated kinase ERK2, which are known to be involved in synaptic plasticity, consistent with depression of normal glutamatergic synaptic trans-mission in the anesthetized mouse cerebral cortex. These effects are somewhat selective in that several other PKA substrates examined are not affected, indicating sub-strate-specific effects rather than a general inhibition of PKA activity.183 Additional studies will be required to determine which anesthetic effects on kinase pathways represent direct effects, as occurs with PKC, and which are indirect because of anesthetic-induced alterations in signaling molecules known to regulate protein kinase and phosphatase activity such as Ca2+ and other second messengers.

Gene ExpressionThe ability of general anesthetics to alter gene expres-sion in the brain was first observed for the highly reac-tive immediate early genes c-fos and c-jun.184 Anesthetic effects on gene expression have since been observed for multiple anesthetics and organs.185 In the hippocampus of aged rats, changes in gene expression persisted for up to 2 days in rats exposed to isoflurane and nitrous oxide,186 and changes in protein expression have been observed 3 days after exposure to desflurane.187 The significance of these changes in gene and protein expression persisting after recovery from the classic signs of anesthesia remains to be established (see review77).

CELLULAR MECHANISMS

NEURONAL EXCITABILITY

Neuronal excitability is determined by resting membrane potential, threshold for action potential initiation, and input resistance (an indication of overall channel activ-ity), which can differ between compartments within an individual neuron owing to subcellular specializations (e.g., soma versus dendrite). Neurons show considerable diversity, and therefore anesthetic effects vary between neuronal populations with the state of the individual neuron and its network—that is, whether it is hyperpo-larized or depolarized, stimulated by synaptic inputs, or quiescent. Therefore it should be recognized that results obtained even from traditional acute slice preparations

Downloaded from ClinicalKey.com at Buddhist Tzu For personal use only. No other uses without permission.

hapter 25: Inhaled Anesthetics: Mechanisms of Action 629

only incompletely reflect anesthetic effects on neurons in vivo.

The intrinsic excitability of lumbar spinal motoneu-rons in vivo appears to be little affected by halothane.188 Anesthetic effects on the firing properties of hippocampal pyramidal neurons are complex: increased or decreased thresholds and regional differences and dose-dependent effects on firing patterns have been reported.189,190 By contrast, neurons in the ventral-posterior nucleus of the thalamus (possibly thalamic relay neurons) hyper-polarize in the presence of isoflurane and are less likely to fire action potentials because of a decrease in input resistance (increased shunting), attributed to increased K+ conductance.191 Similar effects are observed in hypo-glossal motoneurons and neurons from the locus coeru-leus, where a TASK-type K2P channel has been causally implicated.192

A role for GABAA receptors located at extrasynaptic sites may include volatile anesthetic effects (Fig. 25-9). Extrasynaptic GABAA receptors, which differ from synap-tic receptors in their subunit composition, mediate tonic inhibition (as opposed to phasic inhibition mediated by synaptic GABAA receptors) and are remarkably sensitive to many general anesthetics. Extrasynaptic GABAA recep-tors have a high affinity for GABA, desensitize slowly, and are tonically exposed to low ambient GABA concentra-tions,193 at which the potentiating effect of anesthetics is most significant. The effects of anesthetics on tonic inhibi-tion have been characterized in the hippocampus, which is critically involved in learning and memory. Hippocam-pal neurons generate a robust tonic current via activation of α5-subunit–containing GABAA receptors194-197 that are highly sensitive to etomidate, propofol, midazolam, and isoflurane. GABAA receptors containing the α5-subunit are highly sensitive to low concentrations of propofol and isoflurane that produce amnesia but not unconscious-ness, and may also be involved in their longer-lasting cognitive effects (discussed earlier). (Also see Chapters 26, 30, and 99.) Receptors containing α5-subunits also con-tribute to slow phasic (synaptic) currents that have been discovered in many brain regions.198 In the hippocam-pus, these currents termed GABAAslow are substantially modulated by both etomidate and isoflurane at amnesic concentrations. The slow time course and the location of GABAAslow matches that of synaptic NMDA receptor–mediated inputs in hippocampal pyramidal cells, placing them therefore in an ideal position to modulate synaptic plasticity. Thus these receptors provide a potential sub-strate for the amnesic properties of anesthetics.

PRESYNAPTIC VERSUS POSTSYNAPTIC EFFECTS ON SYNAPTIC TRANSMISSION

General anesthetics have potent and specific effects on synaptic transmission, including presynaptic actions (by altering transmitter release) and postsynaptic actions (by altering the postsynaptic responses of neurons to specific transmitters). The relative contributions of presynaptic compared with postsynaptic anesthetic effects on synap-tic transmission have been difficult to resolve, probably because the effects are transmitter and synapse specific. The net effect of anesthetics on synaptic transmission is

Chi General Hospital JC September 17, 2016.Copyright ©2016. Elsevier Inc. All rights reserved.

Page 17: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology630

Figure 25-9. Synaptic and extrasyn-aptic GABAA receptors are targets for inhaled anesthetics. A, On binding of GABA (γ-aminobutyric acid) to the GABAA receptor, its chloride-permeable channel opens, leading to hyperpolar-ization. Volatile anesthetics have a rela-tively low potency but high efficacy at synaptic GABAA receptors and a high potency and low efficacy at extrasyn-aptic GABAA receptors. B, General anesthetics prolong channel opening and increase postsynaptic inhibition. The cartoon illustrates the prolongation of miniature inhibitory postsynaptic currents (mIPSCs) by the slowing of current decay. C, A pentameric GABAA receptor complex in the lipid bilayer membrane (left), with a blow-up of a single subunit showing the location of residues critical to anesthetic efficacy in the second and third transmembrane domains (right). D, A tonic inhibitory conductance is revealed by the appli-cation of a GABAA receptor antagonist (bicuculline or picrotoxin), as shown by the upward shift in the baseline cur-rent. Anesthetics and benzodiazepines increase the tonic conductance as indi-cated by the inward shift in the current. (Modified from Hemmings HC Jr, Akabas MH, Goldstein PA, et al: Emerging molec-ular mechanisms of general anesthetic action, Trends Pharmacol Sci 26:503-510, 2005.)

1 2 3 4

Cl-

Cl-

Cl-

βγ

βαα

A

C

B

Cur

rent

Time

GABArelease Postsynaptic

membraneCa2+

entryAction potentialInhibitory postsynaptic

current

+ Anesthetic

Channelpore

Putativeanesthetic site

Intracellular

Criticalamino acids

Extracellular

Bicuculline, picrotoxinInhibition of tonic current

Potentiation of tonic current

Baseline (Iholding)

Anesthetics, benzodiazepines

D

Isoflurane 2.5 µM Isoflurane 25 µM

determined by the relative magnitude and direction of both their presynaptic and postsynaptic effects.

Synaptic excitation is decreased by volatile anesthet-ics (Fig. 25-10). Experiments in various slice prepara-tions indicate that reduced excitation is primarily caused by presynaptic mechanisms.131,188,199-202 A postsynap-tic mechanism is also involved because the response to directly applied glutamate is reduced to some degree.202-204 Volatile anesthetics have inconsistent effects on cloned AMPA or NMDA glutamate receptors, but they potenti-ate kainite receptors,117,205-207 which is consistent with a predominantly presynaptic mechanism for inhibition of glutamatergic synapses. By contrast, the effects of the nonhalogenated inhaled anesthetics (xenon, nitrous oxide, cyclopropane) appear to be mediated primar-ily by inhibition of postsynaptic NMDA receptors (dis-cussed earlier). Augmentation of GABAergic inhibition by most general anesthetics is mediated by both presyn-aptic and postsynaptic mechanisms. Enhancement of

Downloaded from ClinicalKey.com at BuddhFor personal use only. No other uses without perm

postsynaptic and extrasynaptic GABAA receptors is well recognized.118 Remarkably, however, volatile anesthet-ics also increase spontaneous GABA release and inhibi-tory postsynaptic current (IPSC) frequency208-212—that is, their presynaptic effects at GABAergic terminals are dis-tinct from those at glutamatergic synapses.

The mechanisms for the presynaptic effects of inhaled anesthetics, like those for their postsynaptic effects, are complex and involve multiple targets. Although a syn-apse-specific contribution of presynaptic Ca2+ channels is likely, presynaptic Na+ channels are more sensitive than the Ca2+ channels coupled to glutamate release. This is consistent with observations that the predominant Ca2+ channel coupled to neurotransmitter release at hippo-campal glutamatergic synapses (P/Q-type) is insensitive to isoflurane.149 Other presynaptic mechanisms have been proposed, including actions on the vesicle fusion process as demonstrated in the model organism Cae-norhabditis elegans.213,214 However, isoflurane effects on

ist Tzu Chi General Hospital JC September 17, 2016.ission. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 18: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

Chapter 25: Inhaled Anesthetics: Mechanisms of Action 631

Figure 25-10. Halogenated anesthetics enhance inhibitory and depress excitatory synaptic transmission. Halothane slows the decay of GABAA receptor–mediated postsynaptic inhibitory currents (IPSCs) (A) and depresses the amplitude of glutamatergic excitatory postsynaptic currents (EPSCs) without affecting the decay (B) in hippocampal interneurons. (A, Redrawn with permission from Nishikawa K, MacIver MB: Membrane and synaptic actions of halothane on rat hippocampal pyramidal neurons and inhibitory interneurons, J Neurosci 20:5915-5923, 2000. B, Redrawn from Perouansky M, et al: Effects of halothane on glutamate receptor-mediated excitatory postsynaptic currents: a patch-clamp study in adult mouse hippo-campal slices, Anesthesiology 83:109-119, 1995.)

A B

Interneuron IPSC

ControlGABAA

Halothane0.35 mM Overlay

200 pA

100 msec

100 pA

200 msec

100 pA

20 msec

Halothane0.64 mM

Controlnon-NMDA

Interneuron EPSC

NMDA

C

exocytosis in rat hippocampal neurons occur primarily upstream of vesicle fusion.137,215 (See review117.)

The general effects of inhaled anesthetics are to increase inhibitory synaptic transmission and to inhibit excitatory synaptic transmission (Fig. 25-11). The rela-tive importance of anesthetic effects on excitatory versus inhibitory synaptic transmission, and the mechanisms of these effects, is likely to differ for different anesthetics and for specific synapses and networks.105,216 This con-cept extends to other putative targets and mechanisms and is incorporated in an “agent-specific multiple-site” hypothesis of anesthesia.216,217

SIMPLE CIRCUITS AND COMPLEX NETWORKS

SIMPLE CIRCUIT PHENOMENA

The development of a mechanistic understanding of phenomena involving complex circuits has been greatly facilitated by the study of anatomically reduced (in vitro) or physiologically simplified (in vivo) preparations, com-plemented by computer simulations (in silico). These approaches are essential for integrating reductionist observations of the multiple molecular effects of anes-thetics into functional models relevant to behavioral end points. Anesthetic effects have been studied in acute brain slice preparations from various regions of the CNS (hippocampus, amygdala, cortex, thalamus, brainstem, and spinal cord—most often from rodents). Brain slices preserve native connections but usually lack natural inputs and outputs. Slices from developing mammalian brain can be cultured in vitro. These “organotypic slice cultures” preserve a high degree of synaptic connectiv-ity and display spontaneous network activity, typically absent from “acute slices.” Simplified in vivo preparations involve phenomena (typically evoked responses) with relatively well-understood circuitry. Computer models and simulations can assist in generating hypotheses for

Downloaded from ClinicalKey.com at Buddhist Tzu For personal use only. No other uses without permission. C

experimental testing and developing hypotheses based on experimental data.

Synaptic PlasticityPaired-pulse depression and paired-pulse facilitation are examples of short-term plasticity in response to external stimulation. Synaptic inhibition is prolonged by volatile anesthetics in vivo218 and in vitro,219 in general agreement with the notion that anesthetics enhance functional inhi-bition in the CNS. Enhancement of paired-pulse facilita-tion has been attributed to the presynaptic depressant effect of volatile agents (see Fig. 25-11).131,201

Long-term potentiation, a cellular model of learning and memory, consists of use-dependent strengthening of glutamatergic excitatory synaptic connections (see

5 mV

100 msec

Control

Halothane

Halothane

Pyramidal cell

ControlInterneuron

Figure 25-11. Anesthetics affect excitation and inhibition simulta-neously in opposite directions. Halothane depresses the excitatory depolarization and augments the inhibitory hyperpolarization in both hippocampal pyramidal cells and interneurons. The net result will depend on the underlying neuronal state, the neuronal network, and its function. (Redrawn form Nishikawa K, MacIver MB: Membrane and synaptic actions of halothane on rat hippocampal pyramidal neurons and inhibitory interneurons, J Neurosci 20:5915-5923, 2000.)

hi General Hospital JC September 17, 2016.opyright ©2016. Elsevier Inc. All rights reserved.

Page 19: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology632

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.2

2.4

2.0

-20 -10 0 10 20 30 40 -20 -10 0 10 20 30 40

Nor

mal

ized

fEP

SP

slo

pe100Hz/1sec 100Hz/1sec

Isoflurane 0.2–0.3 mM

Time (min)

Figure 25-12. Isoflurane blocks the induction of synaptic plasticity in vitro (a model of learning and memory). Long-term potentiation (an increase in synaptic strength), indicated by the increase in normalized excitatory postsynaptic potential (EPSP) slope, in hippocampal slices evoked by tetanic stimulation of excitatory synapses, is blocked by 0.2 to 0.3 mM isoflurane. (Redrawn from Simon W, et al: Isoflurane blocks synaptic plastic-ity in the mouse hippocampus, Anesthesiology 94:1058-1065, 2001.)

Chapter 13). Reported volatile anesthetic effects on LTP have been inconsistent: halothane, enflurane, and isoflu-rane do not block LTP induction in vivo, whereas ket-amine does.218 By contrast, isoflurane blocks LTP in the hippocampal slice by enhancing GABAA receptor–medi-ated inhibition (Fig. 25-12).220 Long-term depression, a use-dependent weakening of excitatory connections that is effectively a homeostatic counterpart of LTP, is also blocked by isoflurane.220 The discrepancy between these findings in vivo and in vitro remains unexplained.

Spontaneously Active CircuitsSpontaneous neuronal activity is decreased by volatile anesthetics both in vivo and in cortical brain slices. This effect is largely GABAA receptor dependent and marked even at low sedative concentrations.76 Because the corti-cal brain slices lack subcortical input, these results sug-gest that volatile anesthetics can cause some effects (e.g., sedation) via direct cortical action. However, changes in raw neuronal firing rates do not provide a precise quan-titative measure of higher cognitive function, which is better reflected by the relationship of firing patterns to strength and phase of cortical rhythms (see next sec-tion). Anesthetic effects have also been tested on the circuitry underlying locomotion, a well-studied central pattern generator. Effects of isoflurane on in vitro lam-prey and rat spinal cord preparations support the spinal cord as the primary target for volatile anesthetic–induced immobility.221,222

RHYTHMS AND SIMULATIONS

The brain perpetually generates complex electrical rhythms (oscillations in extracellular field potentials) that range in frequency from fractions to hundreds of Hertz (Hz; cycles per second), as recorded on the surface of the scalp as the electroencephalogram (EEG; the higher fre-quency oscillations cannot be resolved in surface record-ings). All oscillations are behavioral state dependent, and

Downloaded from ClinicalKey.com at BuddhFor personal use only. No other uses without permi

multiple oscillations coexist throughout the sleep-wake cycle. Lower frequency rhythms allow for integration over longer time periods and typically engage larger areas of the brain. By contrast, higher frequency rhythms allow for higher temporal resolution on local scales. Cross- frequency modulation allows for integration of both aspects of information processing. Although their physi-ologic roles are not clear, brain rhythms reflect, subserve, and/or constitute fundamental higher-order processing, such that their modulation by anesthetics is of great interest. The current nomenclature of brain rhythms reflects historical conventions and is not based on under-lying mechanisms.

δ-Rhythms and Other Slow RhythmsOscillations with EEG frequencies from 1.5 to 4 Hz are gen-erally referred to as δ-rhythms. Slower frequency rhythms are termed Slow-1 to Slow-4, with the higher numbers referring to slower rhythms.223 Slow rhythms are promi-nent during non–rapid eye movement sleep (i.e., slow-wave sleep) and appear at loss of consciousness induced by propofol.224 δ-Frequency rhythms are commonly observed under general anesthesia. During natural slow-wave sleep (SWS), δ-rhythms and sleep spindles are phase related to a slower oscillation, suggesting functional interaction.225 Paroxysmal spindlelike waxing and waning oscillations overriding slower rhythms are also present in the cortical EEG under anesthesia, but the underlying physiology and functional significance are unknown.

θ-Rhythmsθ-Rhythms, present in various cortical structures but most prominent in the hippocampus, are thought to signal the “online state.” They are associated with sensorimotor and mnemonic functions during waking behavior.226 One component of the θ-rhythm (type I or atropine resis-tant) can be affected by amnestic concentrations of iso-flurane as well as by the amnestic nonimmobilizer F6,63 indicating a potential network-level signature effect for

ist Tzu Chi General Hospital JC September 17, 2016.ssion. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 20: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

anesthetic-induced amnesia. Type II θ-rhythm (atropine sensitive) can be evoked under anesthesia and is slowed and potentiated by halothane.227

γ-RhythmsThis designation includes an extremely broad and proba-bly functionally and mechanistically heterogeneous spec-trum of rhythms. It is frequently subdivided into slow γ (30 to 50 Hz; i.e., the spectrum above β-rhythms); γ (50 to 90 Hz), and fast-, ultra-γ, or ε-rhythms (>90 and up to hun-dreds of Hz).224 Fast GABAA-ergic synaptic inhibition and intrinsic resonant properties of neurons play important roles in γ-physiology, making these obvious candidates for anesthetic modulation. Isoflurane slows the frequency of evoked γ-oscillations (30 to 90 Hz, also known as “40- Hz rhythms”) in humans.228,229 Studies of γ-oscillations in vitro suggest that their frequency depends on the time constant of decay of GABAA receptor–mediated synap-tic currents in inhibitory networks.230 Isoflurane slows γ-rhythms in hippocampal231 and neocortical slices232 to a comparable degree as it does in humans,229 providing a tentative link between receptor and circuit-level effects. However, the interaction between anesthetics and behav-iorally relevant network effects is likely to be complex because flash-evoked γ-oscillations in the primary visual cortex are not affected by inhaled agents,54 whereas feed-back information transfer at γ-frequencies between the visual and frontal cortices is disrupted.53,54 Moreover, brain rhythms are interlinked (e.g., θ-rhythms modulate γ-oscillations (θ−γ nesting). The nature of their modula-tion by anesthetics as well as relevance is far from clear.

Models and SimulationsOn the atomic scale, modeling the interaction of anes-thetic molecules with targets and model proteins has defined binding motifs that characterize the amphiphilic binding cavities for anesthetic molecules (see “Identifica-tion of Molecular Sites of Anesthetic Action,” discussed earlier).115 Modeling anesthetic interaction with firefly luciferase suggests that anesthetics act by disrupting the modes of motion essential for protein function.233

On the macroscopic scale, computer simulations can provide an integrated picture of modulation of dynamic neuronal and network activity. “Bottom-up” neuron-by-neuron approaches are based on computational models of individual neurons, known anesthetic effects on intrin-sic and synaptic membrane conductances, and simple network models. Computer simulations of anesthetic effects on integrated outputs can thereby be generated (e.g., the firing of pacemaker neurons).234 These models obviously rely on the accuracy of the derived characteris-tics of real neurons and networks as well as knowledge of the effects of anesthetics; the scale of simulations is lim-ited by the complexity of its elements.235 An alternative is a “top-down” approach, such as mean-field modeling, in which molecular and cellular individual accuracy detail is sacrificed in favor of global dynamics. Global cortical phenomena, such as anesthetic-induced seizures,236 are modeled as phase transitions based on mean interactions between populations of averaged neurons (analogous to the EEG signal that also averages the signals of neuro-nal populations). This approach can be extended to other

Downloaded from ClinicalKey.com at Buddhist TzFor personal use only. No other uses without permission

Chapter 25: Inhaled Anesthetics: Mechanisms of Action 633

global cortical phenomena such as consciousness.237 Neu-ronal modeling and computer simulations will likely gain importance in the future as bridges between theoretical and experimental studies of anesthesia.

RESEARCH STRATEGIES FOR THE FUTURE

The search for anesthetic mechanisms is driven by advances in the basic sciences. Some strategies that should facilitate understanding of anesthetic mechanisms include use of agonists or antagonists in vivo, nonanesthetics or nonimmobilizers, transgenic animals, and high-resolution imaging of the functioning brain.

PHARMACOLOGIC APPROACHES

Agonists, Antagonists, and Experimental AnestheticsUse of receptor-specific agonists and antagonists provides a pharmacologic method for bridging in vitro with in vivo studies. In this approach, a receptor may be tested for its contribution to a specific end point (e.g., immobility) according to the criteria presented earlier. This approach was used to exclude an important role for NMDA receptor block in the immobilizing action of volatile anesthetics but did not yield conclusive results for the roles of GABAA and GlyR for immobilization,238,239 probably owing to the intricacies of drug-receptor interactions on different levels of integration in complex networks like the spinal cord. A complementary pharmacologic approach using experimental anesthetics that inhibit NMDA receptors in vitro with different potencies supports the conclu-sion that NMDA receptor blockade does not contribute significantly to immobility by conventional volatile anesthetics.240 Refinement of this strategy involves the anatomically discrete application of drugs to nuclei with known function. For example, the tuberomammillary nucleus (part of the endogenous sleep pathway) medi-ates the sedative component of anesthesia for some intravenous anesthetics (e.g., propofol).23 A discrete site of general anesthetic action for GABAergic drugs in the mesopontine tegmentum has also been proposed based on this strategy.22,241

NonimmobilizersNonimmobilizers are compounds with physicochemi-cal characteristics similar to those of conventional inhaled anesthetics, but their predicted anesthetizing concentrations (based on their lipid solubility and the Meyer-Overton correlation [MACpred]) do not induce immobility.11 Initially termed nonanesthetics, the termi-nology was revised when it was discovered that at least some of them cause amnesia at similar MACpred fractions as classic volatile anesthetics.67 If a molecular or cellular process is affected in similar ways by an anesthetic and a nonimmobilizer, that process is not relevant for the anesthetic state, with the notable exception of amne-sia. Despite this elegant rationale, only a limited num-ber of receptors have been excluded because, compared with volatile anesthetics, nonimmobilizers are relatively

u Chi General Hospital JC September 17, 2016.. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 21: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology634

target selective. These compounds have the potential to provide insights beyond the initially envisaged receptor-level studies by allowing the separation of sedation from amnesia for the study of underlying network activity in vivo.63

Photoreactive AnestheticsRecently, azi-isoflurane, a photoreactive analogue of isoflu-rane, was synthesized. This agent has similar (although not identical) physicochemical properties as isoflurane. When irradiated at a wavelength of 300 nm, azi-isoflurane reacts with amino acid residues surrounding putative isoflurane-binding sites. Use of azi-isoflurane and similar agents could help identify binding sites of inhaled agents on anesthesia-relevant molecular targets.242

GENETIC APPROACHES

Genetic strategies take two forms: forward and reverse.243 The reverse genetic approach focuses on a particular gene, chosen because there are reasons to hypothesize that its product may be important to anesthesia. Exam-ples of this strategy are targeted mutations that alter the sensitivity of specific neurotransmitter receptors to anesthetics.121 Initially, these mutations were used to identify anesthetic binding sites. Subsequently, trans-genic animals rendered resistant to anesthetics, either by deletion of a putative target protein from the genome or by expressing a target receptor engineered to be insen-sitive to an anesthetic, were used to test behavioral rel-evance of the altered gene product for the production of anesthesia. Forward genetics, by contrast, is a discovery process that involves the study of randomly generated mutations (either experimentally induced or naturally occurring polymorphisms) in a population that affect the phenotype of interest (i.e., anesthetic end points).

Knockout and Knock-in AnimalsIn the knockout approach, expression of the gene encod-ing a protein of interest is disrupted by a specific deletion or insertion. Nearly all such studies have been carried out in mice. A well-recognized problem with the global knockout approach is that extensive compensatory changes can be induced, from anomalies that are lethal in utero to insidious (but experimentally confounding) influences that might be expressed only at maturity. A complementary strategy is the conditional knockout, in which the genetic deletion is restricted—either anatomi-cally (limited to certain brain regions) or temporally (at a known point in time). These strategies can minimize developmental anomalies and reduce the likelihood of compensatory changes. In the knock-in approach, a muta-tion, usually of a single amino acid residue, is targeted to produce a protein with altered sensitivity to a drug of interest. Ideally, this mutation remains completely silent in the absence of the drug—that is, it does not perturb the normal expression and function of the protein of interest or alter the expression of other genes.

GABAA ReceptoRs. Results from transgenic animals illus-trate both the utility and the difficulties of the genetic approach with respect to inhaled agents. The conditional

Downloaded from ClinicalKey.com at BuddhFor personal use only. No other uses without permi

forebrain-restricted GABAA receptor α1-subunit knock-out mouse was found to be less sensitive to isoflurane-induced amnesia than wild-type mice, and this led to the conclusion that action at these receptors contributes to its amnesic effects.244 By contrast, a mouse harboring a muta-tion of the GABAA receptor α1-subunit that renders the receptor insensitive to isoflurane in vitro did not show reduced sensitivity to either the amnesic or the immobi-lizing effects of isoflurane, leading to the conclusion that this subunit does not mediate the impairment of learning and memory by isoflurane.31 Similar experiments indi-cate that action at the GABAA receptor β3-subunit does not mediate immobility or amnesia by isoflurane.30 This “bottom-up” genetic approach is a work-intensive but powerful tool that has yielded clear results with the recep-tor-specific intravenous anesthetics,16 but it has proved more challenging to apply it to the more promiscuous inhaled agents.

Glycine α1–contAininG ReceptoRs. Pharmacologic stud-ies supported the notion that glycinergic neurotransmis-sion might be the effector for the immobilizing action of inhaled anesthetics in the spinal cord, where glycine replaces GABA as the principal inhibitory transmit-ter. However, mice harboring mutations that render α1 subunit–containing glycine receptors largely insensitive to alcohol and inhaled ether anesthetics did not demon-strate a concordant change in MAC values. Because α1 is the most widely expressed subunit in adult animals, it is unlikely that action at glycine receptors plays an impor-tant part in the immobilizing action of inhaled agents.245

two-poRe DomAin K chAnnels. Use of mice harboring knockout mutations of several two-pore domain K+ chan-nel (K2P) family members (TASK-1, TASK-3, TREK-1) has demonstrated a role for these channels in volatile anesthe-sia.32-34 For example, TREK-1 knockout mice are partially resistant to all volatile anesthetics tested with respect to both loss of righting reflex (a measure of conscious-ness) and immobility, but anesthesia can still be induced, albeit at higher anesthetic concentrations. Interestingly, responses to pentobarbital are unaffected, indicating that the mutation does not cause a generalized resistance to anesthesia.

FORWARD AND POPULATION GENETICS

The nematode Caenorhabditis elegans and the fruit fly Drosophila melanogaster, with 302 and 100,000 neu-rons, respectively, have also been used as model organ-isms in anesthesia research. Mutations in a number of C. elegans genes affect sensitivity to volatile anesthetics, such as unc64,214 stomatin,246 and gas1. Mutation of the latter gene confers hypersensitivity to enflurane. This gene encodes the worm homologue of a subunit of NADH hydrogenase expressed in neurons.247 Yeasts have also been used as model organisms, with even more obvious limitations with respect to identification of appropriate anesthetic end points.

Sensitivity to anesthetics is a quantitative trait (vary-ing continuously in a population). Quantitative genetics is the study of the heritability of continuous traits. These

ist Tzu Chi General Hospital JC September 17, 2016.ssion. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 22: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

C

traits are controlled by genes represented in quantitative trait loci (QTLs). A top-down population-based approach has been used to localize QTLs that govern susceptibil-ity of individuals to anesthetics, in higher and lower organisms. Starting from the observation that inbred mouse strains vary in their sensitivity to isoflurane, mic-rosatellite-based linkage analysis and, alternatively, sin-gle-nucleotide polymorphism–based analysis of genetic variation localized a QTL for isoflurane immobilization to the proximal part of mouse chromosome 7.248 This type of analysis promises to assist in defining the genetic basis for variability in the susceptibility to both primary anesthetic end points and side effects.

HIGH-DENSITY EEG AND FUNCTIONAL IMAGING

Identification of the anatomic and functional substrates for anesthetic effects on consciousness, memory, and immobility is now approachable with improved imaging techniques. Imaging is based on mapping of either meta-bolic or hemodynamic changes as surrogate measures of neuronal activity, as in positron emission tomography (PET) and functional magnetic resonance imaging, or on mapping electrical activity with high-density EEG, mag-netoencephalography, and low-resolution brain electro-magnetic tomography (see also Chapters 15, 17, and 49). Properties of receptors can also be probed with radioac-tive ligands (PET). These techniques have the capacity to identify neuroanatomic substrates of drug action, with method-specific limitations. Results from functional PET imaging suggest that propofol suppresses episodic memory by targeting the prefrontal and posterior pari-etal cortex as opposed to the medial temporal lobe,249 and that suppression of consciousness is caused by anesthetic action in the thalamus, parts of the medial and posterior parietal cortex, and/or the posterior cingulate and medial parietal cortex.250 Although observations of regionally specific and global suppressive effects of anesthetics on metabolic activity are unlikely to provide a definitive mechanistic understanding, such information can facili-tate hypotheses and experimentally testable predictions.

Advanced analytical approaches based on theories from mathematical and statistical sciences are being increas-ingly applied to enhance the power of existing technolo-gies. Magnetic resonance images and high-density EEG recordings of the brain reveal strong interregional connec-tions, but the considerable potential of this connectivity information to better understand the brain’s response to anesthesia has only recently begun to be tapped. The grow-ing use of invasive recording techniques (e.g., brain surface electrode grids and microelectrodes implanted deep into the brain for functional neurosurgical therapies) is also advancing the frontiers of neuroscience in general and the understanding of anesthetic mechanisms in particular.

SUMMARY

The mechanisms of inhaled anesthetics have proved more difficult to explain than was envisaged a generation ago, when the paradigm shifted from lipids to amphiphilic

Downloaded from ClinicalKey.com at Buddhist Tzu For personal use only. No other uses without permission. C

hapter 25: Inhaled Anesthetics: Mechanisms of Action 635

cavities in proteins as targets for anesthetics. Despite a remarkable accumulation of information, a comprehen-sive theory of general anesthetic action has yet to be for-mulated. Progress toward this goal has been difficult for several reasons. Important pharmacologic characteristics of inhaled anesthetics that have impeded identification of their relevant molecular targets are their low potency (mil-limolar), promiscuous activity at multiple targets, lack of specific antagonists, and limitations in the neuroscience of memory and consciousness. This contrasts with the situa-tion for intravenous anesthetics that exhibit more conven-tional receptor pharmacology. Moreover, accumulating evidence indicates that no universal target exists to explain the actions of every general anesthetic, or even of a single anesthetic agent. It is now clear that the composite state of anesthesia and its core components (amnesia, sedation/unconsciousness, immobility) are separable behavioral states in vivo that have limited reproducibility in vitro. Resolution of these phenomena at the molecular and cel-lular levels represents the cutting edge of contemporary neuroscience. Of the multiple molecular and cellular anes-thetic effects identified, it is unclear which are critical for the desired behavioral end points, which are harmless or beneficial side effects (e.g., preconditioning), and which, if any, could have long-lasting or delayed undesirable conse-quences (e.g., cell death, cognitive dysfunction). Progress in identifying the molecular targets of general anesthet-ics provides a foundation for identification of the network and systems level effects relevant to their behavioral and peripheral end points. As the biologic foundations of behaviors once thought to be exclusively the realm of psychology are unraveled, for which anesthetics provide a valuable investigative tool, a comprehensive theory of anesthesia will also develop.

Complete references available online at expertconsult.com.

RefeRences

1. Perouansky M: Anesthesiology 117:465, 2012. 2. Eger EI: Anesthesiology 96:238, 2002. 3. Eger EI, et al: Anesthesiology 26:756, 1965. 4. Franks NP, Lieb WR: Anesthesiology 84:716, 1996. 5. Seeman P: Experientia 30:759, 1974. 6. Franks NP, Lieb WR: J Mol Biol 133:469, 1979. 7. Featherstone RM, et al: Anesthesiology 22:977, 1961. 8. Ueda I, Kamaya H: Anesthesiology 38:425, 1973. 9. Franks NP, Lieb WR: Anesthesiology 101:235, 2004. 10. Franks NP, Lieb WR: Nature 310:599, 1984. 11. Koblin DD, et al: Anesth Analg 79:1043, 1994. 12. Hall AC, et al: Br J Pharmacol 112:906, 1994. 13. Eger EI 2nd, et al: Anesthesiology 94:915, 2001. 14. Eckenhoff RG: Mol Interv 5:258, 2001. 15. Eckenhoff RG, Johansson JS: Anesthesiology 95:1537, 2001. 16. Rudolph U, Antkowiak B: Nat Rev Neurosci 5:709, 2004. 17. Antognini JF, Schwartz K: Anesthesiology 79:1244, 1993. 18. Rampil IJ, et al: Anesthesiology 78:707, 1993. 19. Stabernack C, et al: Anesth Analg 100:128, 2005. 20. Veselis RA, et al: Anesthesiology 95:896, 2001. 21. Sanders R, et al: Anesthesiology 116:946, 2012. 22. Devor M, Zalkind V: Pain 94:101, 2001. 23. Nelson LE, et al: Nat Neurosci 5:979, 2002. 24. Rampil IJ, Laster MJ: Anesthesiology 77:920, 1992. 25. Antognini JF, et al: Anesthesiology 96:980, 2002. 26. Zhang Y, et al: Anesth Analg 99:85, 2004. 27. Raines DE, et al: Anesth Analg 95:573, 2002. 28. Zhang Y, et al: Br J Pharmacol 159:872, 2010.

Chi General Hospital JC September 17, 2016.opyright ©2016. Elsevier Inc. All rights reserved.

Page 23: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

PART III: Anesthetic Pharmacology636

29. Liao M, et al: Anesth Analg 101:412, 2005. 30. Sonner JM, et al: Anesthesiology 106:107, 2007. 31. Heurteaux C, et al: EMBO J 23:2684, 2004. 32. Linden AM, et al: J Pharmacol Exp Ther 317:615, 2006. 33. Linden AM, et al: J Pharmacol Exp Ther 323:924, 2007. 34. Westphalen RI, et al: Br J Pharmacol 152:939, 2007. 35. Jinks SL, et al: Anesthesiology 108:1016, 2008. 36. Searle JR: Annu Rev Neurosci 23:557, 2000. 37. Tononi G: Biol Bull 215:216, 2008. 38. Crick F, Koch C: Nat Neurosci 6:119, 2003. 39. Alkire MT, et al: Conscious Cogn 9:370, 2000. 40. Ries CR, Puil E: J Neurophysiol 81:1795, 1999. 41. Detsch O, et al: Brain Res 829:77, 1999. 42. John ER, et al: Conscious Cogn 10:165, 2001. 43. Chortkoff BS, et al: Anesth Analg 81:737, 1995. 44. Dwyer R, et al: Anesthesiology 77:888, 1992. 45. Alkire MT, et al: Science 322:876, 2008. 46. Brown E, et al: N Engl J Med 263:2638, 2010. 47. Varela F, et al: Nat Rev Neurosci 2:229, 2001. 48. van den Heuvel MP, Sporns O: J Neurosci 31:15775, 2011. 49. Bullmore E, Sporns O: Nat Rev Neurosci 13:336, 2012. 50. Massimini M, et al: Science 309:2228, 2005. 51. Ferrarelli F, et al: Proc Natl Acad Sci U S A 107:2681, 2010. 52. Imas OA, et al: Anesthesiology 102:937, 2005. 53. Imas OA, et al: Neurosci Lett 402:216, 2006. 54. Ter Mikaelian M, et al: J Neurosci 27:6091, 2007. 55. Burlingame RH, et al: Anesthesiology 106:754, 2007. 56. Dutton RC, et al: Anesthesiology 94:514, 2001. 57. Alkire MT, Nathan SV: Anesthesiology 102:754, 2005. 58. Alkire MT, Gorski LA: Anesthesiology 101:417, 2004. 59. Vertes RP: Hippocampus 15:923, 2005. 60. Pan WX, McNaughton N: Brain Res 764:101, 1997. 61. Robbe D, et al: Nat Neurosci 9:1526, 2006. 62. Perouansky M, et al: Anesthesiology 106:1168, 2007. 63. Seidenbecher T, et al: Science 301:846, 2003. 64. Pryor KO, et al: Anesthesiology 113:313, 2010. 65. Rudolph U, et al: Nature 401:796, 1999. 66. Kandel L, et al: Anesth Analg 82:321, 1996. 67. Mihic SJ, et al: Mol Pharmacol 46:851, 1994. 68. Zarnowska ED, et al: Anesth Analg 101:401, 2005. 69. Jevtovic-Todorovic V, et al: Nat Med 4:460, 1998. 70. Gruss M, et al: Mol Pharmacol 65:443, 2004. 71. Gries DA, et al: Life Sci 76:1667, 2005. 72. Murphy M: Sleep 34:283, 2011. 73. Tung A, et al: Anesthesiology 100:1419, 2004. 74. Nelson A, et al: Sleep 33:1659, 2010. 75. Hentschke H, et al: Eur J Neurosci 21:93, 2005. 76. Perouansky M, Hemmings HC Jr: Anesthesiology 111:1365, 2009. 77. Durieux M, Davis PJ: Anesth Analg 110:1265, 2010. 78. Jevtovic-Todorovic V, et al: J Neurosci 23:876, 2003. 79. Shih J, et al: Anesthesiology 116:586, 2012. 80. Kitano H, et al: J Cereb Blood Flow Metab 27:1108, 2006. 81. Fukuda S, Warner DS: Br J Anaesth 99:10, 2007. 82. Kawaguchi M, et al: J Anesth 19:150, 2005. 83. Inoue S, et al: Anesthesiology 101:75, 2004. 84. Preckel B, et al: Anesthesiology 105:187, 2006. 85. Turner CP, et al: Neuroscience 17:384, 2012. 86. Kitano H, et al: J Cereb Blood Flow Metab 27(6):1108, 2007. 87. Rasmussen L, et al: Acta Anesthesiol Scand 48:1137, 2004. 88. Zurek A, et al: Anesth Analg 114:845, 2012. 89. Wiklund A, et al: Anesthesiology 109:790, 2008. 90. Wan Y, et al: Anesthesiology 106:436, 2007. 91. Terrando N, et al: Ann Neurol 70:986, 2011. 92. Tang J, et al: Alzheimer’s Dement 7:521, 2011. 93. Avidan M, et al: Anesthesiology 111:964, 2009. 94. Avidan M, Evers A: J Alzheimer Dis 24(2):201, 2011. 95. Zaugg M, et al: Br J Anaesth 91:551, 2003. 96. Pagel PS, Warltier DC: Ventricular function. In Warltier DC, editor:

Anesthetics and left ventricular function. Baltimore, 1995, Williams & Wilkins, p 213.

97. Hanley PJ, et al: Anesthesiology 101:999, 2004. 98. Huneke R, et al: Anesthesiology 95:999, 2001. 99. Stowe DF, et al: Anesthesiology 92:516, 2000. 100. Ebert TJ, Kampine JP: Anesth Analg 69:444, 1989. 101. Stowe DF, et al: Anesthesiology 73:1220, 1990. 102. Pagel PS, et al: Anesthesiology 74:1103, 1991.

Downloaded from ClinicalKey.com at BuddhisFor personal use only. No other uses without permis

103. Tanaka K, et al: Anesthesiology 106:984, 2007. 104. Yoshino J, et al: Naunyn Schmied Arch Pharmacol 371:500, 2005. 105. Vulliemoz Y: Toxicol Lett 100-101:103, 1998. 106. Huneke R, et al: Acta Anaesthesiol Scand 48:547, 2004. 107. Stadnicka A, et al: J Anesth 21:212, 2007. 108. Suleiman MS, et al: Br J Pharmacol 153:21, 2008. 109. Pratt PF Jr, et al: Curr Opin Anaesth 19:397, 2006. 110. Weber NC, et al: Anesthesiology 103:1174, 2005. 111. Tanaka K, et al: Anesthesiology 100:707, 2004. 112. Stuth EAE, et al: Anesthesiology 91:804, 1999. 113. Franks NP, Lieb WR: Nature 367:607, 1994. 114. Zeller A, et al: Handb Exp Pharmacol 182:31, 2008. 115. Bertaccini EJ, et al: Anesth Analg 104:318, 2007. 116. Nury H, et al: Nature 469(7330):428, 2011. 117. Dickinson R, et al: Anesthesiology 107:756, 2007. 118. Hemmings HC Jr, et al: Trends Pharmacol Sci 26:503, 2005. 119. Patel AJ, et al: Nat Neurosci 2:422, 1999. 120. Lynch JW: Physiol Rev 84:1051, 2004. 121. Mihic SJ, et al: Nature 389:385, 1997. 122. Solt K, et al: J Pharmacol Exp Ther 315:771, 2005. 123. Role LW, Berg DK: Neuron 16:1077, 1996. 124. Flood P, et al: Anesthesiology 86:859, 1997. 125. Violet JM, et al: Anesthesiology 86:866, 1997. 126. Dingledine R, et al: Pharmacol Rev 51:7, 1999. 127. Franks NP, et al: Nature 396:324, 1998. 128. Solt K, et al: Anesth Analg 102:1407, 2006. 129. Harris RA, et al: FASEB J 9:1454, 1995. 130. Sonner JM, et al: Anesth Analg 101:143, 2005. 131. MacIver MB, et al: Anesthesiology 85:823, 1996. 132. Perouansky M, et al: Anesthesiology 100:470, 2004. 133. Winegar BD, MacIver MB: BMC Neurosci 7:5, 2006. 134. Haydon DA, Urban BW: J Physiol (Lond) 341:429, 1983. 135. Berg-Johnsen J, Langmoen IA: Acta Physiol Scand 127:87, 1986. 136. Mikulec AA, et al: Brain Res 796:231, 1998. 137. Wu XS, et al: Anesthesiology 100:663, 2004. 138. Yu FH, Catterall WA: Sci STKE 2004(253):re15, 2004. 139. Herold KF, et al: Anesthesiology 111:591, 2009. 140. OuYang W, Hemmings HC Jr: J Pharmacol Exp Ther 312:801, 2005. 141. OuYang W, et al: J Pharmacol Exp Ther 322:1076, 2007. 142. Ouyang W, Hemmings HC Jr: Anesthesiology 107:91, 2007. 143. Ratnakumari L, et al: Anesthesiology 92:529, 2000. 144. Catterall WA: Annu Rev Cell Dev Biol 16:521, 2000. 145. Topf N, et al: Actions of general anesthetics on voltage-gated

ion channels. In Antognini JF, Carlens E, Raines D, editors: Neu-ral mechanisms of anesthesia. Totowa, NJ, 2003, Humana Press, p 299.

146. Miao N, et al: Anesthesiology 83:593, 1995. 147. Kameyama K, et al: Br J Anaesth 82:402, 1999. 148. Study RE: Anesthesiology 81:104, 1994. 149. Hall AC, et al: Anesthesiology 81:117, 1994. 150. Takei T, et al: Neurosci Lett 350:41, 2003. 151. Joksovic PM, et al: Br J Pharmacol 144:59, 2005. 152. Todorovic SM, et al: Mol Pharmacol 60:603, 2001. 153. Petrenko AB, et al: Anesthesiology 106:1177, 2007. 154. Rithalia A, et al: Anesth Analg 99:1615, 2004. 155. Davies LA, et al: Anesthesiology 93:1034, 2000. 156. Pabelick CM, et al: Anesthesiology 95:207, 2001. 157. Roberts MC, et al: Anesthesiology 95:716, 2001. 158. Mickelson JR, Louis CF: Physiol Rev 76:537, 1996. 159. Yost CS: Anesthesiology 90:1186, 1999. 160. Friederich P, et al: Anesthesiology 95:954, 2001. 161. Franks NP, Lieb WR: Nature 333:662, 1988. 162. Franks NP, Honore E: Trends Pharmacol Sci 25:601, 2004. 163. Patel AJ, Honore E: Anesthesiology 95:1013, 2001. 164. Farwell D, Gollob MH: Can J Cardiol 23(Suppl A):16A, 2007. 165. Antzelevitch C: J Electrocardiol 34(Suppl):177, 2001. 166. Li J, Correa AM: Anesthesiology 97:921, 2002. 167. Davies LA, et al: Br J Pharmacol 131:223, 2000. 168. Sirois JE, et al: J Physiol 541:717, 2002. 169. Chen X, et al: J Neurosci 25:5803, 2005. 170. Robinson RB, Siegelbaum SA: Annu Rev Physiol 65:453, 2003. 171. Girault JA, Hemmings HC Jr: Cell signaling. In Hemmings HC

Jr, Hopkins PM, editors: Foundations of anesthesia, ed 2. London, 2005, Mosby, p .31.

172. Rebecchi MJ, Pentyala SN: Br J Anaesth 89:62, 2002. 173. Peterlin Z, et al: Mol Cell Neurosci 30:506, 2005.

t Tzu Chi General Hospital JC September 17, 2016.sion. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 24: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

174. Minami K, et al: Mol Pharmacol 53:148, 1998. 175. Minami K, et al: J Pharmacol Exp Ther 281:1136, 1997. 176. Minami K, et al: Eur J Pharmacol 339:237, 1997. 177. Hemmings HC Jr: Toxicol Lett 100:89, 1998. 178. Hasegawa J, et al: Acta Histochem Cytochem 39:163, 2006. 179. Das J, et al: J Biol Chem 279:37964, 2004. 180. Shumilla JA, et al: Anesth Analg 99:82, 2004. 181. Sonner JM, et al: Anesth Analg 89:1030, 1999. 182. Snyder GL, et al: Neuropharmacology 53:619, 2007. 183. Hemmings HC Jr, Adamo AI: Anesthesiology 81:147, 1994. 184. Marota JJ, et al: Anesthesiology 77:365, 1992. 185. Hamaya Y, et al: Anesth Analg 90:1177, 2000. 186. Culley DJ, et al: Eur J Pharmacol 549:71, 2006. 187. Fütterer CD, et al: Anesthesiology 100:302, 2004. 188. Kullmann DM, et al: J Physiol (Lond) 412:277, 1989. 189. Fujiwara N, et al: J Physiol (Lond) 402:155, 1988. 190. MacIver MB, Roth SH: Br J Anaesth 60:680, 1988. 191. Ries CR, Puil E: J Neurophysiol 81:1802, 1999. 192. Sirois JE, et al: J Neurosci 20:6347, 2000. 193. Semyanov A, et al: Trends Neurosci 27:262, 2004. 194. Bai D, et al: Mol Pharmacol 59:814, 2001. 195. Bieda MC, MacIver MB: J Neurophysiol 92:1658, 2004. 196. Caraiscos VB, et al: Proc Natl Acad Sci U S A 101:3662, 2004. 197. Caraiscos VB, et al: J Neurosci 24:8454, 2004. 198. Capogna M, Pearce RA: Trends Neurosci 34:101, 2011. 199. Dai S, et al: Anesthesiology 116:816, 2012. 200. Berg-Johnsen J, Langmoen IA: Acta Anaesthesiol Scand 36:350, 1992. 201. Kirson ED, et al: Br J Pharmacol 124:1607, 1998. 202. Richards CD, Smaje JC: Br J Pharmacol 58:347, 1976. 203. Wakamori M, et al: J Neurophysiol 66:2014, 1991. 204. Yang J, Zorumski CF: Ann NY Acad Sci 625:287, 1991. 205. de Sousa SLM, et al: Anesthesiology 92:1055, 2000. 206. Dildy-Mayfield JE, et al: J Pharmacol Exp Ther 276:1058, 1996. 207. Minami K, et al: J Biol Chem 273:8248, 1998. 208. Banks MI, Pearce RA: Anesthesiology 90:120, 1999. 209. Murugaiah KD, Hemmings HC Jr: Anesthesiology 89:919, 1998. 210. Nishikawa K, MacIver MB: Anesthesiology 94:340, 2001. 211. Westphalen RI, Hemmings HC: J Pharmacol Exp Ther 304:1188,

2003.

Downloaded from ClinicalKey.com at Buddhist TzuFor personal use only. No other uses without permission

Chapter 25: Inhaled Anesthetics: Mechanisms of Action 637

212. Westphalen RI, Hemmings HC Jr: J Pharmacol Exp Ther 316:208, 2006.

213. Nagele P, et al: Anesthesiology 103:768, 2005. 214. van Swinderen B, et al: Proc Natl Acad Sci U S A 96:2479, 1999. 215. Hemmings HC Jr, et al: Mol Pharmacol 67:1591, 2005. 216. Pittson S, et al: BMC Neurosci 5:52, 2004. 217. Eckenhoff ME, et al: J Pharmacol Exp Ther 300:172, 2002. 218. Pearce RA, et al: Anesthesiology 71:591, 1989. 219. Pearce RA: J Physiol 492:823, 1996. 220. Simon W, et al: Anesthesiology 94:1058, 2001. 221. Jinks SL, et al: Anesthesiology 103:567, 2005. 222. Jinks SL, et al: Neuroreport 14(22):655, 2011. 223. Penttonen M, Buzsaki G: Thalamus Relat Syst 2:145, 2003. 224. Lewis LD, et al: Proc Natl Acad Sci U S A 109(49):E3377, 2012. 225. Steriade M: J Neurosci 13:3266, 1993. 226. Buzsaki G: Neuron 33:325, 2002. 227. Bland BH, et al: Hippocampus 13:38, 2003. 228. Madler C, et al: Br J Anaesth 66:81, 1991. 229. Munglani R, et al: Br J Anaesth 71:633, 1993. 230. Buzsaki G: Annu Rev Neurosci 35:203, 2012. 231. Dickinson R, et al: Neuropharmacology 44:864, 2003. 232. Antkowiak B, Hentschke H: Neurosci Lett 231:87, 1997. 233. Szarecka A, et al: Biophys J 93:1895, 2007. 234. Gottschalk A, Haney P: Anesthesiology 98:548, 2003. 235. Storer KP, Reeke GN: Anesthesiology 117:780, 2012. 236. Wilson MT, et al: Anesthesiology 104:588, 2006. 237. Steyn-Ross ML, et al: Prog Biophys Mol Biol 85:369, 2004. 238. Zhang Y, et al: Anesth Analg 92:1585, 2001. 239. Zhang Y, et al: Anesth Analg 92:123, 2001. 240. Eger EI, et al: Anesth Analg 102:1397, 2006. 241. Sukhotinsky I, et al: Eur J Neurosci 25:1417, 2007. 242. Eckenhoff RG, et al: ACS Chem Neurosci 1:139, 2010. 243. Nash HA: Br J Anaesth 89:143, 2002. 244. Sonner JM, et al: Mol Pharmacol 68:61, 2005. 245. Borghese CM, et al: Anesthesiology 117:765, 2012. 246. Sedensky MM, et al: Am J Physiol Cell Physiol 280:C1340, 2001. 247. Kayser EB, et al: Anesthesiology 90:545, 1999. 248. Cascio M, et al: Anesth Analg 105:381, 2007. 249. Veselis RA, et al: Anesthesiology 97:329, 2002. 250. Alkire MT, Miller J: Prog Brain Res 150:229, 2005.

Chi General Hospital JC September 17, 2016.. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 25: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

RefeRences

1. Perouansky M: The quest for a unified model of anesthetic action: a century in Claude Bernard’s shadow, Anesthesiology 117:465-474, 2012.

2. Eger EI: A brief history of the origin of minimum alveolar concen-tration (MAC), Anesthesiology 96:238-239, 2002.

3. Eger EI, Saidman LJ, Brandstater B: Minimum alveolar anesthetic concentration: a standard of anesthetic potency, Anesthesiology 26:756-763, 1965.

4. Franks NP, Lieb WR: Temperature dependence of the potency of volatile general anesthetics: implications for in vitro experiments, Anesthesiology 84:716-720, 1996.

5. Seeman P: The membrane expansion theory of anesthesia: direct evidence using ethanol and a high-precision density meter, Expe-rientia 30:759-760, 1974.

6. Franks NP, Lieb WR: The structure of lipid bilayers and the effects of general anaesthetics. An xray and neutron diffraction study, J Mol Biol 133:469-500, 1979.

7. Featherstone RM, Muehlbaecher CA, Debon FL, Forsaith JA: Inter-actions of inert anesthetic gases with proteins, Anesthesiology 22:977-981, 1961.

8. Ueda I, Kamaya H: Kinetic and thermodynamic aspects of the mechanism of general anesthesia in a model system of firefly luminescence in vitro, Anesthesiology 38:425-436, 1973.

9. Franks NP, Lieb WR: Seeing the light: protein theories of general anesthesia: 1984, Anesthesiology 101:235-237, 2004.

10. Franks NP, Lieb WR: Do general anaesthetics act by competitive binding to specific receptors? Nature 310:599-601, 1984.

11. Koblin DD, Chortkoff BS, Laster MJ, et al: Polyhalogenated and perfluorinated compounds that disobey the Meyer-Overton hypothesis, Anesth Analg 79:1043-1048, 1994.

12. Hall AC, Lieb WR, Franks NP: Stereoselective and non-stereoselective actions of isoflurane on the GABAA receptor, Br J Pharmacol 112: 906-910, 1994.

13. Eger EI 2nd, Fisher DM, Dilger JP, et al: Relevant concentrations of inhaled anesthetics for in vitro studies of anesthetic mechanisms, Anesthesiology 94:915-921, 2001.

14. Eckenhoff RG: Promiscuous ligands and attractive cavities: how do the inhaled anesthetics work? Mol Interv 5:258-268, 2001.

15. Eckenhoff RG, Johansson JS: What are “relevant” concentrations? Anesthesiology 95:1537-1539, 2001.

16. Rudolph U, Antkowiak B: Molecular and neuronal substrates for general anaesthetics, Nat Rev Neurosci 5:709-720, 2004.

17. Antognini JF, Schwartz K: Exaggerated anesthetic requirements in the preferentially anesthetized brain, Anesthesiology 79: 1244-1249, 1993.

18. Rampil IJ, Mason P, Singh H: Anesthetic potency (MAC) is inde-pendent of forebrain structures in the rat, Anesthesiology 78: 707-712, 1993.

19. Stabernack C, Zhang Y, Sonner JM, et al: Thiopental produces immobility primarily by supraspinal actions in rats, Anesth Analg 100:128-136, 2005.

20. Veselis RA, Reinsel RA, Feshchenko VA: Drug-induced amnesia is a separate phenomenon from sedation—electrophysiologic evi-dence, Anesthesiology 95:896-907, 2001.

21. Sanders R, Tononi G, Laureys S, Sleigh JW: Unresponsiveness ≠ unconsciousness, Anesthesiology 116:946-959, 2012.

22. Devor M, Zalkind V: Reversible analgesia, atonia, and loss of con-sciousness on bilateral intracerebral microinjection of pentobarbital, Pain 94:101-112, 2001.

23. Nelson LE, Guo TZ, Lu J, et al: The sedative component of anes-thesia is mediated by GABA(A) receptors in an endogenous sleep pathway, Nat Neurosci 5:979-984, 2002.

24. Rampil IJ, Laster MJ: No correlation between quantitative elec-troencephalographic measurements and movement response to noxious stimuli during isoflurane anesthesia in rats, Anesthesiol-ogy 77:920-925, 1992.

25. Antognini JF, Carstens E, Atherley R: Does the immobilizing effect of thiopental in brain exceed that of halothane? Anesthesiology 96:980-986, 2002.

26. Zhang Y, Sonner JM, Eger EI, et al: Gamma-aminobutyric acidA receptors do not mediate the immobility produced by isoflurane, Anesth Analg 99:85-90, 2004.

Downloaded from ClinicalKey.com at Buddhist Tzu CFor personal use only. No other uses without permission. C

27. Raines DE, Claycomb RJ, Forman SA: Nonhalogenated anesthetic alkanes and perhalogenated nonimmobilizing alkanes inhibit alpha(4)beta(2) neuronal nicotinic acetylcholine receptors [com-ment] [erratum appears in Anesth Analg 95:869, 2002], Anesth Analg 95:573-577, 2002.

28. Zhang Y, Guzinski M, Eger EI 2nd, et al: Bidirectional modulation of isoflurane potency by intrathecal tetrodotoxin and veratridine in rats, Br J Pharmacol 159:872-878, 2010.

29. Liao M, Sonner JM, Jurd R, et al: Beta-3-containing gamma- aminobutyric acid A receptors are not major targets for the amnesic and immobilizing actions of isoflurane, Anesth Analg 101: 412-418, 2005.

30. Sonner JM, Werner DF, Elsen FP, et al: Effect of isoflurane and other potent inhaled anesthetics on minimum alveolar concen-tration, learning, and the righting reflex in mice engineered to express alpha1 gamma-aminobutyric acid type A receptors unre-sponsive to isoflurane, Anesthesiology 106:107-113, 2007.

31. Heurteaux C, Guy N, Laigle C, et al: TREK-1, a K(+) channel involved in neuroprotection and general anesthesia, EMBO J 23:2684-2695, 2004.

32. Linden AM, Aller MI, Leppa E, et al: The in vivo contributions of TASK-1-containing channels to the actions of inhalation anes-thetics, the alpha(2) adrenergic sedative dexmedetomidine, and cannabinoid agonists, J Pharmacol Exp Ther 317:615-626, 2006.

33. Linden AM, Sandu C, Aller MI, et al: TASK-3 knockout mice exhibit exaggerated nocturnal activity, impairments in cogni-tive functions, and reduced sensitivity to inhalation anesthetics, J Pharmacol Exp Ther 323:924-934, 2007.

34. Westphalen RI, Krivitski M, Amarosa A, et al: Reduced inhibition of cortical glutamate and GABA release by halothane in mice lack-ing the K(+) channel, TREK-1, Br J Pharmacol 152:939-945, 2007.

35. Jinks SL, Bravo M, Hayes SG: Volatile anesthetic effects on midbrain-elicited locomotion suggest that the locomotor network in the ventral spinal cord is the primary site for immobility, Anes-thesiology 108:1016-1024, 2008.

36. Searle JR: Consciousness, Annu Rev Neurosci 23:557-578, 2000. 37. Tononi G: Consciousness as integrated information: a provisional

manifesto, The Bio Bull 215:216-242, 2008. 38. Crick F, Koch C: A framework for consciousness, Nat Neurosci

6:119-126, 2003. 39. Alkire MT, Haier RJ, Fallon JH: Toward a unified theory of nar-

cosis: brain imaging evidence for a thalamocortical switch as the neurophysiologic basis of anesthetic-induced unconsciousness [see comments], Conscious Cogn 9:370-386, 2000.

40. Ries CR, Puil E: Mechanism of anesthesia revealed by shunting actions of isoflurane on thalamocortical neurons, J Neurophysiol 81:1795-1801, 1999.

41. Detsch O, Vahle-Hinz C, Kochs E, et al: Isoflurane induces dose-dependent changes of thalamic somatosensory information transfer, Brain Res 829:77-89, 1999.

42. John ER, Prichep LS, Kox W, et al: Invariant reversible QEEG effects of anesthetics, Conscious Cogn 10:165-183, 2001.

43. Chortkoff BS, Eger EI 2nd, Crankshaw DP, et al: Concentrations of desflurane and propofol that suppress response to command in humans, Anesth Analg 81:737-743, 1995.

44. Dwyer R, Bennett HL, Eger EI 2nd, et al: Effects of isoflurane and nitrous oxide in subanesthetic concentrations on memory and responsiveness in volunteers, Anesthesiology 77:888-898, 1992.

45. Alkire MT, Hudetz AG, Tononi G: Consciousness and anesthesia, Science 322:876-880, 2008.

46. Brown E, Lydic R, Schiff ND: General anesthesia, sleep, coma, N Engl J Med 263:2638-2650, 2010.

47. Varela F, Lachaux JP, Rodriguez E, et al: The brainweb: phase synchronization and large-scale integration, Nat Rev Neurosci 2: 229-239, 2001.

48. van den Heuvel MP, Sporns O: Rich-club organization of the human connectome, J Neurosci 31:15775-15786, 2011.

49. Bullmore E, Sporns O: The economy of brain network organiza-tion, Nat Rev Neurosci 13:336-349, 2012.

50. Massimini M, Ferraretti F, Huber R, et al: Breakdown of cortical effective connectivity during sleep, Science 309:2228-2232, 2005.

51. Ferrarelli F, et al: Breakdown in cortical effective connectivity dur-ing midazolam-induced loss of consciousness, Proc Natl Acad Sci U S A 107:2681-2686, 2010.

637.e1

hi General Hospital JC September 17, 2016.opyright ©2016. Elsevier Inc. All rights reserved.

Page 26: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

References637.e2

52. Imas OA, Ropella KM, Ward BD, et al: Volatile anesthetics enhance flash-induced gamma oscillations in rat visual cortex, Anesthesiology 102:937-947, 2005.

53. Imas OA, Ropella KM, Wood JD, et al: Isoflurane disrupts antero-posterior phase synchronization of flash-induced field potentials in the rat, Neurosci Lett 402:216-221, 2006.

54. Ter Mikaelian M, Sanes DH, Semple MN: Transformation of tem-poral properties between auditory midbrain and cortex in the awake Mongolian gerbil, J Neurosci 27:6091-6102, 2007.

55. Burlingame RH, Shrestha S, Rummel MR, Banks MI: Subhypnotic doses of isoflurane impair auditory discrimination in rats, Anes-thesiology 106:754-762, 2007.

56. Dutton RC, Maurer AJ, Sonner JM, et al: The concentration of isoflurane required to suppress learning depends on the type of learning, Anesthesiology 94:514-519, 2001.

57. Alkire MT, Nathan SV: Does the amygdala mediate anesthetic-induced amnesia? Basolateral amygdala lesions block sevoflurane-induced amnesia, Anesthesiology 102:754-760, 2005.

58. Alkire MT, Gorski LA: Relative amnesic potency of five inhala-tional anesthetics follows the Meyer-Overton rule, Anesthesiology 101:417-429, 2004.

59. Vertes RP: Hippocampal theta rhythm: a tag for short-term memory, Hippocampus 15:923-935, 2005.

60. Pan WX, McNaughton N: The medial supramammillary nucleus, spatial learning and the frequency of hippocampal theta activity, Brain Res 764:101-108, 1997.

61. Robbe D, Montgomery SM, Thome A, et al: Cannabinoids reveal importance of spike timing coordination in hippocampal function, Nat Neurosci 9:1526-1533, 2006.

62. Perouansky M, Hentschke H, Perkins M, et al: Amnesic concentra-tions of the nonimmobilizer 1,2-dichlorohexafluorocyclobutane (F6, 2N) and isoflurane alter hippocampal theta oscillations in vivo, Anesthesiology 106:1168-1176, 2007.

63. Seidenbecher T, Laxmi TR, Stork O, et al: Amygdalar and hip-pocampal theta rhythm synchronization during fear memory retrieval, Science 301:846-850, 2003.

64. Pryor KO, Reinsel RA, Mehta M, et al: Visual P2-N2 complex and arousal at the time of encoding predict the time domain charac-teristics of amnesia for multiple intravenous anesthetic drugs in humans, Anesthesiology 113:313-326, 2010.

65. Rudolph U, Crestani F, Benke D, et al: Benzodiazepine actions mediated by specific gamma-aminobutyric acid(A) receptor sub-types, Nature 401:796-800, 1999.

66. Kandel L, Chortkoff BS, Sonner J, et al: Nonanesthetics can sup-press learning, Anesth Analg 82:321-326, 1996.

67. Mihic SJ, McQuilkin SJ, Eger EI 2nd, et al: Potentiation of gamma-aminobutyric acid type A receptor–mediated chloride currents by novel halogenated compounds correlates with their abilities to induce general anesthesia, Mol Pharmacol 46:851-857, 1994.

68. Zarnowska ED, Pearce RA, Saad AA, et al: The gamma-subunit gov-erns the susceptibility of recombinant gamma-aminobutyric acid type A receptors to block by the nonimmobilizer 1,2-dichlorohexa-fluorocyclobutane (F6, 2N), Anesth Analg 101:401-406, 2005.

69. Jevtovic-Todorovic V, Todorovic SM, Mennerick S, et al: Nitrous oxide (laughing gas) is an NMDA antagonist, neuroprotectant and neurotoxin, Nat Med 4:460-463, 1998.

70. Gruss M, Bushell TJ, Bright DP, et al: Two-pore-domain K+channels are a novel target for the anesthetic gases xenon, nitrous oxide, and cyclopropane, Mol Pharmacol 65:443-452, 2004.

71. Gries DA, Condouris GA, Shey Z, et al: Anxiolytic-like action in mice treated with nitrous oxide and oral triazolam or diazepam, Life Sci 76:1667-1674, 2005.

72. Murphy M: Propofol anesthesia and sleep: a high-density EEG study, Sleep 34:283-291, 2011.

73. Tung A, Bergmann BM, Herrera S, et al: Recovery from sleep deprivation occurs during propofol anesthesia, Anesthesiology 100:1419-1426, 2004.

74. Nelson A, Faraguna U, Tononi G, Cirelli C: Effects of anesthesia on the response to sleep deprivation, Sleep 33:1659-1667, 2010.

75. Hentschke H, Schwarz C, Antkowiak B: Neocortex is the major target of sedative concentrations of volatile anaesthetics: strong depression of firing rates and increase of GABAA receptor-mediated inhibition, Eur J Neurosci 21:93-102, 2005.

76. Perouansky M, Hemmings HC Jr: General anesthetics and neuro-toxicity: cause for concern? Anesthesiology 111:1365-1371, 2009.

Downloaded from ClinicalKey.com at BuddhistFor personal use only. No other uses without permiss

77. Durieux M, Davis PJ: The Safety of Key Inhaled and Intravenous Drugs in Pediatrics (SAFEKIDS): a update, Anesth Analg 110: 1265-1267, 2010.

78. Jevtovic-Todorovic V, Hartman RE, Izumi Y, et al: Early exposure to common anesthetic agents causes widespread neurodegenera-tion in the developing rat brain and persistent learning deficits, J Neurosci 23:876-882, 2003.

79. Shih J, May LD, Gonzalez HE, et al: Delayed environmental enrich-ment reverses sevoflurane-induced memory impairment in rats, Anesthesiology 116:586-602, 2012.

80. Kitano H, Kirsch JR, Hurn PD, et al: Inhalational anesthetics as neuroprotectants or chemical preconditioning agents in ischemic brain, J Cereb Blood Flow Metab 27:1108-1128, 2006.

81. Fukuda S, Warner DS: Cerebral protection, Br J Anaesth 99:10-17, 2007.

82. Kawaguchi M, Furuya H, Patel PM: Neuroprotective effects of anesthetic agents, J Anesth 19:150-156, 2005.

83. Inoue S, Drummond JC, Davis DP, et al: Combination of isoflu-rane and caspase inhibition reduces cerebral injury in rats sub-jected to focal cerebral ischemia, Anesthesiology 101:75-81, 2004.

84. Preckel B, Weber NC, Sanders RD, et al: Molecular mechanisms transducing the anesthetic, analgesic, and organ-protective actions of xenon, Anesthesiology 105:187-197, 2006.

85. Turner CP, Gutierrez S, Liu C, et al: Strategies to defeat ketamine-induced neonatal brain injury, Neuroscience 17:384-392, 2012.

86. Kitano H, Kirsch JR, Hurn PD, Murphy SJ: Inhalational anesthetics as neuroprotectants or chemical preconditioning agents in isch-emic brain, J Cereb Blood Flow Metab 27:1108-1128, 2007.

87. Rasmussen L, Siersma VD, ISPOCD Group: Postoperative cogni-tive dysfunction: true deterioration versus random variation, Acta Anaesthesiol Scand 48:1137-1143, 2004.

88. Zurek A, Bridgwater EM, Orser BA: Inhibition of alpha5 gamma-Aminobutyric acid type A receptors restores recognition memory after general anesthesia Short-term alterations in mental function, Anesth Analg 114:845-855, 2012.

89. Wiklund A, Granon S, Faure P, et al: Object memory in young and aged mice after sevoflurane anaesthesia, Anesthesiology 109:790-798, 2008.

90. Wan Y, Xu J, Ma D, et al: Postoperative impairment of cognitive function in rats: a possible role for cytokine-mediated inflamma-tion in the hippocampus, Anesthesiology 106:436-443, 2007.

91. Terrando N, Eriksson LI, Ryu JK, et al: Resolving postoperative neuroinflammation and cognitive decline, Ann Neurol 70:986-995, 2011.

92. Tang J, Mardini F, Caltagarone BM, et al: Anesthesia in presymptom-atic Alzheimer’s disease: a study using the triple-transgenic mouse model, Alzheimer’s Dement 7:521-531, 2011.

93. Avidan M, Searleman AC, Storandt M, et al: Long-term cognitive decline in older subjects was not attributable to noncardiac surgery or major illness, Anesthesiology 111:964-970, 2009.

94. Avidan M, Evers A: Review of clinical evidence for persistent cog-nitive decline or incident dementia attributable to surgery or gen-eral anesthesia, J Alzheimers Dis 24:201-216, 2011.

95. Zaugg M, Lucchinetti E, Uecker M, et al: Anaesthetics and cardiac preconditioning: I. Signalling and cytoprotective mechanisms, Br J Anaesth 91:551-565, 2003.

96. Pagel PS, Warltier DC: Ventricular function. In Warltier DC, edi-tor: Anesthetics and left ventricular function. Baltimore, 1995, Wil-liams & Wilkins, pp 213-252.

97. Hanley PJ, ter Keurs HE, Cannell MB: Excitation-contraction cou-pling in the heart and the negative inotropic action of volatile anesthetics, Anesthesiology 101:999-1014, 2004.

98. Huneke R, Jungling E, Skasa M, et al: Effects of the anesthetic gases xenon, halothane, and isoflurane on calcium and potas-sium currents in human atrial cardiomyocytes, Anesthesiology 95: 999-1006, 2001.

99. Stowe DF, Rehmert GC, Kwok WM, et al: Xenon does not alter cardiac function or major cation currents in isolated guinea pig hearts or myocytes, Anesthesiology 92:516-522, 2000.

100. Ebert TJ, Kampine JP: Nitrous oxide augments sympathetic out-flow: direct evidence from human peroneal nerve recordings, Anesth Analg 69:444-449, 1989.

101. Stowe DF, Monroe SM, Marijic J, et al: Effects of nitrous oxide on contractile function and metabolism of the isolated heart, Anes-thesiology 73:1220-1226, 1990.

Tzu Chi General Hospital JC September 17, 2016.ion. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 27: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

102. Pagel PS, Kampine JP, Schmeling WT, et al: Alteration of left ventricular diastolic function by desflurane, isoflurane, and halo-thane in the chronically instrumented dog with autonomic ner-vous system blockade, Anesthesiology 74:1103-1114, 1991.

103. Tanaka K, Kawano T, Nakamura A, et al: Isoflurane activates sarco-lemmal adenosine diphosphate-sensitive potassium channels in vascular smooth muscle cells—a role for protein kinase A, Anes-thesiology 106:984-991, 2007.

104. Yoshino J, Akata T, Izumi K, et al: Multiple actions of halothane on contractile response to noradrenaline in isolated mesenteric resis-tance arteries, Naunyn Schmied Arch Pharmacol 371:500-515, 2005.

105. Vulliemoz Y: The nitric oxide-cyclic 3’,5’-guanosine monophos-phate signal transduction pathway in the mechanism of action of general anesthetics, Toxicol Lett 100-101:103-108, 1998.

106. Huneke R, Fassl J, Rossaint R, et al: Effects of volatile anesthetics on cardiac ion channels, Acta Anaesthesiol Scand 48:547-561, 2004.

107. Stadnicka A, Marinovic J, Ljubkovic M, et al: Volatile anesthetic-induced cardiac preconditioning, J Anesth 21:212-219, 2007.

108. Suleiman MS, Zacharowski K, Angelini GD: Inflammatory response and cardioprotection during open-heart surgery: the importance of anaesthetics, Br J Pharmacol 153:21-33, 2008.

109. Pratt PF Jr, Wang C, Weihrauch D, et al: Cardioprotection by vola-tile anesthetics: new applications for old drugs? Curr Opin Anaesth 19:397-403, 2006.

110. Weber NC, Toma O, Awan S, et al: Effects of nitrous oxide on the rat heart in vivo—another inhalational anesthetic that precondi-tions the heart? Anesthesiology 103:1174-1182, 2005.

111. Tanaka K, Ludwig LM, Kersten JR, et al: Mechanisms of cardiopro-tection by volatile anesthetics, Anesthesiology 100:707-721, 2004.

112. Stuth EAE, Krolo M, Tonkovic-Capin M, et al: Effects of halothane on synaptic neurotransmission to medullary expiratory neurons in the ventral respiratory group of dogs, Anesthesiology 91:804-814, 1999.

113. Franks NP, Lieb WR: Molecular and cellular mechanisms of gen-eral anaesthesia, Nature 367:607-614, 1994.

114. Zeller A, Jurd R, Lambert S, et al: Inhibitory ligand-gated ion chan-nels as substrates for general anesthetic actions, Handb Exp Phar-macol 182:31-51, 2008.

115. Bertaccini EJ, Trudell JR, Franks NP: The common chemical motifs within anesthetic binding sites, Anesth Analg 104:318-324, 2007.

116. Nury H, Van Renterghem C, Weng Y, et al: X-ray structures of general anaesthetics bound to a pentameric ligand-gated ion channel, Nature 469(7330):428-431, 2011.

117. Dickinson R, Peterson BK, Banks P, et al: Competitive inhibition at the glycine site of the N-methyl-d-aspartate receptor by the anes-thetics xenon and Isoflurane, Anesthesiology 107:756-767, 2007.

118. Hemmings HC Jr, Akabas MH, Goldstein PA, et al: Emerging molecular mechanisms of general anesthetic action, Trends Phar-macol Sci 26:503-510, 2005.

119. Patel AJ, Honore E, Lesage F, et al: Inhalational anesthetics acti-vate two-pore-domain background K+channels, Nat Neurosci 2:422-426, 1999.

120. Lynch JW: Molecular structure and function of the glycine recep-tor chloride channel, Physiol Rev 84:1051-1095, 2004.

121. Mihic SJ, Ye Q, Wick MJ, et al: Sites of alcohol and volatile anaes-thetic action on GABA(A) and glycine receptors, Nature 389: 385-389, 1997.

122. Solt K, Stevens RJ, Davies PA, et al: General anesthetic-induced channel gating enhancement of 5-hydroxytryptamine type 3 receptors depends on receptor subunit composition, J Pharmacol Exp Ther 315:771-776, 2005.

123. Role LW, Berg DK: Nicotinic receptors in the development and modulation of CNS synapses, Neuron 16:1077-1085, 1996.

124. Flood P, Ramirez-Latorre J, Role L: Alpha 4 beta 2 neuronal nicotinic acetylcholine receptors in the central nervous system are inhibited by isoflurane and propofol, but alpha 7-type nicotinic acetylcholine receptors are unaffected, Anesthesiology 86:859-865, 1997.

125. Violet JM, Downie DL, Nakisa RC, et al: Differential sensitivities of mammalian neuronal and muscle nicotinic acetylcholine recep-tors to general anesthetics, Anesthesiology 86:866-874, 1997.

126. Dingledine R, Borges K, Bowie D, et al: The glutamate receptor ion channels, Pharmacol Rev 51:7-61, 1999.

127. Franks NP, Dickinson R, de Sousa SL, et al: How does xenon pro-duce anaesthesia? Nature 396:324, 1998.

Downloaded from ClinicalKey.com at Buddhist Tzu CFor personal use only. No other uses without permission. C

References 637.e3

128. Solt K, Eger EI, Raines DE: Differential modulation of human N-methyl-d-aspartate receptors by structurally diverse general anesthetics, Anesth Analg 102:1407-1411, 2006.

129. Harris RA, Mihic SJ, Dildy-Mayfield JE, et al: Actions of anesthetics on ligand-gated ion channels: role of receptor subunit composi-tion, FASEB J 9:1454-1462, 1995.

130. Sonner JM, Vissel B, Royle G, et al: The effect of three inhaled anesthetics in mice harboring mutations in the GluR6 (kainate) receptor gene, Anesth Analg 101:143-148, 2005.

131. MacIver MB, Mikulec AA, Amagasu SM, et al: Volatile anesthetics depress glutamate transmission via presynaptic actions, Anesthesi-ology 85:823-834, 1996.

132. Perouansky M, Hemmings HC Jr, Pearce RA: Anesthetic effects on glutamatergic neurotransmission: lessons learned from a large synapse, Anesthesiology 100:470-472, 2004.

133. Winegar BD, MacIver MB: Isoflurane depresses hippocampal CA1 glutamate nerve terminals without inhibiting fiber volleys, BMC Neurosci 7:5, 2006.

134. Haydon DA, Urban BW: The effects of some inhalation anaesthet-ics on the sodium current of the squid giant axon, J Physiol (Lond) 341:429-439, 1983.

135. Berg-Johnsen J, Langmoen IA: The effect of isoflurane on unmy-elinated and myelinated fibres in the rat brain, Acta Physiol Scand 127:87-93, 1986.

136. Mikulec AA, Pittson S, Amagasu SM, et al: Halothane depresses action potential conduction in hippocampal axons, Brain Res 796:231-238, 1998.

137. Wu XS, Sun JY, Evers AS, et al: Isoflurane inhibits transmit-ter release and the presynaptic action potential, Anesthesiology 100:663-670, 2004.

138. Yu FH, Catterall WA: The VGL-chanome: a protein superfamily specialized for electrical signaling and ionic homeostasis, Sci STKE 2004:re15, 2004.

139. Herold KF, Nau C, Ouyang W, Hemmings HC Jr: Isoflurane inhibits the tetrodotoxin-resistant voltage-gated sodium channel Nav1.8, Anesthesiology 111:591-599, 2009.

140. OuYang W, Hemmings HC Jr: Depression by isoflurane of the action potential and underlying voltage-gated ion currents in iso-lated rat neurohypophysial nerve terminals, J Pharmacol Exp Ther 312:801-808, 2005.

141. OuYang W, Jih T-Y, Zhang T-T, et al: Isoflurane inhibits NaChBac, a prokaryotic voltage-gated sodium channel, J Pharmacol Exp Ther 322:1076-1083, 2007.

142. Ouyang W, Hemmings HC Jr: Isoform-selective effects of isoflu-rane on voltage-gated Na+ channels, Anesthesiology 107:91-98, 2007.

143. Ratnakumari L, Vysotskaya TN, Duch DS, et al: Differential effects of anesthetic and nonanesthetic cyclobutanes on neuro-nal voltage-gated sodium channels, Anesthesiology 92:529-541, 2000.

144. Catterall WA: Structure and regulation of voltage-gated Ca2+ channels, Annu Rev Cell Dev Biol 16:521-555, 2000.

145. Topf N, Recio-Pinto E, Blanck TJ, et al: Actions of general anes-thetics on voltage-gated ion channels. In Antognini JF, Carlens E, Raines D, editors: Neural mechanisms of anesthesia. Totowa, NJ, 2003, Humana Press, pp 299-318.

146. Miao N, Frazer MJ, Lynch C 3rd: Volatile anesthetics depress Ca2+ transients and glutamate release in isolated cerebral synapto-somes, Anesthesiology 83:593-603, 1995.

147. Kameyama K, Aono K, Kitamura K: Isoflurane inhibits neuronal Ca2+ channels through enhancement of current inactivation, Br J Anaesth 82:402-411, 1999.

148. Study RE: Isoflurane inhibits multiple voltage-gated calcium cur-rents in hippocampal pyramidal neurons [see comments], Anes-thesiology 81:104-116, 1994.

149. Hall AC, Lieb WR, Franks NP: Insensitivity of P-type calcium channels to inhalational and intravenous general anesthetics, Anesthesiology 81:117-123, 1994.

150. Takei T, Saegusa H, Zong S, et al: Increased sensitivity to halo-thane but decreased sensitivity to propofol in mice lacking the N-type Ca2+ channel, Neurosci Lett 350:41-45, 2003.

151. Joksovic PM, Brimelow BC, Murbartian J, et al: Contrasting anes-thetic sensitivities of T-type Ca2+ channels of reticular thalamic neurons and recombinant Ca(v)3.3 channels, Br J Pharmacol 144:59-70, 2005.

hi General Hospital JC September 17, 2016.opyright ©2016. Elsevier Inc. All rights reserved.

Page 28: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

References637.e4

152. Todorovic SM, Jevtovic-Todorovic V, Mennerick S, et al: Ca(v)3.2 channel is a molecular substrate for inhibition of T-type calcium currents in rat sensory neurons by nitrous oxide, Mol Pharmacol 60:603-610, 2001.

153. Petrenko AB, Tsujita M, Kohno T, et al: Mutation of alpha(1G) T-type calcium channels in mice does not change anesthetic requirements for loss of the righting reflex and minimum alveolar concentration but delays the onset of anesthetic induction, Anes-thesiology 106:1177-1185, 2007.

154. Rithalia A, Hopkins PM, Harrison SM: The effects of halothane, isoflurane, and sevoflurane on Ca2+ current and transient outward K+ current in subendocardial and subepicardial myocytes from the rat left ventricle, Anesth Analg 99:1615-1622, 2004.

155. Davies LA, Gibson CN, Boyett MR, et al: Effects of isoflurane, sevoflurane, and halothane on myofilament Ca2+ sensitivity and sarcoplasmic reticulum Ca2+ release in rat ventricular myocytes, Anesthesiology 93:1034-1044, 2000.

156. Pabelick CM, Prakash YS, Kannan MS, et al: Effects of halothane on sarcoplasmic reticulum calcium release channels in porcine airway smooth muscle cells, Anesthesiology 95:207-215, 2001.

157. Roberts MC, Mickelson JR, Patterson EE, et al: Autosomal domi-nant canine malignant hyperthermia is caused by a mutation in the gene encoding the skeletal muscle calcium release channel (RYR1), Anesthesiology 95:716-725, 2001.

158. Mickelson JR, Louis CF: Malignant hyperthermia: excitation- contraction coupling, Ca2+ release channel, and cell Ca2+ regula-tion defects, Physiol Rev 76:537-592, 1996.

159. Yost CS: Potassium channels: basic aspects, functional roles, and medical significance, Anesthesiology 90:1186-1203, 1999.

160. Friederich P, Benzenberg D, Trellakis S, et al: Interaction of vola-tile anesthetics with human Kv channels in relation to clinical concentrations, Anesthesiology 95:954-958, 2001.

161. Franks NP, Lieb WR: Volatile general anesthetics activate a novel neuronal K+ current, Nature 333:662-664, 1988.

162. Franks NP, Honore E: The TREK K-2P channels and their role in gen-eral anaesthesia and neuro, Trends Pharmacol Sci 25:601-608, 2004.

163. Patel AJ, Honore E: Anesthetic-sensitive 2P domain K+ channels, Anesthesiology 95:1013-1021, 2001.

164. Farwell D, Gollob MH: Electrical heart disease: genetic and molec-ular basis of cardiac arrhythmias in normal structural hearts, Can J Cardiol 23(Suppl A):16A-22A, 2007.

165. Antzelevitch C: Molecular biology and cellular mechanisms of Brugada and long QT syndromes in infants and young children, J Electrocardiol 34(Suppl):177-181, 2001.

166. Li J, Correa AM: Kinetic modulation of HERG potassium channels by the volatile anesthetic halothane, Anesthesiology 97:921-930, 2002.

167. Davies LA, Hopkins PM, Boyett MR, et al: Effects of halothane on the transient outward K(+) current in rat ventricular myocytes, Br J Pharmacol 131:223-230, 2000.

168. Sirois JE, Lynch C III, Bayliss DA: Convergent and reciprocal mod-ulation of a leak K+ current and I(h) by an inhalational anaesthetic and neurotransmitters in rat brainstem motoneurones, J Physiol 541:717-729, 2002.

169. Chen X, Sirois JE, Lei Q, et al: HCN subunit-specific and cAMP-modulated effects of anesthetics on neuronal pacemaker currents, J Neurosci 25:5803-5814, 2005.

170. Robinson RB, Siegelbaum SA: Hyperpolarization-activated cat-ion currents: from molecules to physiological function, Annu Rev Physiol 65:453-480, 2003.

171. Girault JA, Hemmings HC Jr: Cell signaling. In Hemmings HC Jr, Hopkins PM, editors: Foundations of anesthesia, ed 2. London, 2005, Mosby, pp 31-50.

172. Rebecchi MJ, Pentyala SN: Anaesthetic actions on other targets: protein kinase C and guanine nucleotide-binding proteins, Br J Anaesth 89:62-78, 2002.

173. Peterlin Z, Ishizawa Y, Araneda R, et al: Selective activation of G-protein–coupled receptors by volatile anesthetics, Mol Cell Neu-rosci 30:506-512, 2005.

174. Minami K, Gereau RW, Minami M, et al: Effects of ethanol and anes-thetics on type 1 and 5 metabotropic glutamate receptors expressed in Xenopus laevis oocytes, Mol Pharmacol 53:148-156, 1998.

175. Minami K, Minami M, Harris RA: Inhibition of 5-hydroxytrypta-mine type 2A receptor-induced currents by n-alcohols and anes-thetics, J Pharmacol Exp Ther 281:1136-1143, 1997.

176. Minami K, Vanderah TW, Minami M, et al: Inhibitory effects of anesthetics and ethanol on muscarinic receptors expressed in Xenopus oocytes, Eur J Pharmacol 339:237-244, 1997.

Downloaded from ClinicalKey.com at BuddhisFor personal use only. No other uses without permis

177. Hemmings HC Jr: General anesthetic effects on protein kinase C, Toxicol Lett 100:89-95, 1998.

178. Hasegawa J, Takekoshi S, Nagata H, et al: Sevoflurane stimulates MAP kinase signal transduction through the activation of PKC alpha and betaII in fetal rat cerebral cortex cultured neuron, Acta Histochem Cytochem 39:163-172, 2006.

179. Das J, Addona GH, Sandberg WS, et al: Identification of a general anesthetic binding site in the diacylglycerol-binding domain of protein kinase C-delta, J Biol Chem 279:37964-37972, 2004.

180. Shumilla JA, Sweitzer SM, Eger EI 2nd, et al: Inhibition of spi-nal protein kinase C-epsilon or -gamma isozymes does not affect halothane minimum alveolar anesthetic concentration in rats, Anesth Analg 99:82-84, 2004.

181. Sonner JM, Gong D, Li J, et al: Mouse strain modestly influences minimum alveolar anesthetic concentration and convulsivity of inhaled compounds, Anesth Analg 89:1030-1034, 1999.

182. Snyder GL, Galdi S, Hendrick JP, et al: General anesthetics selec-tively modulate glutamatergic and dopaminergic signaling via site-specific phosphorylation in vivo, Neuropharmacology 53: 619-630, 2007.

183. Hemmings HC Jr, Adamo AI: Effects of halothane and propo-fol on purified brain protein kinase C activation, Anesthesiology 81:147-155, 1994.

184. Marota JJ, Crosby G, Uhl GR: Selective effects of pentobarbital and halothane on c-fos and jun-B gene expression in rat brain, Anesthesiology 77:365-371, 1992.

185. Hamaya Y, Takeda T, Dohi S, et al: The effects of pentobarbital, isoflurane, and propofol on immediate-early gene expression in the vital organs of the rat, Anesth Analg 90:1177-1183, 2000.

186. Culley DJ, Yukhananov RY, Xie ZC, et al: Altered hippocampal gene expression 2 days after general anesthesia in rats, Eur J Phar-macol 549:71-78, 2006.

187. Fütterer CD, Maurer MH, Schmitt A, et al: Alterations in rat brain pro-teins after desflurane anesthesia, Anesthesiology 100:302-308, 2004.

188. Kullmann DM, Martin RL, Redman SJ: Reduction by general anaesthetics of group Ia excitatory postsynaptic potentials and currents in the cat spinal cord, J Physiol (Lond) 412:277-296, 1989.

189. Fujiwara N, Higashi H, Nishi S, et al: Changes in spontaneous firing patterns of rat hippocampal neurones induced by volatile anaesthetics, J Physiol (Lond) 402:155-175, 1988.

190. MacIver MB, Roth SH: Inhalation anaesthetics exhibit path-way-specific and differential actions on hippocampal synaptic responses in vitro, Br J Anaesth 60:680-691, 1988.

191. Ries CR, Puil E: Ionic mechanism of isoflurane’s actions on thala-mocortical neurons, J Neurophysiol 81:1802-1809, 1999.

192. Sirois JE, Lei Q, Talley EM, et al: The TASK-1 two-pore domain K+ channel is a molecular substrate for neuronal effects of inhalation anesthetics, J Neurosci 20:6347-6354, 2000.

193. Semyanov A, Walker MC, Kullmann DM, et al: Tonically active GABA(A) receptors: modulating gain and maintaining the tone, Trends Neurosci 27:262-269, 2004.

194. Bai D, Zhu G, Pennefather P, et al: Distinct functional and phar-macological properties of tonic and quantal inhibitory postsyn-aptic currents mediated by γ-aminobutyric acidA receptors in hippocampal neurons, Mol Pharmacol 59:814-824, 2001.

195. Bieda MC, MacIver MB: Major role for tonic GABAA conductances in anesthetic suppression of intrinsic neuronal excitability, J Neu-rophysiol 92:1658-1667, 2004.

196. Caraiscos VB, Elliott EM, You T, et al: Tonic inhibition in mouse hippocampal CA1 pyramidal neurons is mediated by α 5 subunit-containing gamma-aminobutyric acid type A receptors, Proc Natl Acad Sci U S A 101:3662-3667, 2004.

197. Caraiscos VB, Newell JG, You T, et al: Selective enhancement of tonic GABAergic inhibition in murine hippocampal neurons by low concentrations of the volatile anesthetic isoflurane, J Neurosci 24:8454-8458, 2004.

198. Capogna M, Pearce RA: GABAA,slow: causes and consequences, Trends Neurosci 34:101-112, 2011.

199. Dai S, Perouansky M, Pearce RA: Isoflurane enhances both fast and slow synaptic inhibition in the hippocampus at amnestic concentrations, Anesthesiology 116:816-823, 2012.

200. Berg-Johnsen J, Langmoen IA: The effect of isoflurane on excit-atory synaptic transmission in the rat hippocampus, Acta Anaes-thesiol Scand 36:350-355, 1992.

201. Kirson ED, Yaari Y, Perouansky M: Presynaptic and postsynaptic actions of halothane at glutamatergic synapses in the mouse hip-pocampus, Br J Pharmacol 124:1607-1614, 1998.

t Tzu Chi General Hospital JC September 17, 2016.sion. Copyright ©2016. Elsevier Inc. All rights reserved.

Page 29: Chapter 25 - Inhaled Anesthetics: Mechanisms of Action...Inhaled Anesthetics: Mechanisms . of Action. MISHA PEROUANSKY • ROBERT A. PEARCE • HUGH C. HEMMINGS, JR. Despite the widespread

202. Richards CD, Smaje JC: Anaesthetics depress the sensitivity of cor-tical neurones to L-glutamate, Br J Pharmacol 58:347-357, 1976.

203. Wakamori M, Ikemoto Y, Akaike N: Effects of two volatile anes-thetics and a volatile convulsant on the excitatory and inhibitory amino acid responses in dissociated CNS neurons of the rat, J Neu-rophysiol 66:2014-2021, 1991.

204. Yang J, Zorumski CF: Effects of isoflurane on N-methyl-d-aspartate gated ion channels in cultured rat hippocampal neurons, Ann NY Acad Sci 625:287-289, 1991.

205. de Sousa SLM, Dickinson R, Lieb WR, et al: Contrasting synap-tic actions of the inhalational general anesthetics isoflurane and xenon, Anesthesiology 92:1055-1066, 2000.

206. Dildy-Mayfield JE, Eger EI 2nd, Harris RA: Anesthetics produce subunit-selective actions on glutamate receptors, J Pharmacol Exp Ther 276:1058-1065, 1996.

207. Minami K, Wick MJ, Stern-Bach Y, et al: Sites of volatile anesthetic action on kainate (Glutamate receptor 6) receptors, J Biol Chem 273:8248-8255, 1998.

208. Banks MI, Pearce RA: Dual actions of volatile anesthetics on GABA(A) IPSCs: dissociation of blocking and prolonging effects, Anesthesiology 90:120-134, 1999.

209. Murugaiah KD, Hemmings HC Jr: Effects of intravenous general anesthetics on [3H]GABA release from rat cortical synaptosomes, Anesthesiology 89:919-928, 1998.

210. Nishikawa K, MacIver MB: Agent-selective effects of volatile anes-thetics on GABA(A) receptor-mediated synaptic inhibition in hip-pocampal interneurons, Anesthesiology 94:340-347, 2001.

211. Westphalen RI, Hemmings HC: Selective depression by general anesthetics of glutamate versus GABA release from isolated corti-cal nerve terminals, J Pharmacol Exp Ther 304:1188-1196, 2003.

212. Westphalen RI, Hemmings HC Jr: Volatile anesthetic effects on glutamate versus GABA release from isolated rat cortical nerve ter-minals: basal release, J Pharmacol Exp Ther 316:208-215, 2006.

213. Nagele P, Mendel JB, Placzek WJ, et al: Volatile anesthetics bind rat synaptic snare proteins, Anesthesiology 103:768-778, 2005.

214. van Swinderen B, Saifee O, Shebester L, et al: A neomorphic syn-taxin mutation blocks volatile-anesthetic action in Caenorhabditis elegans, Proc Natl Acad Sci U S A 96:2479-2484, 1999.

215. Hemmings HC Jr, Yan W, Westphalen RI, Ryan TA: The general anesthetic isoflurane depresses synaptic vesicle exocytosis, Mol Pharmacol 67:1591-1599, 2005.

216. Pittson S, Himmel AM, MacIver MB: Multiple synaptic and mem-brane sites of anesthetic action in the CA1 region of rat hippo-campal slices, BMC Neurosci 5:52, 2004.

217. Eckenhoff ME, Chan K, Eckenhoff RG: Multiple specific binding targets for inhaled anesthetics in the mammalian brain, J Pharma-col Exp Ther 300:172-179, 2002.

218. Pearce RA, Stringer JL, Lothman EW: Effect of volatile anesthetics on synaptic transmission in the rat hippocampus, Anesthesiology 71:591-598, 1989.

219. Pearce RA: Volatile anaesthetic enhancement of paired-pulse depression investigated in the rat hippocampus in vitro, J Physiol 492:823-840, 1996.

220. Simon W, Hapfelmeier G, Kochs E, et al: Isoflurane blocks synap-tic plasticity in the mouse hippocampus, Anesthesiology 94:1058-1065, 2001.

221. Jinks SL, Atherley RJ, Dominguez CL, et al: Isoflurane disrupts central pattern generator activity and coordination in the lam-prey isolated spinal cord, Anesthesiology 103:567-575, 2005.

222. Jinks SL, Andrada J, Satter O: Anesthetic effect on fictive locomo-tion in the rat isolated spinal cord, Neuroreport 14:655-659, 2011.

223. Penttonen M, Buzsaki G: Natural logarithmic relationships between brain oscillators, Thalamus Relat Syst 2:145-152, 2003.

224. Lewis LD, Weiner VS, Mukamel JA et al: Rapid fragmentation of neuronal networks at the onset of propofol-induced unconscious-ness. Proc Natl Acad Sci U S A 109(49):E3377-E3386, 2012.

225. Steriade M: Intracellular analysis of relations between the slow (<1 Hz) neocortical oscillation and other sleep rhythms of the electroencephalogram, J Neurosci 13:3266-3283, 1993.

226. Imas OA, Ropella KMWard BD, et al: Volatile anesthestics dis-rupt fontal-posterior recurrent information transfer at gamma fre-quencies in rat, Neurosci Lett 387:145-150, 2005.

227. Buzsaki G: Theta oscillations in the hippocampus, Neuron 33: 325-340, 2002.

228. Bland BH, Bland CE, Colom LV, et al: Effect of halothane on type 2 immobility-related hippocampal theta field activity and theta-on/theta-off cell discharges, Hippocampus 13:38-47, 2003.

Downloaded from ClinicalKey.com at Buddhist Tzu CFor personal use only. No other uses without permission. C

References 637.e5

229. Madler C, Keller I, Schwender D, Poeppel E: Sensory informa-tion processing during general anaesthesia: effect of isoflurane on auditory evoked neuronal oscillations, Br J Anaesth 66:81-87, 1991.

230. Munglani R, Andrade J, Sapsford DJ, et al: A measure of conscious-ness and memory during isoflurane administration: the coherent frequency, Br J Anaesth 71:633-641, 1993.

231. Buzsaki G, Wang X-J: Mechanisms of gamma oscillations, Annu Rev Neurosci 35:203-225, 2012.

232. Dickinson R, Awaiz S, Whittington MA, et al: The effects of gen-eral anaesthetics on carbachol-evoked gamma oscillations in the rat hippocampus in vitro, Neuropharmacology 44:864-872, 2003.

233. Antkowiak B, Hentschke H: Cellular mechanisms of gamma rhythms in rat neocortical brain slices probed by the volatile anaesthetic isoflurane, Neurosci Lett 231:87-90, 1997.

234. Szarecka A, Xu Y, Tang P: Dynamics of firefly luciferase inhibition by general anesthetics: gaussian and anisotropic network analy-ses, Biophys J 93:1895-1905, 2007.

235. Gottschalk A, Haney P: Computational aspects of anesthetic action in simple neural models, Anesthesiology 98:548-564, 2003.

236. Storer KP, Reeke GN: γ-Aminobutyric acid receptor type A recep-tor potentiation reduces firing of neuronal assemblies in a compu-tational cortical model, Anesthesiology 117:780-790, 2012.

237. Wilson MT, Sleigh JW, Steyn-Ross DA, et al: General anesthetic-induced seizures can be explained by a mean-field model of corti-cal dynamics, Anesthesiology 104:588-593, 2006.

238. Steyn-Ross ML, Steyn-Ross DA, Sleigh JW: Modelling general anaesthesia as a first-order phase transition in the cortex, Prog Bio-phys Mol Biol 85:369-385, 2004.

239. Zhang Y, Stabernack C, Sonner J, et al: Both cerebral GABA(A) receptors and spinal GABA(A) receptors modulate the capacity of isoflurane to produce immobility, Anesth Analg 92:1585-1589, 2001.

240. Zhang Y, Wu S, Eger EI 2nd, et al: Neither GABA(A) nor strych-nine-sensitive glycine receptors are the sole mediators of MAC for isoflurane, Anesth Analg 92:123-127, 2001.

241. Eger EI, Liao M, Laster MJ, et al: Contrasting roles of the N-methyl-D-aspartate receptor in the production of immobilization by con-ventional and aromatic anesthetics, Anesth Analg 102:1397-1406, 2006.

242. Sukhotinsky I, Zalkind V, Lu J, et al: Neural pathways associated with loss of consciousness caused by intracerebral microinjec-tion of GABA A-active anesthetics, Eur J Neurosci 25:1417-1436, 2007.

243. Eckenhoff RG, Xi J, et al: Azi-isoflurane, a photolabel analog of the commonly used inhaled general anesthetic isoflurane, ACS Chem Neurosci 1:139-145, 2010.

244. Nash HA: In vivo genetics of anaesthetic action, Br J Anaesth 89:143-155, 2002.

245. Sonner JM, Cascio M, Xing Y, et al: Alpha 1 subunit–containing GABA type A receptors in forebrain contribute to the effect of inhaled anesthetics on conditioned fear, Mol Pharmacol 68:61-68, 2005.

246. Borghese CM, Xiong W, Oh SI, et al: Mutations M287L and Q266I in the glycine receptor α1 subunit change sensitivity to volatile anesthetics in oocytes and neurons, but not the minimal alveo-lar concentration in knockin mice, Anesthesiology 117:765-771, 2012.

247. Sedensky MM, Siefker JM, Morgan PG: Model organisms: new insights into ion channel and transporter function: stomatin homologues interact in Caenorhabditis elegans, Am J Physiol Cell Physiol 280:C1340-C1348, 2001.

248. Kayser EB, Morgan PG, Sedensky MM: GAS-1: a mitochondrial protein controls sensitivity to volatile anesthetics in the nema-tode Caenorhabditis elegans, Anesthesiology 90:545-554, 1999.

249. Cascio M, Xing Y, Gong D, et al: Mouse chromosome 7 harbors a quantitative trait locus for isoflurane minimum alveolar concen-tration, Anesth Analg 105:381-385, 2007.

250. Veselis RA, Reinsel RA, Feshchenko VA, et al: A neuroanatomi-cal construct for the amnesic effects of propofol, Anesthesiology 97:329-337, 2002.

251. Alkire MT, Miller J: General anesthesia and the neural correlates of consciousness, Prog Brain Res 150:229-244, 2005.

hi General Hospital JC September 17, 2016.opyright ©2016. Elsevier Inc. All rights reserved.