Abstract
General anesthetics depress excitatory and/or enhance inhibitory synaptic transmission principally by modulating the function of glutamatergic or GABAergic synapses, respectively, with relative anesthetic agent-specific mechanisms. Synaptic signaling proteins, including ligand- and voltage-gated ion channels, are targeted by general anesthetics to modulate various synaptic mechanisms, including presynaptic neurotransmitter release, postsynaptic receptor signaling, and dendritic spine dynamics to produce their characteristic acute neurophysiological effects. As synaptic structure and plasticity mediate higher-order functions such as learning and memory, long-term synaptic dysfunction following anesthesia may lead to undesirable neurocognitive consequences depending on the specific anesthetic agent and the vulnerability of the population. Here we review the cellular and molecular mechanisms of transient and persistent general anesthetic alterations of synaptic transmission and plasticity.
Keywords: Anesthesia, synaptic plasticity, presynaptic function, postsynaptic structure, ion channels, synaptic transmission
1. INTRODUCTION
The first public demonstration of a surgical procedure with general anesthesia was performed in 1846 using sulfuric ether. Since that time, various chemically diverse structures have been identified to have potent and selective effects on neuronal transmission to produce the cardinal features of general anesthesia: amnesia, unconsciousness, and immobility. General anesthetic agents are categorized based on whether they are administered by inhalation (e.g., volatile) or intravenously, but their chemical structures, molecular targets, and binding sites are quite diverse. Although mechanistically distinct, general anesthetics depress fast excitatory and/or enhance fast inhibitory synaptic transmission mediated primarily by glutamate and GABA, respectively [1, 2]. The relative importance of anesthetic effects on excitatory vs inhibitory synapses to potentiate overall synaptic inhibition varies among anesthetic agents [3]: modulation of postsynaptic NMDA or postsynaptic and extrasynaptic GABAA receptors is a major contributor to the effects of intravenous anesthetics like ketamine or propofol, respectively; presynaptic and postsynaptic inhibition of excitatory glutamatergic transmission contributes to the depressant effects of volatile anesthetics [3-7]. The mechanisms for the acute anti-excitatory effects of volatile anesthetics include depression of neuronal excitability [8] or action potential conduction [9-12], inhibition of Ca2+ influx [13, 14] and synaptic vesicle exocytosis [15-17], and/or blockade of postsynaptic glutamate receptors [18]. Excitatory glutamate receptors are largely targeted to dendritic spines [19]; their identification as a cellular substrate for anesthetic action [20, 21] highlights their influence on advanced functions such as synaptic plasticity, learning, memory and suggests a role for spine plasticity in the acute and enduring neurocognitive effects of general anesthetics. Here we highlight some of the critical targets involved in the modulation of synaptic transmission and plasticity and the neurophysiological effects of general anesthetics as well as their role in lasting cognitive changes.
Unless indicated otherwise, the agent-specific cellular and pharmacological differences described here occur at clinical concentrations of anesthetics, defined as concentrations measured in vivo required for therapeutic endpoints of general anesthesia.
2. PRESYNAPTIC ANESTHETIC EFFECTS
2.1. Neuronal Excitability
2.1.1. Sodium Channels
Volatile anesthetics reduce excitatory postsynaptic potentials (EPSPs) primarily through presynaptic actions [3, 5, 22]. Reports in the 1970s and 1980s implicated the effects of volatile anesthetics on modulation of protein function, shifting attention away from direct interactions with the lipid bilayer [23]. In fact, clinically relevant concentrations of various general anesthetics only minimally affect lipid bilayer properties [24]. Voltage-gated Na+ channels (Nav) control neuronal excitability, action-potential (AP) driven Ca2+ influx, and Ca2+-dependent neurotransmitter release [25, 26]. At least three states have been identified for neuronal Nav depending on membrane potential: resting (closed), activated (open), and inactivated [27]. Upon neuronal depolarization, sodium channels rapidly activate and initiate an AP, followed by fast inactivation and return of baseline sodium conductance within a few milliseconds [28]. Both local and volatile anesthetic exhibit a voltage- and frequency-dependent block of Nav [29-31]. Volatile anesthetics reduce peak Na+ currents in two ways: 1) preferential interaction with inactivated states of Nav to shift steady-state inactivation toward more negative membrane potentials, reducing channel availability and slowing recovery from fast- inactivation, and 2) interaction with the open and/or resting state to produce tonic block [30, 31]. Substantial neurochemical and electrophysiological evidence supports direct inhibition of presynaptic Nav and subsequent depression of APs and nerve terminal depolarization by volatile anesthetics in heterologous expression systems [32] as well as in more physiologically relevant neuronal preparations, including isolated nerve terminals from rat cerebral cortex [33], rat neurohypophysis, and giant calyx of Held slices [11, 12, 34]. Alternatively, prolonged depolarization or repetitive depolarization drives Nav into a distinct slow-inactivated state from which recovery is very slow [35, 36]; slowly inactivating Nav or resurgent sodium currents that occur with repolarization can enhance repetitive firing and modulate overall neuronal excitability as opposed to AP initiation and propagation [27, 37]. In acute mouse brain slices, these persistent neuronal sodium currents are also inhibited by isoflurane to reduce hippocampal pyramidal neuron excitability [38].
Nav blockade may also contribute to the actions of propofol, a structurally distinct intravenous anesthetic that attenuates increases in Na+ flux, intracellular free Na+ levels and Na+ channel-dependent glutamate release in isolated rat cerebrocortical nerve terminals [39] and calyx of Held [40]. Like volatile anesthetics, propofol primarily enhances inactivation of Nav, with some contribution of reduced activation at higher concentrations in neurohypophysial nerve terminals [34] and isolated nerve terminals [41]. Similarly, higher than clinically relevant plasma concentrations of ketamine, another commonly used intravenous anesthetic, also inhibit Nav conductance in human neuroblastoma cells [42] and isolated nerve terminals prepared from human cortical tissue [43].
Nav channels consist of a large, pore-forming α subunit in association with auxiliary β subunits [44, 45]. Of the nine subtypes of α subunit (Nav1.1-Nav1.9), Nav1.1, Nav1.2, Nav1.3, and Nav1.6 are highly expressed in the central nervous system [46, 47], with Nav1.3 preferentially expressed during development [44]. Multiple expression systems have identified neuronal Nav subtype- and agent-specific effects of inhaled anesthetics to reduce peak Nav current at clinically relevant concentrations in a voltage-dependent manner. At physiologic holding potentials, isoflurane induces fast inactivation and inhibition of peak Na+ currents of Nav1.2 and Nav1.6 compared with Nav1.1 expressed in a mammalian cell line [48, 49]. Reduced sensitivity of Nav1.1 to isoflurane is consistent with previous findings reported in Chinese hamster ovary cells [31] and Xenopus oocytes [50]. Distinct gating properties contribute to these sensitivity differences as similar voltage-dependence of activation, but a positive shift in the voltage-dependence of inactivation was revealed for Nav1.1 compared to Nav1.2 and Nav1.6 [49]. In contrast, suppression of Nav1.2 (50-70%) by propofol in various expression systems requires supratherapeutic concentrations [34, 50, 51]. Furthermore, Nav subtypes show subcellular-, regional-, and neurotransmitter-selective expression [52-55], and distinct sensitivity differences may further contribute to selective anesthetic effects on synaptic transmission.
Regulation of neuronal excitability by Nav supports their roles in the behavioral endpoints of anesthesia. Intrathecal and intravenous administrations of Nav blockers in rats and humans increase the potencies of isoflurane, halothane, or sevoflurane [56-59], while drugs that activate Nav antagonize general anesthesia as reflected in increased MAC (minimum alveolar concentration of anesthetic that eliminates movement in response to noxious stimulation in 50% of subjects) and reduced potency [59-61]. This finding is consistent with a role for Nav as a mediator of immobility produced by anesthetics. Additional reports have also observed reduced activity of a specific Nav to increased sensitivity to volatile anesthetics in transgenic mice. Mutant mice with reduced Nav1.6 activity exhibit reduced theta power in the waking state, suggestive of decreased arousal and neuronal excitability [62]. These mice exhibit increased sensitivity to isoflurane and sevoflurane during induction of unconsciousness [62], exemplifying how reductions in AP conduction by anesthetics at the neuronal level leads to alterations in oscillatory activity at the neural network level [63], influencing disruptions in signal propagation and cortical communication involved in higher-order functions like consciousness which rely on precisely tuned integration of inputs [64].
Nav function is also regulated by second messenger-mediated protein phosphorylation, an important neuromodulation mechanism and an additional target for volatile anesthetics. Phosphorylation of Nav by protein kinase A (PKA) and protein kinase C (PKC) reduces Na+ channel activity by altering channel kinetics [65-67] with differential subtype sensitivity: Nav1.2 is more sensitive to PKA and PKC modulation compared to Nav1.6 [68, 69]. In turn, anesthetics can mediate PKC phosphorylation to indirectly affect Nav inhibition as brain PKC activity is increased by halothane and/or propofol purified from rat brain or in isolated nerve terminals [70, 71].
2.2. Calcium Dynamics
2.2.1. Calcium Channels
Presynaptic voltage-gated Ca2+ channels (Cav) are essential to neurotransmission by mediating Ca2+ influx to trigger synaptic vesicle exocytosis. Over the course of an AP, membrane Cav open at the presynaptic terminal, allowing Ca2+ entry. The increase in Ca2+ influx, along with efflux from store-operated Ca2+ channels (SOCs), increases intracellular free calcium concentrations regulating crucial second messenger-mediated biochemical processes, including neurotransmitter release [72, 73]. In the CNS, there are five major classes of Ca2+ currents: L-, N-, P/Q-, R-, and T- type, with the latter characterized as low voltage-activated channels (LVA) based on the degree of membrane depolarization required for activation [74, 75]. This section will largely focus on the function of presynaptic Cav (N-, P/Q-, R-type currents) in anesthetic effects.
As ion channels are a principal target of general anesthetics, inhibition of Cav to reduce AP-evoked exocytosis is a plausible action of general anesthetics [76, 77]. However, early observations of anesthetic inhibition of synaptic Cav were often conflicting. Volatile anesthetics such as halothane were found to increase resting cytoplasmic Ca2+ in isolated mouse brain nerve terminals [78] and hippocampal brain slices [79], but likely resulted in depressed neuronal excitability via potassium channel- [80] or GABAA-mediated chloride conductance [79] as later work observed significant reductions in the frequency of miniature excitatory postsynaptic currents (mEPSCs) and miniature inhibitory postsynaptic currents (mIPSCs) [81]. With electrical or chemical depolarization, halothane, enflurane, and isoflurane cause a marked and reversible suppression of inward Ca2+ current and synaptic excitation in rat hippocampal slices [82] and cultured neurons [83, 84]. Intracellular Ca2+ dynamics are further perturbated due to delayed clearance by volatile anesthetics. Clinical concentrations of halothane not only reduce AP amplitude, but also prolong the repolarization phase 2-to 4-fold by inhibition of plasma membrane Ca2+-ATPase, an ion pump that ejects Ca2+ from the cell following influx into the cytoplasm [85]. Impaired synaptic transmission contributed by Ca2+-dependent presynaptic mechanisms is also produced by supratherapeutic concentrations of propofol as inhibition of EPSCs due to reduced Ca2+ influx was observed by an increased paired-pulse ratio (PPR) and prolonged slowing of endocytosis [40]. Voltage-gated Cav are hetero-multimeric protein complexes composed of α1, α2-δ, β, and γ subunits, with the α1 subunit containing an ion-conducting pore, voltage sensor, gating machinery and drug-binding sites that determine channel subtype [44]. There are three major presynaptic Cav family subtypes: P/Q- (Cav2.1), N- (Cav2.2), and R-type (Cav-2.3), with P/Q- and N-types primarily coupled to neurotransmitter release with distinct physiological and pharmacological characteristics [86-88]. Anesthetics alter Ca2+ currents contributed by multiple subtypes of Cav, but detailed analysis of the influence on individual channel types was inconclusive in multiple physiological systems [89, 90]. For example, P-type channels, named from their initial discovery in cerebellar Purkinje cells, were found to be insensitive to inhibition by volatile (halothane, isoflurane) and intravenous (propofol, thiopental) anesthetics [91], although they contribute to ~80% of voltage-gated Ca2+ uptake in nerve terminals based on inhibition by ω-Aga-IVA, a specific P-channel blocker [92, 93]. However, these studies were conducted in dissociated Purkinje neurons and the possibility of heterogenous subpopulations of Cav could not be ruled out [91]. Separate expression of specific neuronal Cav in Xenopus oocytes in later studies revealed inhibition of P/Q- and N-mediated Ca2+ currents by both halothane and isoflurane in a concentration dependent manner via steady-state inactivation [94], an effect further modulated by PKC translocation and activation [95]. Further examination of excitatory neurotransmitter release by specific Cav blockers in rat hippocampal neurons revealed a larger contribution by P/Q-type compared to N-type currents, but again with no significant differences in their functional sensitivities to isoflurane [14]. Behaviorally, N-type channels participate in both excitatory and inhibitory synaptic transmission which are sensitive to the actions of anesthetics. Mutant mice lacking Cav2.2 show increased sensitivities to halothane-induced immobility and hypnosis consistent with reduced field excitatory postsynaptic potentials (fEPSPs) recorded from Schaffer collateral CA1 synapses, while displaying decreased sensitivity to propofol possibly due to reduced GABA release from inhibitory presynaptic terminals [96]. Similar findings were also reported for Cav2.3 [96], a subtype that contributes to transmitter release with lower efficacy compared to Cav2.1 and Cav2.2 [97].
2.2.2. Intracellular Calcium Stores
With critical influence over cytoplasmic free Ca2+, intracellular Ca2+ stores provide an additional target for anesthetics. The endoplasmic reticulum (ER) is the largest intracellular store of releasable Ca2+ [98], regulating intracellular Ca2+ critical to neuronal excitability [99] via Ca2+ influx pumps [sarcoplasmic/ER Ca2+ ATPase (SERCA)] and receptor- gated Ca2+ efflux channels [inositol 1,4,5-triphosphate receptors (IP3R), ryanodine receptors (RyR)] [100-102]. Axonal ER Ca2+ controls presynaptic intracellular Ca2+ through Ca2+ sensing proteins and decreased ER Ca2+ is linked to reduced presynaptic Ca2+ influx [103], providing possible targets of anesthetic action; efflux of Ca2+ via IP3R or RyR have been implicated in sevoflurane- or halothane-induced increase in intracellular Ca2+, respectively [104, 105]. Furthermore, volatile anesthetics have several significant side effects, with some individuals developing malignant hyperthermia (MH), a potentially fatal pharmacogenetic disorder triggered by uncontrolled Ca2+ release from the SR mediated by RyR1 [106, 107].
2.3. Synaptic Vesicle Exocytosis
2.3.1. Small Vesicle Exocytosis
Neurotransmitter release is primarily determined by Ca2+ entering the bouton [108]. General anesthetics differentially act on presynaptic Ca2+ via actions on ion channels or vesicle fusion mechanisms to inhibit evoked synaptic vesicle (SV) exocytosis [109, 110] and reduce EPSCs [11] associated directly with attenuated release probability and a number of functional release sites [81, 111]. AP-evoked depolarization can be pharmacologically mimicked by 4- aminopyridine (4AP) and veratridine, a K+ channel blocker and a Nav agonist, respectively, while elevated K+ elicits Nav-independent depolarization [25]. Although one study reported that isoflurane, enflurane, and halothane can directly act on Cav in isolated nerve terminals to decrease glutamate release [76], Cav are relatively insensitive to isoflurane compared to Nav [11, 91]. Considerable evidence supports inhibition of Nav as a major contributor to the presynaptic effects of anesthetics on veratridine-4AP-evoked glutamate release from nerve terminals isolated from several species and brain regions, including from rat, mouse, or guinea pig cerebral cortex and rat striatum and hippocampus [110, 112]. These findings are consistent with a target upstream of Ca2+ entry as they were not reproduced for release evoked by elevated K+, a mechanism that directly opens Cav [113]. Inhibition of evoked GABA release also occurs for isoflurane [13, 114], but is balanced by potentiation of GABAA receptors and increased asynchronous GABA release [115], leading overall to enhanced net inhibition [3, 116]. Moreover, isoflurane inhibits Nav-dependent glutamate release with greater potency than GABA release from cerebral cortex, striatum, and hippocampus [15, 16, 113]. Consistent with selective inhibition of excitatory synaptic transmission, isoflurane inhibition of Ca2+ influx was also greater in glutamatergic compared with GABAergic boutons assayed using genetically encoded Ca2+ biosensors [13]. These effects are not dependent on differential sensitivities to isoflurane or variable expression or coupling of presynaptic Cav subtypes to synaptic vesicle exocytosis [14]. Neurotransmitter release is supra-linearly dependent on presynaptic Ca2+ influx due to the highly cooperative binding of Ca2+ to synaptotagmin 1, the principal neuronal Ca2+ sensor for triggering synaptic vesicle fusion [117, 118]. Clinical concentrations of isoflurane inhibit single AP-evoked glutamate exocytosis; lowering external Ca2+ to mimic the isoflurane-induced reduction in Ca2+ leads to an equivalent reduction in exocytosis, suggesting that anesthetic inhibition of neurotransmitter release occurs primarily through reduced axon terminal Ca2+ entry without significant direct effects on Ca2+-exocytosis coupling or on the vesicle fusion machinery [13].
The diversity of GABAergic interneurons makes comparing differences in presynaptic anesthetic pharmacology between glutamatergic and GABAergic neurons difficult, as these interneurons vary in their expression of key neurotransmission-related proteins [119] including ion channels, Ca2+-binding proteins involved in modulating intracellular Ca2+, and intracellular signaling proteins [120]. For example, hippocampal parvalbumin (PV+) interneurons are distinguished by a supra-critical density of Nav1.1 at distal and terminal axons thought to enable their fast-spiking phenotype [121], while excitatory pyramidal neurons express more Nav1.2 and Nav1.6 [53, 54, 122]. Expression of presynaptic Nav subtypes coupled to neurotransmitter release not only differs between transmitter types, but also between CNS regions in a nerve terminal-specific manner [113, 123] as inhibition of Nav-dependent glutamate release occurs with ~50% greater potency than inhibition of GABA release in cortical, striatal, and hippocampal nerve terminals [123]. Differential pre- and post-synaptic expression of Nav subtypes in rat hippocampus [55] further contributes to variations in neurotransmitter suppression as Nav subtypes are differentially inhibited by volatile anesthetics Nav1.1< Nav1.2/1.6; [49], determining cell-specific inhibition of synaptic vesicle exocytosis. For example, Nav1.1 is less sensitive to isoflurane than Nav1.2 and Nav1.6 are due to its unique gating properties [49], suggesting that Nav1.1 expression in GABAergic boutons underlies their reduced sensitivity to isoflurane inhibition. Like volatile anesthetics, propofol, thiopental, and ketamine also reduce glutamate release from rat cortical slices, while GABA release was increased via an Nav-independent mechanism by propofol, etomidate, or pentobarbital [115, 124].
2.3.2. Large Vesicle Exocytosis
Monoamine neurotransmitters such as dopamine are packaged into both small synaptic vesicles and large dense core vesicles (LDCV) for release. However, only small synaptic vesicles localize to synaptic boutons, while LDCVs engage primarily in extrasynaptic exocytosis [125, 126]. Aminergic neurons have distinct mechanisms of LDCV exocytosis [127] and play central roles in wakefulness and arousal [128]. Stimulation of dopaminergic neurons in the rat ventral tegmental area (VTA), a principal midbrain dopaminergic region [129], induces emergence from isoflurane in rats [130]. Conversely, inhibitors of the dopamine transporter restore conscious behaviors in rats anesthetized with isoflurane, propofol, or sevoflurane [131-133]. Early reports on basal and chemically evoked dopamine release by various anesthetics have been inconsistent, showing increased or decreased release based on the type of anesthetic agent and stimulation [134-136]. Recent work using fluorescence imaging, under electrically evoked action potentials in cultured rat VTA neurons, show that isoflurane inhibits exocytosis in dopaminergic neurons with intermediate potency compared to glutamate and GABA synaptic vesicle exocytosis [17]. Although considerable evidence supports that presynaptic Ca2+ channels are not the principal targets involved in inhibition of glutamate and GABA release by volatile anesthetics [123, 137], exocytosis in dopaminergic neurons evoked by elevated KCl is inhibited, a mechanism mediated exclusively by reduced Ca2+ entry through both Cav2.1 (P/Q-type) and Cav2.2 (N-type) and is independent of Nav activation [17]. Halothane and isoflurane also alter the presynaptic regulation of dopamine and GABA release mediated by presynaptic acetylcholine receptors in rat striatum, suggesting that cholinergic transmission is a presynaptic target for volatile anesthetics [138].
Brain-derived neurotrophic factor (BDNF), a secreted growth factor packaged in LDCV, plays a critical role in modulating excitatory synaptic transmission. Activity-dependent presynaptic BDNF signaling via its receptor, TrkB, potentiates glutamate release by multiple mechanisms: activation of the phospholipase Cγ (PLCγ) pathway increases intracellular Ca2+ [139]; phosphorylation of synapsin 1 by mitogen-activated protein kinase (MAPK) and increased Rab3 expression act directly on vesicle machinery [140, 141]; and activation of an actin motor complex increases neurotransmitter replenishment and release [142, 143]. Synapses in BDNF knockout mice have fewer docked vesicles [144] and reduced probability of neurotransmitter release [145]. Based on stimulation frequency, BDNF can be released from axons or dendrites [146, 147] and have diverse pre- and post-synaptic modulatory effects on mature glutamatergic synapses, indicating a plausible presynaptic target for volatile anesthetics. BDNF release evoked by tetanic high-frequency stimulation (16 bursts of 50 action potentials @ 50 Hz) was significantly inhibited by clinical concentrations of isoflurane (unpublished data), suggesting a novel contribution to isoflurane-induced reductions of excitatory transmission.
2.4. Additional Targets
2.4.1. Potassium and Hyperpolarization-activated and Cyclic Nucleotide-gated Channels
Other anesthetic-sensitive ion channels or presynaptic proteins critical to CNS excitability and synaptic transmission include two-pore domain K+ channels (K2P), hyperpolarization-activated and cyclic nucleotide-gated (HCN) channels, and SNARE proteins. Potassium channels regulate the passive flow of K+ ions to maintain the membrane potential following changes in transmembrane potential. Out of the four subfamilies of potassium channels, K2P and Kv channels are modulated by general anesthetics [57, 148-150]. K2P channels are widely expressed in the nervous system and contribute to background membrane conductance [151]. Electrophysiological measurements show that isoflurane opens K2P channels leading to reduced neuronal excitability [152, 153], altering network oscillations [154], underlying loss of consciousness and immobilization [155-158]. HCN channels also contribute to sensitivity to general anesthetics [159-161], as slow-wave oscillations within the thalamocortical circuits, indicative of the anesthetized state, are associated with isoflurane-mediated inhibition of HCN channels [162]. Induction of both anterograde amnesia and hypnosis under isoflurane and sevoflurane are attenuated in HCN1 knockout mice [160]. Comparably, K2P and HCN channels are also targeted by propofol and ketamine to mediate anesthetic endpoints [163-165].
2.4.2. Synaptic Vesicle Machinery
Downstream of ion channel regulation, the synaptic vesicle release machinery itself may also be involved in anesthetic inhibition of exocytosis. Recent findings support distinct dual mechanisms based on the frequency of input. With short depolarizing pulses in calyx of Held slices, isoflurane inhibits exocytosis by reducing Ca2+ influx without altering Ca2+-exocytosis coupling as previously reported in hippocampal neurons [13, 111], whereas with long presynaptic depolarizations associated with greater exocytosis, isoflurane directly inhibits exocytic machinery downstream of Ca2+ influx [111]. Moreover, in simultaneous recordings of pre- and post-synaptic APs, as well as in unit recordings from cerebral cortical neurons in mice in vivo, isoflurane preferentially inhibits monosynaptic transmission evoked by higher-frequency stimulation [111], further suggesting a critical role for exocytic machinery.
The SNARE complex, comprised of syntaxin 1, SNAP-25/23, and synaptobrevin [166, 167] regulates synaptic vesicle docking, priming, and fusion [168]. Mutations in the syntaxin homolog in Caenorhabaditis elegans confer resistance to the behavioral effects of isoflurane and halothane [169], results that are recapitulated in Drosophila melanogaster [170]. In rodent model systems, clinical concentrations of isoflurane and halothane directly bind rat synaptic SNARE proteins [171] and inhibit neurotransmitter release in hippocampal neurons, an effect abolished by overexpression of syntaxin 1 [172] or knockdown of SNAP-25/23 [173]. Interference with SNARE complex formation by direct interaction with syntaxin 1 [174] and inhibition of neurotransmitter release were also observed at clinical concentrations of propofol and etomidate in neurosecretory cells and rat hippocampal neurons [175], although these effects occur at higher anesthetic concentrations.
2.4.3. Gap Junctions
Gap junctions provide electrical coupling of neurons and glial cells to mediate direct cell-to-cell communication and synchronized activation of cellular networks [176]. Volatile anesthetics like halothane, enflurane, sevoflurane, and isoflurane at high supratherapeutic concentrations inhibit gap-junction-mediated neural activities by reducing gap junction conductance [177] and permeability in vitro [178, 179], and contribute to anesthetic-induced immobilization in vivo [180]. These effects are largely observed under higher than clinical concentrations, unlike depression of intercellular communication via gap junctions by intravenous anesthetics like propofol and thiopental [181]. Preclinically, based on the loss or recovery of righting reflex, used as a proxy for conscious states, genetic deletion of connexin 36, a gap-junction protein, increases sensitivity to the hypnotic effects of both isoflurane and propofol [182]. These studies indicate that desynchronization of saltatory conduction between clustered neurons contributes to reduced overall cortical network excitation by anesthetics.
3. POSTSYNAPTIC ANESTHETIC EFFECTS
The neurophysiological effects of general anesthetics involving postsynaptic modulation are largely mediated through ligand-gated ion channels that play essential physiological roles in inhibitory and excitatory synaptic transmission: GABAA, nicotinic acetylcholine, and AMPA- and NMDA-type glutamate receptors [77, 183]. Glutamate receptors are anchored by actin filaments, mostly in dendritic spines, postsynaptic structures that compartmentalize biochemical and cell biological processes critical for excitatory synaptic transmission and plasticity [19, 184, 185]. Transient alterations in spine structure and function and the underlying actin dynamics also provide plausible mechanisms for acute anesthetic action [20, 21, 186].
3.1. Receptor Signaling
3.1.1. γ-Aminobutyric Acid Receptors
Considerable evidence indicates that the neurodepressive actions of general anesthetics involve enhancement of inhibitory synaptic transmission [1, 187-191]. GABA, the major inhibitory neurotransmitter in the mammalian brain, activates two distinct receptor subtypes: GABAA receptors, pentamers composed of multiple subunits that form a Cl--selective ion channel; and GABAB receptors, single subunit receptors that couple to G-proteins and modulate K+ and Ca2+ channels [192]. GABAA receptors mediate both phasic synaptic (fast) and tonic extrasynaptic (slow) inhibitory transmission; they are sites of action for most general anesthetics [116, 193, 194].
The halogenated ethers enflurane, halothane, and isoflurane [195-198] as well as the intravenous anesthetics propofol [199] and etomidate [200] all enhance GABAA receptor function at clinically relevant concentrations shown by enhancement of GABAA-mediated currents and Cl- flux in cells, brain slices and brain homogenates (reviewed in [201, 202]). Intracerebroventricular picrotoxin, a GABAA receptor antagonist, increases isoflurane MAC, requiring higher doses to achieve a surgical plane [203]. GABAA receptor modulation by propofol or pentobarbital has distinct dose-dependent effects likely involving multiple sites of action [204]; clinical concentrations of propofol potentiate GABA-activated currents, increase open channel frequency, and reduce the rate of desensitization, while intermediate concentrations directly activate GABAA channels, and even higher concentrations inhibit receptor function [205-209]. Electrophysiological and radioligand binding experiments suggest that etomidate shares this ability to interact with GABAA receptors [210-212], potentiating synaptic and extrasynaptic GABAA- mediated currents by increased channel open time and probability of channel opening [213, 214]. Similarly, most volatile anesthetics enhance the amplitude of GABAA-mediated currents and prolong synaptic inhibition [215] by direct effects on channel gating, as suggested by the use of partial agonists [216]. Heterogenous effects on miniature-IPSCs and evoked-IPSCs suggest a larger contribution by extrasynaptic receptors [194, 217]. Modulation of extrasynaptic GABAA-mediated tonic conductance by volatile anesthetics was first demonstrated in CA1 pyramidal neurons [218], followed by single synapse preparations from CNS slices [215].
GABAA receptors are pentamers formed by different glycoprotein subunits (α1-6, β1-3, γ1-3, δ, ρ1-3, and ε, φ, π), providing numerous potential combinations with distinct pharmacological profiles [219-221] that show variable sensitivity to allosteric modulators [222]. Subunit requirements for the formation of functional recombinant GABAA receptors in cell lines and Xenopus oocytes are various [223-226], but most GABAA receptors in vivo consist of α, β, and γ subunit complexes [227]. Receptors composed of α1β3 can be expressed in vitro and form functional GABA- activated channels [228-231] that are fully sensitive to general anesthetics [232, 233]. Differences in β subunits may be species-specific; rodent β1 and murine β2 and β3 form homomeric receptors that are selectively opened by pentobarbital, etomidate, and propofol, but are less sensitive to GABA [234-237], unlike human and bovine β1 subunits expressed in oocytes [235, 238], suggesting distinct drug- and neurotransmitter-bound states [239]. Site-directed mutagenesis has identified key subunits involved in the anesthetic modulation of GABAA receptors. A point mutation in the β1 subunit (M286W) abolishes potentiation of GABA effect by propofol, without altering direct activation by higher concentrations [240]. Replacing the asparagine residue at position 265 on either β2 or β3 subunits inhibits the modulatory actions of enflurane and isoflurane [241, 242] as well as both the modulatory and direct effects of etomidate and propofol [243], findings that are further substantiated in mouse models. Cultured hippocampal neurons or cortical brain slices from β2 (N265S) mice show diminished potentiation of GABA-induced Cl- currents by etomidate or less effective reduction of spontaneous firing by etomidate and enflurane, respectively [244]. Direct in vivo measurements of anesthetic endpoints show that selective agonism of receptors containing the β3 subunit contributes to etomidate anesthesia, with both β2 and β3 subunits contributing to sedation [245, 246] and loss of righting reflex (LORR) [244, 245]. Thus, GABAA receptors containing the β2 and β3 subunits are sufficient for most anesthetic endpoints, but potentiation caused by volatile anesthetics are smaller compared to intravenous anesthetics at equi-anesthetic concentrations, suggesting additional targets. Genetically engineered mice that lack the β3 subunit are more resistant to the immobilizing effects of enflurane, isoflurane, and halothane, but not loss of consciousness, suggesting separate anesthetic states may also be mediated by regional GABAA receptor heterogeneity [247-250]. In the neocortex, both α1 and β2 subunit-containing GABAA receptors contribute to the sedative effects of volatile anesthetics [191], including depression of action potential firing correlated with increased GABAergic inhibition by isoflurane, enflurane, and halothane [251].
The GABAA receptor is made up of five subunits, each of which is proposed to contain four transmembrane segments (TM1-4), including a short extracellular loop between TM2 and TM3 that is involved in channel gating [252]. Chimeric studies demonstrate that transmembrane domains of α and β subunits are essential for positive modulation of GABA binding and modulation by anesthetics. The GABA-potentiating action of propofol is influenced by the α1 subunit, particularly the TM2 region [253], while the TM3 domains of the α2 and β2 subunits are associated with binding [254, 255] and increased affinity for etomidate and propofol [240]. In contrast, enhancement of submaximal GABA- activated currents by intravenous anesthetics are eliminated by mutation of the G219 residue in the TM1 region of the β2 subunit [256]. Moreover, residues within TM2 and TM3 of GABAA α2 and β1 subunits are necessary for positive receptor modulation by isoflurane [240, 257, 258] with some evidence of α1 and α4 involvement in isoflurane binding [259] and its hypnotic [260, 261] and amnestic actions [262], respectively. In the dentate gyrus, levels of α1 and γ2 subunits are higher at the synapse compared to α4 and δ subunits that are expressed more extrasynaptically [263], suggesting further regional and cellular specificity as the γ2 subunit can influence pharmacologic sensitivity as effects of isoflurane on peak amplitude and on the kinetics of deactivation and desensitization are substantially reduced in γ2-containing receptors [264].
Impaired memory is a potent effect of anesthetics and was initially thought to involve tonic inhibitory conductance generated primarily by α5 subunit-containing GABAA receptors [265]; hippocampal slices and primary hippocampal neurons exhibit a tonic conductance generated by α5 GABAA receptors that is enhanced by several classes of anesthetics, including propofol [116], isoflurane [265], and etomidate [266]. Subsequent studies support differential modulation of both synaptic and extrasynaptic GABAA receptors in effects of anesthetic on memory [267-269]. In hippocampal slices, long-term potentiation, normally impaired by etomidate, remains intact in mice with a null mutation of the α5 subunit compared to wild-type littermates [266]. Correspondingly, wild-type mice exhibit increased sensitivity to the amnestic effects of etomidate as well as impaired performance in hippocampal-dependent learning and memory tasks compared to α5-subunit-deficient mice [266, 270].
3.1.2. Nicotinic Acetylcholine Receptors
GABAA receptors are members of the pentameric ligand-gated ion channel superfamily that also includes muscle- and neuronal-type nicotinic acetylcholine receptors (nAChR; [271]. Due to their greater accessibility, initial studies focused on anesthetic inhibition of muscle-type nAChRs at the neuromuscular junction, a target of local anesthetics [272]. However, neuronal nAChRs that play a crucial role in synaptic transmission, both as postsynaptic mediators of fast synaptic responses and at presynaptic sites, were later found to have greater anesthetic sensitivity, particularly to volatile anesthetics [273]. Neuronal receptors are expressed as homomeric or heteromeric complexes [274, 275] with repeating α7-α9 subunits or heteromers of two or three α and β subunits in different stoichiometries, respectively [275, 276]. The different subunit combinations enable functional diversity in ion conductance, selectivity, and kinetics in both native and heterologous systems [275, 277].
Initially, the effects of general anesthetics were tested on nicotinic receptors in mollusks [278, 279] and bovine adrenal chromaffin cells [280, 281], which showed inhibition by inhalational agents. Voltage-clamp studies on acetylcholine- activated currents of rodent α4β2 nAChR expressed in Xenopus oocytes confirmed high sensitivity to inhalational anesthetics including halothane, isoflurane, and sevoflurane at anesthetic and sub-anesthetic doses, with higher concentrations required for inhibition by propofol [273]. Heteromeric human neuronal nAChRs (α2β4, α3β4, α4β2) expressed in oocytes or human embryonic kidney cells were also inhibited by isoflurane, sevoflurane, halothane and ketamine with subunit-specific sensitivities [282-284], largely determined by a single amino acid residue (β2-Val 253; β4-Phe 255) identified by chimeric and single amino acid mutants of nAChR [285]. In contrast, α7 subunits directly interact with anesthetics [286, 287], but homomeric α7 receptors were mostly insensitive to isoflurane or propofol at clinical concentrations, requiring supraclinical doses for inhibition [288, 289]. However, native channels in dissociated rat cortical neurons exhibit inhibition by clinical concentrations of halothane for both α4β2-type and α7-type currents, with lower sensitivity for the latter, by accelerating the decay phase of α7-type currents while slowing decay of α4β2- type currents [290]. Complementary dynamic simulation models further propose multiple putative binding sites for halothane interaction with α4β2 nAChR in the open confirmation [291]. nAChR modulation by various anesthetics possibly contributes to analgesia [272], but is not directly involved in producing hypnosis or immobility [292-294]. However, specific loss of cholinergic basal forebrain neurons increases time to emergence from propofol and pentobarbital, and reduces behavioral excitation during halothane exposure [295].
3.1.3. Glutamate Receptors
Two major classes of ionotropic glutamate receptors have been identified: N-methyl-D-aspartate (NMDA) and α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid (AMPA) receptors [296]. The NMDA receptor is blocked in a voltage-dependent manner by physiologic concentrations of Mg2+ and requires binding of both glutamate and the co-agonist glycine for activation, allowing influx of Ca2+ and Na+ [297]. Considerable evidence suggests interactions between anesthetics and NMDA receptors: volatile anesthetic potency is increased by NMDA receptor antagonism [298, 299], and NMDA-stimulated currents are inhibited by isoflurane in cultured hippocampal neurons, decreasing both the frequency of channel opening and channel open time [300]. In rat brain homogenates, disruption of glutamate- stimulated binding of the NMDA pore blocker, MK-801, with reversal by the NMDA receptor positive modulator, glycine, show direct interaction of enflurane and halothane with NMDA receptors and subsequent ion channel inhibition [301, 302]. Volatile anesthetic effects on glutamatergic synaptic transmission are most likely heterosynaptic as both NMDA and AMPA receptor antagonists, administered intrathecally, can reduce the MAC of isoflurane in rats [303, 304], and they preferentially block AMPA over NMDA receptor-mediated currents at high doses of halothane [22, 305, 306].
Unlike volatile anesthetics, ketamine and propofol have opposing effects on NMDA receptors. Ketamine directly blocks NMDA receptors to produce anesthetic and analgesic effects [307-309] by two mechanisms: an open-channel block and closed-channel block characterized by a decrease in open frequency with no change in open time [310]. In contrast, clinical concentrations of propofol do not show significant effects on excitatory transmission based on effects on NMDA and AMPA glutamate receptors expressed in Xenopus oocytes [311] or on the release of glutamate from rat synaptosomes [39, 312]; any direct effects of propofol on NMDA receptors occur only at supratherapeutic concentrations [313].
NMDA receptors are heteromeric protein complexes comprising at least two of seven known subunit types: GluN1, GluN2 (A-D), and GluN3 (A-B), with proper assembly requiring GluN1 and at least one GluN2 subunits [314]. The functional AMPA receptor is usually a homomeric or heteromeric tetrameric complex with various compositions of four different subunits: GluA1-4 [315]. Using recombinant GluN1/GluN2A and GluN1/GluN2B glutamate receptors, [316] showed that clinical concentrations of isoflurane, sevoflurane, and desflurane inhibit NMDA receptors in a reversible, dose-dependent, and voltage-insensitive manner. Comparing wild-type and mutant GluN1/GluN2A receptors extend these findings to show that amino acid substitutions in TM3 of GluN1 and TM4 of GluN2A subunits reduce the action of various anesthetics [317]. Although equi-anesthetic concentrations of most anesthetics inhibit NMDA receptors to some extent [318], genetic deletion of NMDA receptor subunits does not show specificity to ketamine in reducing sensitivity to anesthesia/hypnosis [319] or to volatile anesthetics in producing immobilization [320].
Glutamate receptors are functionally modulated by signaling that regulates their state of phosphorylation and, in turn, activity [321]. Although direct channel effects are lacking, investigation of propofol effects on GluN1 subunit phosphorylation in cultured neurons found reduced phosphorylation and activation of ERK leading to alterations in intracellular Ca2+ and transcription [322-324] at amnestic concentrations of propofol [325]. Moreover, propofol increases phosphorylation of GluA1, although it is unclear whether this contributes to any specific pharmacological action of propofol [326]. Similarly, both in vitro [327, 328] and in vivo studies [329] show alterations of NMDA GluN1 and AMPA GluA1 subunit phosphorylation by ketamine and isoflurane, with a possible role in analgesia [330].
3.2. Dendritic Spine Dynamics
Modulation of neurotransmission by general anesthetics produces both therapeutic and undesirable effects, including neurotoxicity and cognitive dysfunction. Multiple targets have been identified for the neurophysiological effects of general anesthetics (vida supra), but cumulative and downstream effects on synaptic plasticity can lead to acute or delayed cognitive dysfunction [331-333]. Dendritic spines form the postsynaptic contact sites for most excitatory synapses and represent the structural basis of glutamatergic synaptic plasticity [19]. They are of significant interest as a subcellular substrate for anesthetic action and enduring cognitive effects as alterations in spine number and shape are associated with cognitive and developmental dysfunction in various neurological disorders [334].
Spine changes following anesthetic exposure depend on neuronal age: during development, anesthetic exposure can increase or decrease dendritic spine and filopodial density depending on the stage of synaptogenesis [335-338]. Synaptogenesis is a critical period of development, shaping connectivity and regulating the balance between excitatory and inhibitory synaptic function [339]. In the mouse somatosensory cortex and hippocampus, anesthesia with propofol or ketamine increased the density of dendritic spines [335, 340]; similar structural modifications are also observed in rat prefrontal cortex after treatment with isoflurane, sevoflurane, or desflurane [336]. In mature neurons, there is transient and reversible anesthetic-induced disruption of spine structure [20, 21]. Intravital imaging of young adult mouse cortical neurons showed that ketamine/xylazine or isoflurane had no effects on spine formation or elimination but temporarily reduced elimination of filopodia, precursors of established spines [20].
3.2.1. Actin Dynamics
Spine stability and structure are directly regulated by dynamic changes in the actin cytoskeleton that help anchor glutamate receptors to postsynaptic sites [184, 185] and modulate learning and memory [341-343]. Previous studies have implicated filamentous actin in anesthetic effects on neurons and astrocytes [344, 345]. In developing neurons, sevoflurane can reduce the length of filopodia by an actin-dependent mechanism [346], and inhibition of Rho-A activated kinase (ROCK), a kinase involved in actin dynamics, attenuates the effects of high concentrations of ketamine [347]. Similarly, Matus and colleagues [186] demonstrated that volatile anesthetics, including isoflurane and enflurane, reversibly block rapid actin-based spine motility in the spine head (often called “morphing”). Dendritic spines undergo transient reductions in area and number following isoflurane exposure in the hippocampus, effects that are prevented by actin filament stabilization [21].
3.2.2. BDNF Signaling
As a key modulator of synaptic plasticity, the mature form of brain derived neurotrophic factor (mBDNF), cleaved from its precursor proBDNF, regulates spine structure [348, 349]. ProBDNF and mBDNF signal via distinct receptors to mediate divergent actions on neuronal survival, structure, and synaptic plasticity [350, 351]. Cleavage of proBDNF to mBDNF is developmentally regulated; in the early postnatal brain, high levels of proBDNF activate p75 receptors to promote cell death and attenuate synaptic transmission [352, 353]. Quantitative histology and evidence from immunofluorescence microscopy suggest that isoflurane [344], propofol [354] and a cocktail of general and local anesthetics [355] produce neurotoxicity in the developing brain via proBDNF signaling. Upregulating cleavage of proBDNF to mBDNF or direct pharmacological inhibition of p75 receptors attenuates both isoflurane- and propofol-induced neuronal apoptosis [354] and isoflurane-induced destabilization of dendritic filopodia [344]. In contrast, in the adult brain, mBDNF regulates the density and morphology of dendritic spines [348, 349], promoting neuronal survival and enhancing synaptic plasticity via TrKB signaling [356-358]. In vivo studies support a role for down-regulation of mBDNF by anesthetics. In humans, propofol or isoflurane significantly reduces plasma BDNF concentration intraoperatively and 24 h after surgery [359]. Moreover, epigenetic enhancement of BDNF signaling improves cognitive impairments induced by isoflurane in aged rats [360].
Collectively, these studies identify dendritic spines as a structural target for acute anesthetic action on hippocampus- dependent memory, and possibly persistent effects on network function if original connections are not restored.
4. FUNCTIONAL OUTCOMES OF ANESTHETICINDUCED SYNAPTIC PLASTICITY
4.1. Preclinical Assessments of Cognitive Function
4.1.1. Synaptic Potentiation and Depression
Synaptic plasticity is the strengthening or weakening of synapses in response to activity patterns; it involves complex modulation by distinct signaling proteins that alter cellular activity. The two main types of synaptic plasticity are either a persistent decrease [long-term depression (LTD)] or increase [long-term potentiation (LTP)] in synaptic efficiency [361], involving in part Ca2+ entry and depolarization. LTP and LTD are critically dependent on excitatory synaptic transmission through hippocampal dendritic spines, which are modulated during memory consolidation [351] and blocked by anesthetics [362, 363] both acutely [364, 365] and chronically [366-368], and following anesthetic exposure during critical periods of neurodevelopment [369, 370].
In CA1 pyramidal neurons of rodent hippocampal slices, clinical concentrations of halothane, isoflurane, and sevoflurane reduce the probability of LTP induction likely associated with decreased postsynaptic depolarization [362, 371, 372], while ketamine blocks LTP [373] via inhibition of NMDA receptor-mediated Ca2+ influx [374, 375] or AMPA receptor signaling [376]. Isoflurane-induced persistent changes in synaptic strength have been attributed to p otentiation of GABAA and glycine receptor function [269], as well as inhibition of nAChR [377-379], AMPA and NMDA receptors [18, 380, 381]; modulations that contribute to reduced overall glutamatergic transmission and decreased postsynaptic Ca2+ currents [382]. In vivo, depression of LTP induction is recapitulated in electrophysiological recordings of adult rats with exposure to clinical concentrations of sevoflurane [383], although confounding effects of halothane cannot be completely ruled out here. Effects of propofol on synaptic plasticity in the CA1 region are variable and dose-dependent: low doses of propofol enhance the development of LTD and impair maintenance of LTP by an NMDA receptor-dependent mechanism [366], while supratherapeutic concentrations inhibit LTP mediated by GABAA receptors [384-386]. GABAA receptors on non-pyramidal, inhibitory neurons also play a role in impairing LTP by etomidate [387].
4.1.2. Animal Behavior
These functional deficits in synaptic plasticity intersect the cellular and behavioral changes described in preclinical studies [331, 368, 388-391]. Non-human primates are a good animal model due to the similarities in the physiology, pharmacology, metabolism, and reproductive systems to those of humans. Exposure of fetal or neonatal monkeys to isoflurane, ketamine, or propofol for 3, 5, or >5 h causes cortical neuronal cell death [390, 392-397]; these cellular changes are associated with long-lasting deficits in brain function following ketamine [398] and motor, social, and emotional deficits following isoflurane [399, 400].
Exposure to common anesthetic agents in young or adult rodents results in performance impairment in hippocampus- dependent [331, 401, 402] or motor cortex-dependent [403] learning tasks with differences based on type of anesthetic agent [404], the number of exposures [363], and sex [405, 406]. Possible molecular mechanisms include caspase activation [407, 408], suppression of ERK signaling [409-411], increased JNK signaling [412], and reduced PSD-95 expression [413-415] in young rodents, as well as reduced hippocampal acetylcholine levels [416-419] and nACh receptor signaling [420, 421], and elevated calcineurin-mediated neuroinflammation [422] in aged rodents. In contrast, evidence for cognitive changes following general anesthesia in adult rodents is inconclusive and studies have produced mixed results with some showing no persistent changes in memory function [369, 423-427], and others showing impaired hippocampus-dependent learning as assessed by Morris water maze and Barnes maze, weeks following isoflurane exposure [367, 428]. Sustained increases in α5 GABAA receptor activity in the adult mouse brain following a single exposure to isoflurane or etomidate is implicated in lasting impairments in hippocampal memory performance and synaptic plasticity [429-431]. Moreover, in adolescent rats, isoflurane-elicited affective and cognitive deficits are mitigated by blocking extrasynaptic GABAA receptor function [432], consistent with a wider range for the developmental window of susceptibility to enduring effects of general anesthesia. Reduction of BDNF expression, including via epigenetic mechanisms [433], has also been associated in learning and memory dysfunction in adult rats following isoflurane [433, 434], sevoflurane [435], and ketamine [436] administration.
4.1.3. Intersection with Alzheimer’s Disease
Several studies have suggested that perioperative factors, including anesthetics, contribute to Alzheimer’s disease pathogenesis. In various cell lines transfected with Alzheimer’s disease mutant human amyloid precursor proteins (APP), clinically relevant concentrations of isoflurane can induce apoptosis via disruption of calcium homeostasis [437-439], alter APP processing, and increase amyloid beta (Aβ) levels [440, 441]. Accumulation of brain Aβ plaque burden was also recapitulated in vivo in Alzheimer’s disease mice with halothane exposure [442]. Moreover, in wild- type rodents and in Alzheimer’s disease model mice, sevoflurane, isoflurane, or enflurane induced Aβ oligomerization [443], tau hyperphosphorylation and cognitive decline [444-447] mediated by Akt and ERK activation [448, 449].
4.2. Clinical Assessment of Cognitive Function
As major pharmacological modulators of synaptic transmission, anesthetics impair memory and learning in animal studies, but translational relevance and extrapolation to clinical practice remain highly debated. Postoperative cognitive impairments recently termed perioperative neurocognitive disorders (PND; [450]), include delirium and postoperative cognitive dysfunction (POCD), manifested as an acute, confusional state occurring hours to days after surgery, to more durable deficits in memory, attention, concentration, and executive functions, respectively. Although the duration and onset of perioperative neurocognitive disorders are variable, longitudinal studies suggest that they are distinct manifestations of neurocognitive deficits, with variable incidence, triggered by the intersection between surgery, anesthesia, age, preexisting vulnerability, and individual cognitive trajectory [451-456].
4.2.1. Risk Factors for Elderly Populations
Many prospective and retrospective clinical studies have reported perioperative neurocognitive disorders affecting up to 41% of patients older than 60 years [457-459], while other reports did not detect a difference in objectively measured cognitive decline [460, 461]. The elderly are particularly affected but additional risk factors include pre-existing cognitive impairment or disease as well as education level [462, 463], type of surgery [464-468] and specific anesthetic agents and techniques [469]. For example, elderly patients undergoing desflurane or propofol anesthesia exhibit better cognitive function than those undergoing sevoflurane exposure based on better quality of emergence [470-474] and/or lower surgery-induced pro-inflammatory effects [475].
Several studies have explored the possible association between a specific genotype and perioperative neurocognitive disorders; polymorphisms of the human gene C-reactive protein [476], P-selectin [477], and platelet glycoprotein IIIa [478] suggest an additional vulnerability in genotype may increase susceptibility. Specific genetic variations in apolipoprotein E alleles are linked to increases in the risk of perioperative neurocognitive disorders and Alzheimer’s disease [479, 480] as both general anesthesia and Alzheimer’s disease are associated with deficits in cholinergic transmission [481-483] and synapse dysfunction [484]. Human biomarkers of Alzheimer’s disease such as Aβ levels in the cerebrospinal fluid are increased 24 h following surgery under isoflurane anesthesia [485]. However, observational reports have been conflicting. In retrospective studies, no association between the risk of Alzheimer’s disease and exposure to anesthesia in the 1- and 5-years preceding disease onset was found, nor between the risk of Alzheimer’s disease and the number of surgical procedures [486-488]. Furthermore, no conclusive evidence from observational studies, including meta-analysis of case control studies or retrospective cohort studies, shows a link between anesthesia and surgery with the development of clinical dementia [489]. In contrast, a national case control study showed an associated risk between multiple surgeries with general anesthesia and a reduction in the onset of dementia [490]. Collectively, perioperative neurocognitive disorders are multifactorial with various biological and socioeconomic predisposing factors. Clinical observations may be confounded by temporal differences in onset of cognitive impairments, type of standardized testing, as well as inability to differentiate causation by anesthesia, surgery and/or inflammation.
4.2.2. Risk Factors for Juvenile Populations
Juvenile animals are highly susceptible to anesthetic-induced cognitive impairments. There is considerable preclinical evidence describing how general anesthetics alter brain and behavioral development in young animals [491]. Some human studies have found an association between exposure to anesthesia in early childhood and increased risk of poor neurodevelopmental outcome [492-497], while some cohort studies have found no association [498-500].
In a randomized controlled trial comparing the neurodevelopmental outcome in children receiving general or spinal anesthesia (GAS study), exposure to sevoflurane for an average of just under one hour in infancy did not increase the risk of adverse neurodevelopmental outcomes at two or five years of age [501, 502]. The Pediatric Anesthesia Neurodevelopmental Assessment (PANDA) study used a sibling-matched cohort design to test whether a single exposure to general anesthesia (intravenous or volatile) in healthy children younger than 36 months was associated with increased risk of impaired global cognitive function in early childhood (8 to 15 years), also showing no significant differences [503]. However, changes in primary or secondary measures may be dependent on the neuropsychological domain and specific patterns of vulnerability and insult. The Mayo Anesthesia Safely in Kids (MASK) study tested exposure to multiple procedures requiring anesthesia prior to age 3 years and found no associated deficits in general intelligence but did report modest impairments in processing speed and fine motor coordination [504]. These three studies provide strong evidence that a single short exposure to general anesthesia at a young age does not result in measurable alterations in neurodevelopmental outcome, but long-term effects of longer and multiple exposures on emotional, physical, or social development remain unknown.
CONCLUSION
Various presynaptic and postsynaptic signaling proteins involved in synaptic transmission and plasticity have been identified as major targets for the neurophysiological effects of general anesthesia. The differential effects between various anesthetics on these multiple targets likely contribute to their agent-specific pharmacological actions but may also mediate distinct undesirable toxic effects. Based on animal studies, understanding the mechanisms involved is essential for the development of more specific and safer anesthetic drugs, as well as for mechanism-based application of currently available agents. To date, the overwhelming translational plausibility of anesthetic neurotoxicity, even in vulnerable populations, has not been convincingly supported clinically, but the potential significance of an effect mandates further inquiry.
ACKNOWLEDGEMENTS
Declared none.
CONSENT FOR PUBLICATION
Not applicable.
FUNDING
Dr. Hemmings receives research funding from the US National Institutes of Health and from Instrumentation Laboratory, and consulting fees from Elsevier.
Supported in part by US NIH grants GM130722 (JP) and GM058055 (HCH).
CONFLICT OF INTEREST
The authors declare no conflict of interest, financial or otherwise.
REFERENCES
- 1.Lukatch H.S., MacIver M.B. Voltage-clamp analysis of halothane effects on GABA(A fast) and GABA(A slow) inhibitory currents. Brain Res. 1997;765(1):108–112. doi: 10.1016/S0006-8993(97)00516-7. [DOI] [PubMed] [Google Scholar]
- 2.Sonner J.M., Zhang Y., Stabernack C., Abaigar W., Xing Y., Laster M.J. GABA(A) receptor blockade antagonizes the immobilizing action of propofol but not ketamine or isoflurane in a dose-related manner. Anesth. Analg. 2003;96(3):706–712. doi: 10.1213/01.ANE.0000048821.23225.3A. [DOI] [PubMed] [Google Scholar]
- 3.MacIver M.B. Anesthetic agent-specific effects on synaptic inhibition. Anesth. Analg. 2014;119(3):558–569. doi: 10.1213/ANE.0000000000000321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Perouansky M., Baranov D., Salman M., Yaari Y. Effects of halothane on glutamate receptor-mediated excitatory postsynaptic currents. A patch-clamp study in adult mouse hippocampal slices. Anesthesiology. 1995;83(1):109–119. doi: 10.1097/00000542-199507000-00014. [DOI] [PubMed] [Google Scholar]
- 5.Maclver M.B., Mikulec A.A., Amagasu S.M., Monroe F.A. Volatile anesthetics depress glutamate transmission via presynaptic actions. Anesthesiology. 1996;85(4):823–834. doi: 10.1097/00000542-199610000-00018. [DOI] [PubMed] [Google Scholar]
- 6.Jevtović-Todorović V., Todorović S.M., Mennerick S., Powell S., Dikranian K., Benshoff N., Zorumski C.F., Olney J.W. Nitrous oxide (laughing gas) is an NMDA antagonist, neuroprotectant and neurotoxin. Nat. Med. 1998;4(4):460–463. doi: 10.1038/nm0498-460. [DOI] [PubMed] [Google Scholar]
- 7.Wakasugi M., Hirota K., Roth S.H., Ito Y. The effects of general anesthetics on excitatory and inhibitory synaptic transmission in area CA1 of the rat hippocampus in vitro. Anesth. Analg. 1999;88(3):676–680. doi: 10.1213/00000539-199903000-00039. [DOI] [PubMed] [Google Scholar]
- 8.Pittson S., Himmel A.M., MacIver M.B. Multiple synaptic and membrane sites of anesthetic action in the CA1 region of rat hippocampal slices. BMC Neurosci. 2004;5:52. doi: 10.1186/1471-2202-5-52. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Berg-Johnsen J., Langmoen I.A. The effect of isoflurane on unmyelinated and myelinated fibres in the rat brain. Acta Physiol. Scand. 1986;127(1):87–93. doi: 10.1111/j.1748-1716.1986.tb07879.x. [DOI] [PubMed] [Google Scholar]
- 10.Mikulec A.A., Pittson S., Amagasu S.M., Monroe F.A., MacIver M.B. Halothane depresses action potential conduction in hippocampal axons. Brain Res. 1998;796(1-2):231–238. doi: 10.1016/S0006-8993(98)00348-5. [DOI] [PubMed] [Google Scholar]
- 11.Wu X.S., Sun J.Y., Evers A.S., Crowder M., Wu L.G. Isoflurane inhibits transmitter release and the presynaptic action potential. Anesthesiology. 2004;100(3):663–670. doi: 10.1097/00000542-200403000-00029. [DOI] [PubMed] [Google Scholar]
- 12.Ouyang W., Hemmings H.C., Jr Depression by isoflurane of the action potential and underlying voltage-gated ion currents in isolated rat neurohypophysial nerve terminals. J. Pharmacol. Exp. Ther. 2005;312(2):801–808. doi: 10.1124/jpet.104.074609. [DOI] [PubMed] [Google Scholar]
- 13.Baumgart J.P., Zhou Z.Y., Hara M., Cook D.C., Hoppa M.B., Ryan T.A., Hemmings H.C., Jr Isoflurane inhibits synaptic vesicle exocytosis through reduced Ca2+ influx, not Ca2+-exocytosis coupling. Proc. Natl. Acad. Sci. USA. 2015;112(38):11959–11964. doi: 10.1073/pnas.1500525112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Koyanagi Y., Torturo C.L., Cook D.C., Zhou Z., Hemmings H.C., Jr Role of specific presynaptic calcium channel subtypes in isoflurane inhibition of synaptic vesicle exocytosis in rat hippocampal neurones. Br. J. Anaesth. 2019;123(2):219–227. doi: 10.1016/j.bja.2019.03.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Westphalen R.I., Hemmings H.C., Jr Selective depression by general anesthetics of glutamate versus GABA release from isolated cortical nerve terminals. J. Pharmacol. Exp. Ther. 2003;304(3):1188–1196. doi: 10.1124/jpet.102.044685. [DOI] [PubMed] [Google Scholar]
- 16.Westphalen R.I., Hemmings H.C. Jr Volatile anesthetic effects on glutamate versus GABA release from isolated rat cortical nerve terminals: basal release. J. Pharmacol. Exp. Ther. 2006;316(1):208–215. doi: 10.1124/jpet.105.090647. [DOI] [PubMed] [Google Scholar]
- 17.Torturo C.L., Zhou Z.Y., Ryan T.A., Hemmings H.C. Isoflurane inhibits dopaminergic synaptic vesicle exocytosis coupled to CaV2.1 and CaV2.2 in rat midbrain neurons. eNeuro, 2019, 6(1), ENEURO.0278-18.2018. [DOI] [PMC free article] [PubMed]
- 18.Dickinson R., Peterson B.K., Banks P., Simillis C., Martin J.C., Valenzuela C.A., Maze M., Franks N.P. Competitive inhibition at the glycine site of the N-methyl-D-aspartate receptor by the anesthetics xenon and isoflurane: evidence from molecular modeling and electrophysiology. Anesthesiology. 2007;107(5):756–767. doi: 10.1097/01.anes.0000287061.77674.71. [DOI] [PubMed] [Google Scholar]
- 19.Rochefort N.L., Konnerth A. Dendritic spines: from structure to in vivo function. EMBO Rep. 2012;13(8):699–708. doi: 10.1038/embor.2012.102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Yang G., Chang P.C., Bekker A., Blanck T.J., Gan W.B. Transient effects of anesthetics on dendritic spines and filopodia in the living mouse cortex. Anesthesiology. 2011;115(4):718–726. doi: 10.1097/ALN.0b013e318229a660. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Platholi J., Herold K.F., Hemmings H.C., Jr, Halpain S. Isoflurane reversibly destabilizes hippocampal dendritic spines by an actin-dependent mechanism. PLoS One. 2014;9(7):e102978. doi: 10.1371/journal.pone.0102978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Kirson E.D., Yaari Y., Perouansky M. Presynaptic and postsynaptic actions of halothane at glutamatergic synapses in the mouse hippocampus. Br. J. Pharmacol. 1998;124(8):1607–1614. doi: 10.1038/sj.bjp.0701996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Seeman P. The membrane expansion theory of anesthesia: direct evidence using ethanol and a high-precision density meter. Experientia. 1974;30(7):759–760. doi: 10.1007/BF01924170. [DOI] [PubMed] [Google Scholar]
- 24.Herold K.F., Sanford R.L., Lee W., Andersen O.S., Hemmings H.C., Jr Clinical concentrations of chemically diverse general anesthetics minimally affect lipid bilayer properties. Proc. Natl. Acad. Sci. USA. 2017;114(12):3109–3114. doi: 10.1073/pnas.1611717114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Tibbs G.R., Barrie A.P., Van Mieghem F.J., McMahon H.T., Nicholls D.G. Repetitive action potentials in isolated nerve terminals in the presence of 4-aminopyridine: effects on cytosolic free Ca2+ and glutamate release. J. Neurochem. 1989;53(6):1693–1699. doi: 10.1111/j.1471-4159.1989.tb09232.x. [DOI] [PubMed] [Google Scholar]
- 26.Catterall W.A. Molecular mechanisms of gating and drug block of sodium channels. Novartis Found. Symp. 2002;241:206–218. [PubMed] [Google Scholar]
- 27.Lewis A.H., Raman I.M. Resurgent current of voltage-gated Na(+) channels. J. Physiol. 2014;592(22):4825–4838. doi: 10.1113/jphysiol.2014.277582. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Hodgkin A.L., Huxley A.F. A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol. 1952;117(4):500–544. doi: 10.1113/jphysiol.1952.sp004764. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Ragsdale D.S., McPhee J.C., Scheuer T., Catterall W.A. Molecular determinants of state-dependent block of Na+ channels by local anesthetics. Science. 1994;265(5179):1724–1728. doi: 10.1126/science.8085162. [DOI] [PubMed] [Google Scholar]
- 30.Rehberg B., Xiao Y.H., Duch D.S. Central nervous system sodium channels are significantly suppressed at clinical concentrations of volatile anesthetics. Anesthesiology. 1996;84(5):1223–1233. doi: 10.1097/00000542-199605000-00025. [DOI] [PubMed] [Google Scholar]
- 31.OuYang W., Hemmings H.C., Jr Isoform-selective effects of isoflurane on voltage-gated Na+ channels. Anesthesiology. 2007;107(1):91–98. doi: 10.1097/01.anes.0000268390.28362.4a. [DOI] [PubMed] [Google Scholar]
- 32.Herold K.F., Hemmings H.C., Jr Sodium channels as targets for volatile anesthetics. Front. Pharmacol. 2012;3:50. doi: 10.3389/fphar.2012.00050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Ratnakumari L., Hemmings H.C., Jr Inhibition of presynaptic sodium channels by halothane. Anesthesiology. 1998;88(4):1043–1054. doi: 10.1097/00000542-199804000-00025. [DOI] [PubMed] [Google Scholar]
- 34.Ouyang W., Wang G., Hemmings H.C., Jr Isoflurane and propofol inhibit voltage-gated sodium channels in isolated rat neurohypophysial nerve terminals. Mol. Pharmacol. 2003;64(2):373–381. doi: 10.1124/mol.64.2.373. [DOI] [PubMed] [Google Scholar]
- 35.Palti Y., Adelman W.J., Jr Measurement of axonal membrane conductances and capacity by means of a varying potential control voltage clamp. J. Membr. Biol. 1969;1(1):431–458. doi: 10.1007/BF01869791. [DOI] [PubMed] [Google Scholar]
- 36.Rudy B. Slow inactivation of the sodium conductance in squid giant axons. Pronase resistance. J. Physiol. 1978;283:1–21. doi: 10.1113/jphysiol.1978.sp012485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Stafstrom C.E. Persistent sodium current and its role in epilepsy. Epilepsy Curr. 2007;7(1):15–22. doi: 10.1111/j.1535-7511.2007.00156.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Zhao W., Zhang M., Liu J., Liang P., Wang R., Hemmings H.C., Zhou C. Isoflurane modulates hippocampal cornu ammonis pyramidal neuron excitability by inhibition of both transient and persistent sodium currents in mice. Anesthesiology. 2019;131(1):94–104. doi: 10.1097/ALN.0000000000002753. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Ratnakumari L., Hemmings H.C., Jr Effects of propofol on sodium channel-dependent sodium influx and glutamate release in rat cerebrocortical synaptosomes. Anesthesiology. 1997;86(2):428–439. doi: 10.1097/00000542-199702000-00018. [DOI] [PubMed] [Google Scholar]
- 40.Liu Q.Z., Hao M., Zhou Z.Y., Ge J.L., Wu Y.C., Zhao L.L., Wu X., Feng Y., Gao H., Li S., Xue L. Propofol reduces synaptic strength by inhibiting sodium and calcium channels at nerve terminals. Protein Cell. 2019;10(9):688–693. doi: 10.1007/s13238-019-0624-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Ratnakumari L., Hemmings H.C. Jr Inhibition by propofol of [3H]-batrachotoxinin-A 20-alpha-benzoate binding to voltage-dependent sodium channels in rat cortical synaptosomes. Br. J. Pharmacol. 1996;119(7):1498–1504. doi: 10.1111/j.1476-5381.1996.tb16064.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Reckziegel G., Friederich P., Urban B.W. Ketamine effects on human neuronal Na+ channels. Eur. J. Anaesthesiol. 2002;19(9):634–640. doi: 10.1017/S0265021502001047. [DOI] [PubMed] [Google Scholar]
- 43.Frenkel C., Urban B.W. Molecular actions of racemic ketamine on human CNS sodium channels. Br. J. Anaesth. 1992;69(3):292–297. doi: 10.1093/bja/69.3.292. [DOI] [PubMed] [Google Scholar]
- 44.Catterall W.A. From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels. Neuron. 2000;26(1):13–25. doi: 10.1016/S0896-6273(00)81133-2. [DOI] [PubMed] [Google Scholar]
- 45.Goldin A.L. Resurgence of sodium channel research. Annu. Rev. Physiol. 2001;63:871–894. doi: 10.1146/annurev.physiol.63.1.871. [DOI] [PubMed] [Google Scholar]
- 46.Black J.A., Waxman S.G. Sodium channel expression: a dynamic process in neurons and non-neuronal cells. Dev. Neurosci. 1996;18(3):139–152. doi: 10.1159/000111403. [DOI] [PubMed] [Google Scholar]
- 47.Wood J.N., Baker M. Voltage-gated sodium channels. Curr. Opin. Pharmacol. 2001;1(1):17–21. doi: 10.1016/S1471-4892(01)00007-8. [DOI] [PubMed] [Google Scholar]
- 48.Purtell K., Gingrich K.J., Ouyang W., Herold K.F., Hemmings H.C., Jr Activity-dependent depression of neuronal sodium channels by the general anaesthetic isoflurane. Br. J. Anaesth. 2015;115(1):112–121. doi: 10.1093/bja/aev203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Zhou C., Johnson K.W., Herold K.F., Hemmings H.C., Jr Differential inhibition of neuronal sodium channel subtypes by the general anesthetic isoflurane. J. Pharmacol. Exp. Ther. 2019;369(2):200–211. doi: 10.1124/jpet.118.254938. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Shiraishi M., Harris R.A. Effects of alcohols and anesthetics on recombinant voltage-gated Na+ channels. J. Pharmacol. Exp. Ther. 2004;309(3):987–994. doi: 10.1124/jpet.103.064063. [DOI] [PubMed] [Google Scholar]
- 51.Rehberg B., Duch D.S. Suppression of central nervous system sodium channels by propofol. Anesthesiology. 1999;91(2):512–520. doi: 10.1097/00000542-199908000-00026. [DOI] [PubMed] [Google Scholar]
- 52.Lai H.C., Jan L.Y. The distribution and targeting of neuronal voltage-gated ion channels. Nat. Rev. Neurosci. 2006;7(7):548–562. doi: 10.1038/nrn1938. [DOI] [PubMed] [Google Scholar]
- 53.Ogiwara I., Miyamoto H., Morita N., Atapour N., Mazaki E., Inoue I., Takeuchi T., Itohara S., Yanagawa Y., Obata K., Furuichi T., Hensch T.K., Yamakawa K. Nav1.1 localizes to axons of parvalbumin-positive inhibitory interneurons: a circuit basis for epileptic seizures in mice carrying an Scn1a gene mutation. J. Neurosci. 2007;27(22):5903–5914. doi: 10.1523/JNEUROSCI.5270-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Lorincz A., Nusser Z. Cell-type-dependent molecular composition of the axon initial segment. J. Neurosci. 2008;28(53):14329–14340. doi: 10.1523/JNEUROSCI.4833-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Johnson K.W., Herold K.F., Milner T.A., Hemmings H.C., Jr, Platholi J. Sodium channel subtypes are differentially localized to pre- and post-synaptic sites in rat hippocampus. J. Comp. Neurol. 2017;525(16):3563–3578. doi: 10.1002/cne.24291. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Hodgson P.S., Liu S.S., Gras T.W. Does epidural anesthesia have general anesthetic effects? A prospective, randomized, double-blind, placebo-controlled trial. Anesthesiology. 1999;91(6):1687–1692. doi: 10.1097/00000542-199912000-00021. [DOI] [PubMed] [Google Scholar]
- 57.De Santi, L.; Polimeni, G.; Cuzzocrea, S.; Esposito, E.; Sessa, E.; Annunziata, P.; Bramanti, P. Neuroinflammation and neuroprotection: an update on (future) neurotrophin-related strategies in multiple sclerosis treatment. Curr. Med. Chem., 2011;18(12):1775–1784. doi: 10.2174/092986711795496881. [DOI] [PubMed] [Google Scholar]
- 58.Zhang Y., Laster M.J., Eger E.I., II, Sharma M., Sonner J.M. Lidocaine, MK-801, and MAC. Anesth. Analg. 2007;104(5):1098–1102. doi: 10.1213/01.ane.0000260318.60504.a9. [DOI] [PubMed] [Google Scholar]
- 59.Zhang Y., Guzinski M., Eger E.I., II, Laster M.J., Sharma M., Harris R.A., Hemmings H.C., Jr Bidirectional modulation of isoflurane potency by intrathecal tetrodotoxin and veratridine in rats. Br. J. Pharmacol. 2010;159(4):872–878. doi: 10.1111/j.1476-5381.2009.00583.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Laster M.J., Zhang Y., Eger E.I., II, Shnayderman D., Sonner J.M. Alterations in spinal, but not cerebral, cerebrospinal fluid Na+ concentrations affect the isoflurane minimum alveolar concentration in rats. Anesth. Analg. 2007;105(3):661–665. doi: 10.1213/01.ane.0000278090.88402.26. [DOI] [PubMed] [Google Scholar]
- 61.Zhang Y., Sharma M., Eger E.I., II, Laster M.J., Hemmings H.C., Jr, Harris R.A. Intrathecal veratridine administration increases minimum alveolar concentration in rats. Anesth. Analg. 2008;107(3):875–878. doi: 10.1213/ane.0b013e3181815fbc. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Pal D., Jones J.M., Wisidagamage S., Meisler M.H., Mashour G.A. Reduced Nav1.6 sodium channel activity in mice increases In Vivo sensitivity to volatile anesthetics. PLoS One. 2015;10(8):e0134960. doi: 10.1371/journal.pone.0134960. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Arhem P., Klement G., Nilsson J. Mechanisms of anesthesia: towards integrating network, cellular, and molecular level modeling. Neuropsychopharmacology. 2003;28(Suppl. 1):S40–S47. doi: 10.1038/sj.npp.1300142. [DOI] [PubMed] [Google Scholar]
- 64.Hentschke H., Raz A., Krause B.M., Murphy C.A., Banks M.I. Disruption of cortical network activity by the general anaesthetic isoflurane. Br. J. Anaesth. 2017;119(4):685–696. doi: 10.1093/bja/aex199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Cantrell A.R., Catterall W.A. Neuromodulation of Na+ channels: an unexpected form of cellular plasticity. Nat. Rev. Neurosci. 2001;2(6):397–407. doi: 10.1038/35077553. [DOI] [PubMed] [Google Scholar]
- 66.Carr D.B., Day M., Cantrell A.R., Held J., Scheuer T., Catterall W.A., Surmeier D.J. Transmitter modulation of slow, activity-dependent alterations in sodium channel availability endows neurons with a novel form of cellular plasticity. Neuron. 2003;39(5):793–806. doi: 10.1016/S0896-6273(03)00531-2. [DOI] [PubMed] [Google Scholar]
- 67.Yin L., Rasch M.J., He Q., Wu S., Dou F., Shu Y. selective modulation of axonal sodium channel subtypes by 5-HT1A receptor in cortical pyramidal neuron. Cereb. Cortex. 2017;27(1):509–521. doi: 10.1093/cercor/bhv245. [DOI] [PubMed] [Google Scholar]
- 68.Maurice N., Tkatch T., Meisler M., Sprunger L.K., Surmeier D.J. D1/D5 dopamine receptor activation differentially modulates rapidly inactivating and persistent sodium currents in prefrontal cortex pyramidal neurons. J. Neurosci. 2001;21(7):2268–2277. doi: 10.1523/JNEUROSCI.21-07-02268.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Chen Y., Yu F.H., Sharp E.M., Beacham D., Scheuer T., Catterall W.A. Functional properties and differential neuromodulation of Na(v)1.6 channels. Mol. Cell. Neurosci. 2008;38(4):607–615. doi: 10.1016/j.mcn.2008.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Hemmings H.C., Jr, Adamo A.I. Effects of halothane and propofol on purified brain protein kinase C activation. Anesthesiology. 1994;81(1):147–155. doi: 10.1097/00000542-199407000-00021. [DOI] [PubMed] [Google Scholar]
- 71.Hemmings H.C., Jr, Adamo A.I. Activation of endogenous protein kinase C by halothane in synaptosomes. Anesthesiology. 1996;84(3):652–662. doi: 10.1097/00000542-199603000-00021. [DOI] [PubMed] [Google Scholar]
- 72.Jahn R., Fasshauer D. Molecular machines governing exocytosis of synaptic vesicles. Nature. 2012;490(7419):201–207. doi: 10.1038/nature11320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Südhof T.C. The molecular machinery of neurotransmitter release (Nobel lecture). Angew. Chem. Int. Ed. Engl. 2014;53(47):12696–12717. doi: 10.1002/anie.201406359. [DOI] [PubMed] [Google Scholar]
- 74.Catterall W.A. Voltage-gated calcium channels. Cold Spring Harb. Perspect. Biol. 2011;3(8):a003947. doi: 10.1101/cshperspect.a003947. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Simms B.A., Zamponi G.W. Neuronal voltage-gated calcium channels: structure, function, and dysfunction. Neuron. 2014;82(1):24–45. doi: 10.1016/j.neuron.2014.03.016. [DOI] [PubMed] [Google Scholar]
- 76.Miao N., Frazer M.J., Lynch C., III Volatile anesthetics depress Ca2+ transients and glutamate release in isolated cerebral synaptosomes. Anesthesiology. 1995;83(3):593–603. doi: 10.1097/00000542-199509000-00019. [DOI] [PubMed] [Google Scholar]
- 77.Hemmings H.C., Jr, Yan W., Westphalen R.I., Ryan T.A. The general anesthetic isoflurane depresses synaptic vesicle exocytosis. Mol. Pharmacol. 2005;67(5):1591–1599. doi: 10.1124/mol.104.003210. [DOI] [PubMed] [Google Scholar]
- 78.Daniell L.C., Harris R.A. Neuronal intracellular calcium concentrations are altered by anesthetics: relationship to membrane fluidization. J. Pharmacol. Exp. Ther. 1988;245(1):1–7. [PubMed] [Google Scholar]
- 79.Mody I., Tanelian D.L., MacIver M.B. Halothane enhances tonic neuronal inhibition by elevating intracellular calcium. Brain Res. 1991;538(2):319–323. doi: 10.1016/0006-8993(91)90447-4. [DOI] [PubMed] [Google Scholar]
- 80.Nicoll R.A., Madison D.V. General anesthetics hyperpolarize neurons in the vertebrate central nervous system. Science. 1982;217(4564):1055–1057. doi: 10.1126/science.7112112. [DOI] [PubMed] [Google Scholar]
- 81.Kitamura A., Marszalec W., Yeh J.Z., Narahashi T. Effects of halothane and propofol on excitatory and inhibitory synaptic transmission in rat cortical neurons. J. Pharmacol. Exp. Ther. 2003;304(1):162–171. doi: 10.1124/jpet.102.043273. [DOI] [PubMed] [Google Scholar]
- 82.Krnjević K., Puil E. Halothane suppresses slow inward currents in hippocampal slices. Can. J. Physiol. Pharmacol. 1988;66(12):1570–1575. doi: 10.1139/y88-257. [DOI] [PubMed] [Google Scholar]
- 83.el-Beheiry H., Puil E. Anaesthetic depression of excitatory synaptic transmission in neocortex. Exp. Brain Res. 1989;77(1):87–93. doi: 10.1007/BF00250570. [DOI] [PubMed] [Google Scholar]
- 84.Bleakman D., Jones M.V., Harrison N.L. The effects of four general anesthetics on intracellular [Ca2+] in cultured rat hippocampal neurons. Neuropharmacology. 1995;34(5):541–551. doi: 10.1016/0028-3908(95)00022-X. [DOI] [PubMed] [Google Scholar]
- 85.Wamil A.W., Franks J.J., Janicki P.K., Horn J.L., Franks W.T. Halothane alters electrical activity and calcium dynamics in cultured mouse cortical, spinal cord, and dorsal root ganglion neurons. Neurosci. Lett. 1996;216(2):93–96. doi: 10.1016/0304-3940(96)13003-2. [DOI] [PubMed] [Google Scholar]
- 86.Olivera B.M., Miljanich G.P., Ramachandran J., Adams M.E. Calcium channel diversity and neurotransmitter release: the omega-conotoxins and omega-agatoxins. Annu. Rev. Biochem. 1994;63:823–867. doi: 10.1146/annurev.bi.63.070194.004135. [DOI] [PubMed] [Google Scholar]
- 87.Catterall W.A. Structure and function of neuronal Ca2+ channels and their role in neurotransmitter release. Cell Calcium. 1998;24(5-6):307–323. doi: 10.1016/S0143-4160(98)90055-0. [DOI] [PubMed] [Google Scholar]
- 88.Wheeler D.B., Sather W.A., Randall A., Tsien R.W. Distinctive properties of a neuronal calcium channel and its contribution to excitatory synaptic transmission in the central nervous system. Adv. Second Messenger Phosphoprotein Res. 1994;29:155–171. doi: 10.1016/S1040-7952(06)80014-5. [DOI] [PubMed] [Google Scholar]
- 89.Study R.E. Isoflurane inhibits multiple voltage-gated calcium currents in hippocampal pyramidal neurons. Anesthesiology. 1994;81(1):104–116. doi: 10.1097/00000542-199407000-00016. [DOI] [PubMed] [Google Scholar]
- 90.Xu F., Sarti P., Zhang J., Blanck T.J. Halothane and isoflurane alter calcium dynamics in rat cerebrocortical synaptosomes. Anesth. Analg. 1998;87(3):701–710. doi: 10.1097/00000539-199809000-00040. [DOI] [PubMed] [Google Scholar]
- 91.Hall A.C., Lieb W.R., Franks N.P. Insensitivity of P-type calcium channels to inhalational and intravenous general anesthetics. Anesthesiology. 1994;81(1):117–123. doi: 10.1097/00000542-199407000-00017. [DOI] [PubMed] [Google Scholar]
- 92.Mintz I.M., Venema V.J., Swiderek K.M., Lee T.D., Bean B.P., Adams M.E. P-type calcium channels blocked by the spider toxin omega-Aga-IVA. Nature. 1992;355(6363):827–829. doi: 10.1038/355827a0. [DOI] [PubMed] [Google Scholar]
- 93.Uchitel O.D., Protti D.A., Sanchez V., Cherksey B.D., Sugimori M., Llinás R. P-type voltage-dependent calcium channel mediates presynaptic calcium influx and transmitter release in mammalian synapses. Proc. Natl. Acad. Sci. USA. 1992;89(8):3330–3333. doi: 10.1073/pnas.89.8.3330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Kamatchi G.L., Chan C.K., Snutch T., Durieux M.E., Lynch C. III Volatile anesthetic inhibition of neuronal Ca channel currents expressed in Xenopus oocytes. Brain Res. 1999;831(1-2):85–96. doi: 10.1016/S0006-8993(99)01401-8. [DOI] [PubMed] [Google Scholar]
- 95.Rajagopal S., Fang H., Lynch C., III, Sando J.J., Kamatchi G.L. Effects of isoflurane on the expressed Cav2.2 currents in Xenopus oocytes depend on the activation of protein kinase Cδ and its phosphorylation sites in the Cav2.2α1 subunits. Neuroscience. 2011;182:232–240. doi: 10.1016/j.neuroscience.2011.02.041. [DOI] [PubMed] [Google Scholar]
- 96.Takei T., Saegusa H., Zong S., Murakoshi T., Makita K., Tanabe T. Anesthetic sensitivities to propofol and halothane in mice lacking the R-type (Cav2.3) Ca2+ channel. Anesth. Analg. 2003;97(1):96–103. doi: 10.1213/01.ANE.0000065548.83253.5C. [DOI] [PubMed] [Google Scholar]
- 97.Wu L.G., Borst J.G., Sakmann B. R-type Ca2+ currents evoke transmitter release at a rat central synapse. Proc. Natl. Acad. Sci. USA. 1998;95(8):4720–4725. doi: 10.1073/pnas.95.8.4720. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Ashby M.C., Tepikin A.V. ER calcium and the functions of intracellular organelles. Semin. Cell Dev. Biol. 2001;12(1):11–17. doi: 10.1006/scdb.2000.0212. [DOI] [PubMed] [Google Scholar]
- 99.Wei H., Liang G., Yang H., Wang Q., Hawkins B., Madesh M., Wang S., Eckenhoff R.G. The common inhalational anesthetic isoflurane induces apoptosis via activation of inositol 1,4,5-trisphosphate receptors. Anesthesiology. 2008;108(2):251–260. doi: 10.1097/01.anes.0000299435.59242.0e. [DOI] [PubMed] [Google Scholar]
- 100.Murayama T., Ogawa Y. Properties of Ryr3 ryanodine receptor isoform in mammalian brain. J. Biol. Chem. 1996;271(9):5079–5084. doi: 10.1074/jbc.271.9.5079. [DOI] [PubMed] [Google Scholar]
- 101.Verkhratsky A. Physiology and pathophysiology of the calcium store in the endoplasmic reticulum of neurons. Physiol. Rev. 2005;85(1):201–279. doi: 10.1152/physrev.00004.2004. [DOI] [PubMed] [Google Scholar]
- 102.Liu X., Betzenhauser M.J., Reiken S., Meli A.C., Xie W., Chen B.X., Arancio O., Marks A.R. Role of leaky neuronal ryanodine receptors in stress-induced cognitive dysfunction. Cell. 2012;150(5):1055–1067. doi: 10.1016/j.cell.2012.06.052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.de Juan-Sanz J., Holt G.T., Schreiter E.R., de Juan F., Kim D.S., Ryan T.A. Axonal endoplasmic reticulum Ca2+ content controls release probability in CNS nerve terminals. Neuron. 2017;93(4):867–881.e6. doi: 10.1016/j.neuron.2017.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Gomez R.S., Guatimosim C., Barbosa J., Jr, Massensini A.R., Gomez M.V., Prado M.A. Halothane-induced intracellular calcium release in cholinergic cells. Brain Res. 2001;921(1-2):106–114. doi: 10.1016/S0006-8993(01)03098-0. [DOI] [PubMed] [Google Scholar]
- 105.Pinheiro A.C., Gomez R.S., Guatimosim C., Silva J.H., Prado M.A., Gomez M.V. The effect of sevoflurane on intracellular calcium concentration from cholinergic cells. Brain Res. Bull. 2006;69(2):147–152. doi: 10.1016/j.brainresbull.2005.11.016. [DOI] [PubMed] [Google Scholar]
- 106.Jiang D., Chen W., Xiao J., Wang R., Kong H., Jones P.P., Zhang L., Fruen B., Chen S.R. Reduced threshold for luminal Ca2+ activation of RyR1 underlies a causal mechanism of porcine malignant hyperthermia. J. Biol. Chem. 2008;283(30):20813–20820. doi: 10.1074/jbc.M801944200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Rosenberg H., Pollock N., Schiemann A., Bulger T., Stowell K. Malignant hyperthermia: a review. Orphanet J. Rare Dis. 2015;10:93. doi: 10.1186/s13023-015-0310-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Kim S.H., Ryan T.A. Balance of calcineurin Aα and CDK5 activities sets release probability at nerve terminals. J. Neurosci. 2013;33(21):8937–8950. doi: 10.1523/JNEUROSCI.4288-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Pocock G., Richards C.D. Hydrogen ion regulation in rat cerebellar granule cells studied by single-cell fluorescence microscopy. Eur. J. Neurosci. 1992;4(2):136–143. doi: 10.1111/j.1460-9568.1992.tb00860.x. [DOI] [PubMed] [Google Scholar]
- 110.Schlame M., Hemmings H.C., Jr Inhibition by volatile anesthetics of endogenous glutamate release from synaptosomes by a presynaptic mechanism. Anesthesiology. 1995;82(6):1406–1416. doi: 10.1097/00000542-199506000-00012. [DOI] [PubMed] [Google Scholar]
- 111.Wang H.Y., Eguchi K., Yamashita T., Takahashi T. Frequency-dependent block of excitatory neurotransmission by isoflurane via dual presynaptic mechanisms. J. Neurosci. 2020;40(21):4103–4115. doi: 10.1523/JNEUROSCI.2946-19.2020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Lingamaneni R., Birch M.L., Hemmings H.C., Jr Widespread inhibition of sodium channel-dependent glutamate release from isolated nerve terminals by isoflurane and propofol. Anesthesiology. 2001;95(6):1460–1466. doi: 10.1097/00000542-200112000-00027. [DOI] [PubMed] [Google Scholar]
- 113.Westphalen R.I., Kwak N.B., Daniels K., Hemmings H.C., Jr Regional differences in the effects of isoflurane on neurotransmitter release. Neuropharmacology. 2011;61(4):699–706. doi: 10.1016/j.neuropharm.2011.05.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Vanini G., Watson C.J., Lydic R., Baghdoyan H.A. Gamma-aminobutyric acid-mediated neurotransmission in the pontine reticular formation modulates hypnosis, immobility, and breathing during isoflurane anesthesia. Anesthesiology. 2008;109(6):978–988. doi: 10.1097/ALN.0b013e31818e3b1b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Murugaiah K.D., Hemmings H.C., Jr Effects of intravenous general anesthetics on [3H]GABA release from rat cortical synaptosomes. Anesthesiology. 1998;89(4):919–928. doi: 10.1097/00000542-199810000-00017. [DOI] [PubMed] [Google Scholar]
- 116.Bieda M.C., MacIver M.B. Major role for tonic GABAA conductances in anesthetic suppression of intrinsic neuronal excitability. J. Neurophysiol. 2004;92(3):1658–1667. doi: 10.1152/jn.00223.2004. [DOI] [PubMed] [Google Scholar]
- 117.Augustine G.J. How does calcium trigger neurotransmitter release? Curr. Opin. Neurobiol. 2001;11(3):320–326. doi: 10.1016/S0959-4388(00)00214-2. [DOI] [PubMed] [Google Scholar]
- 118.Chapman E.R. How does synaptotagmin trigger neurotransmitter release? Annu. Rev. Biochem. 2008;77:615–641. doi: 10.1146/annurev.biochem.77.062005.101135. [DOI] [PubMed] [Google Scholar]
- 119.Paul A., Crow M., Raudales R., He M., Gillis J., Huang Z.J. Transcriptional Architecture of synaptic communication delineates GABAergic neuron identity. Cell. 2017;171(3):522–539.e20. doi: 10.1016/j.cell.2017.08.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Hemmings H.C. In: Suppressing the Mind: Anesthetic Modulation of Memory and Consciousness. Hudetz A., Pearce R., editors. Totowa, NJ: Humana Press; 2010. Molecular Targets of General Anesthetics in the Nervous System. pp. 11–31. [Google Scholar]
- 121.Hu H., Jonas P. A supercritical density of Na(+) channels ensures fast signaling in GABAergic interneuron axons. Nat. Neurosci. 2014;17(5):686–693. doi: 10.1038/nn.3678. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Li T., Tian C., Scalmani P., Frassoni C., Mantegazza M., Wang Y., Yang M., Wu S., Shu Y. Action potential initiation in neocortical inhibitory interneurons. PLoS Biol. 2014;12(9):e1001944. doi: 10.1371/journal.pbio.1001944. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Westphalen R.I., Yu J., Krivitski M., Jih T.Y., Hemmings H.C., Jr Regional differences in nerve terminal Na+ channel subtype expression and Na+ channel-dependent glutamate and GABA release in rat CNS. J. Neurochem. 2010;113(6):1611–1620. doi: 10.1111/j.1471-4159.2010.06722.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Buggy D.J., Nicol B., Rowbotham D.J., Lambert D.G. Effects of intravenous anesthetic agents on glutamate release: a role for GABAA receptor-mediated inhibition. Anesthesiology. 2000;92(4):1067–1073. doi: 10.1097/00000542-200004000-00025. [DOI] [PubMed] [Google Scholar]
- 125.Thureson-Klein A. Exocytosis from large and small dense cored vesicles in noradrenergic nerve terminals. Neuroscience. 1983;10(2):245–259. doi: 10.1016/0306-4522(83)90132-X. [DOI] [PubMed] [Google Scholar]
- 126.Rizo J., Südhof T.C. The membrane fusion enigma: SNAREs, Sec1/Munc18 proteins, and their accomplices--guilty as charged? Annu. Rev. Cell Dev. Biol. 2012;28:279–308. doi: 10.1146/annurev-cellbio-101011-155818. [DOI] [PubMed] [Google Scholar]
- 127.Liu C., Kershberg L., Wang J., Schneeberger S., Kaeser P.S. Dopamine secretion is mediated by sparse active zone-like release sites. Cell. 2018;172(4):706–718.e15. doi: 10.1016/j.cell.2018.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Monti J.M., Monti D. The involvement of dopamine in the modulation of sleep and waking. Sleep Med. Rev. 2007;11(2):113–133. doi: 10.1016/j.smrv.2006.08.003. [DOI] [PubMed] [Google Scholar]
- 129.Barrot M. The ventral tegmentum and dopamine: A new wave of diversity. Neuroscience. 2014;282:243–247. doi: 10.1016/j.neuroscience.2014.10.017. [DOI] [PubMed] [Google Scholar]
- 130.Solt K., Van Dort C.J., Chemali J.J., Taylor N.E., Kenny J.D., Brown E.N. Electrical stimulation of the ventral tegmental area induces reanimation from general anesthesia. Anesthesiology. 2014;121(2):311–319. doi: 10.1097/ALN.0000000000000117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Solt K., Cotten J.F., Cimenser A., Wong K.F., Chemali J.J., Brown E.N. Methylphenidate actively induces emergence from general anesthesia. Anesthesiology. 2011;115(4):791–803. doi: 10.1097/ALN.0b013e31822e92e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Chemali J.J., Van Dort C.J., Brown E.N., Solt K. Active emergence from propofol general anesthesia is induced by methylphenidate. Anesthesiology. 2012;116(5):998–1005. doi: 10.1097/ALN.0b013e3182518bfc. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Kenny J.D., Taylor N.E., Brown E.N., Solt K. Dextroamphetamine (but Not Atomoxetine) induces reanimation from general anesthesia: implications for the roles of dopamine and norepinephrine in active emergence. PLoS One. 2015;10(7):e0131914. doi: 10.1371/journal.pone.0131914. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Mantz J., Varlet C., Lecharny J.B., Henzel D., Lenot P., Desmonts J.M. Effects of volatile anesthetics, thiopental, and ketamine on spontaneous and depolarization-evoked dopamine release from striatal synaptosomes in the rat. Anesthesiology. 1994;80(2):352–363. doi: 10.1097/00000542-199402000-00015. [DOI] [PubMed] [Google Scholar]
- 135.Keita H., Henzel-Rouellé D., Dupont H., Desmonts J.M., Mantz J. Halothane and isoflurane increase spontaneous but reduce the N-methyl-D-aspartate-evoked dopamine release in rat striatal slices: evidence for direct presynaptic effects. Anesthesiology. 1999;91(6):1788–1797. doi: 10.1097/00000542-199912000-00033. [DOI] [PubMed] [Google Scholar]
- 136.Adachi Y.U., Watanabe K., Higuchi H., Satoh T., Zsilla G. Halothane decreases impulse-dependent but not cytoplasmic release of dopamine from rat striatal slices. Brain Res. Bull. 2001;56(6):521–524. doi: 10.1016/S0361-9230(01)00619-0. [DOI] [PubMed] [Google Scholar]
- 137.Westphalen R.I., Desai K.M., Hemmings H.C., Jr Presynaptic inhibition of the release of multiple major central nervous system neurotransmitter types by the inhaled anaesthetic isoflurane. Br. J. Anaesth. 2013;110(4):592–599. doi: 10.1093/bja/aes448. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Salord F., Keita H., Lecharny J.B., Henzel D., Desmonts J.M., Mantz J. Halothane and isoflurane differentially affect the regulation of dopamine and gamma-aminobutyric acid release mediated by presynaptic acetylcholine receptors in the rat striatum. Anesthesiology. 1997;86(3):632–641. doi: 10.1097/00000542-199703000-00016. [DOI] [PubMed] [Google Scholar]
- 139.Gärtner A., Polnau D.G., Staiger V., Sciarretta C., Minichiello L., Thoenen H., Bonhoeffer T., Korte M. Hippocampal long-term potentiation is supported by presynaptic and postsynaptic tyrosine receptor kinase B-mediated phospholipase C gamma signaling. J. Neurosci. 2006;26(13):3496–3504. doi: 10.1523/JNEUROSCI.3792-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Jovanovic J.N., Czernik A.J., Fienberg A.A., Greengard P., Sihra T.S. Synapsins as mediators of BDNF-enhanced neurotransmitter release. Nat. Neurosci. 2000;3(4):323–329. doi: 10.1038/73888. [DOI] [PubMed] [Google Scholar]
- 141.Thakker-Varia S., Alder J., Crozier R.A., Plummer M.R., Black I.B. Rab3A is required for brain-derived neurotrophic factor-induced synaptic plasticity: transcriptional analysis at the population and single-cell levels. J. Neurosci. 2001;21(17):6782–6790. doi: 10.1523/JNEUROSCI.21-17-06782.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Tyler W.J., Pozzo-Miller L.D. BDNF enhances quantal neurotransmitter release and increases the number of docked vesicles at the active zones of hippocampal excitatory synapses. J. Neurosci. 2001;21(12):4249–4258. doi: 10.1523/JNEUROSCI.21-12-04249.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Tyler W.J., Zhang X.L., Hartman K., Winterer J., Muller W., Stanton P.K., Pozzo-Miller L. BDNF increases release probability and the size of a rapidly recycling vesicle pool within rat hippocampal excitatory synapses. J. Physiol. 2006;574(Pt 3):787–803. doi: 10.1113/jphysiol.2006.111310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Pozzo-Miller L.D., Gottschalk W., Zhang L., McDermott K., Du J., Gopalakrishnan R., Oho C., Sheng Z.H., Lu B. Impairments in high-frequency transmission, synaptic vesicle docking, and synaptic protein distribution in the hippocampus of BDNF knockout mice. J. Neurosci. 1999;19(12):4972–4983. doi: 10.1523/JNEUROSCI.19-12-04972.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Lin P.Y., Kavalali E.T., Monteggia L.M. Genetic dissection of presynaptic and postsynaptic BDNF-TrkB signaling in synaptic efficacy of CA3-CA1 Synapses. Cell Rep. 2018;24(6):1550–1561. doi: 10.1016/j.celrep.2018.07.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Kojima M., Takei N., Numakawa T., Ishikawa Y., Suzuki S., Matsumoto T., Katoh-Semba R., Nawa H., Hatanaka H. Biological characterization and optical imaging of brain-derived neurotrophic factor-green fluorescent protein suggest an activity-dependent local release of brain-derived neurotrophic factor in neurites of cultured hippocampal neurons. J. Neurosci. Res. 2001;64(1):1–10. doi: 10.1002/jnr.1080. [DOI] [PubMed] [Google Scholar]
- 147.Matsuda N., Lu H., Fukata Y., Noritake J., Gao H., Mukherjee S., Nemoto T., Fukata M., Poo M.M. Differential activity-dependent secretion of brain-derived neurotrophic factor from axon and dendrite. J. Neurosci. 2009;29(45):14185–14198. doi: 10.1523/JNEUROSCI.1863-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Patel A.J., Honoré E., Lesage F., Fink M., Romey G., Lazdunski M. Inhalational anesthetics activate two-pore-domain background K+ channels. Nat. Neurosci. 1999;2(5):422–426. doi: 10.1038/8084. [DOI] [PubMed] [Google Scholar]
- 149.Covarrubias M., Barber A.F., Carnevale V., Treptow W., Eckenhoff R.G. Mechanistic insights into the modulation of voltage-gated ion channels by inhalational anesthetics. Biophys. J. 2015;109(10):2003–2011. doi: 10.1016/j.bpj.2015.09.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Steinberg E.A., Wafford K.A., Brickley S.G., Franks N.P., Wisden W. The role of K2p channels in anaesthesia and sleep. Pflugers Arch. 2015;467(5):907–916. doi: 10.1007/s00424-014-1654-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Yost C.S. Potassium channels: basic aspects, functional roles, and medical significance. Anesthesiology. 1999;90(4):1186–1203. doi: 10.1097/00000542-199904000-00035. [DOI] [PubMed] [Google Scholar]
- 152.Andres-Enguix I., Caley A., Yustos R., Schumacher M.A., Spanu P.D., Dickinson R., Maze M., Franks N.P. Determinants of the anesthetic sensitivity of two-pore domain acid-sensitive potassium channels: molecular cloning of an anesthetic-activated potassium channel from Lymnaea stagnalis. J. Biol. Chem. 2007;282(29):20977–20990. doi: 10.1074/jbc.M610692200. [DOI] [PubMed] [Google Scholar]
- 153.Conway K.E., Cotten J.F. Covalent modification of a volatile anesthetic regulatory site activates TASK-3 (KCNK9) tandem-pore potassium channels. Mol. Pharmacol. 2012;81(3):393–400. doi: 10.1124/mol.111.076281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Pang D.S., Robledo C.J., Carr D.R., Gent T.C., Vyssotski A.L., Caley A., Zecharia A.Y., Wisden W., Brickley S.G., Franks N.P. An unexpected role for TASK-3 potassium channels in network oscillations with implications for sleep mechanisms and anesthetic action. Proc. Natl. Acad. Sci. USA. 2009;106(41):17546–17551. doi: 10.1073/pnas.0907228106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Linden A.M., Aller M.I., Leppä E., Vekovischeva O., Aitta-Aho T., Veale E.L., Mathie A., Rosenberg P., Wisden W., Korpi E.R. The in vivo contributions of TASK-1-containing channels to the actions of inhalation anesthetics, the alpha(2) adrenergic sedative dexmedetomidine, and cannabinoid agonists. J. Pharmacol. Exp. Ther. 2006;317(2):615–626. doi: 10.1124/jpet.105.098525. [DOI] [PubMed] [Google Scholar]
- 156.Linden A.M., Sandu C., Aller M.I., Vekovischeva O.Y., Rosenberg P.H., Wisden W., Korpi E.R. TASK-3 knockout mice exhibit exaggerated nocturnal activity, impairments in cognitive functions, and reduced sensitivity to inhalation anesthetics. J. Pharmacol. Exp. Ther. 2007;323(3):924–934. doi: 10.1124/jpet.107.129544. [DOI] [PubMed] [Google Scholar]
- 157.Chae Y.J., Zhang J., Au P., Sabbadini M., Xie G.X., Yost C.S. Discrete change in volatile anesthetic sensitivity in mice with inactivated tandem pore potassium ion channel TRESK. Anesthesiology. 2010;113(6):1326–1337. doi: 10.1097/ALN.0b013e3181f90ca5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Lazarenko R.M., Willcox S.C., Shu S., Berg A.P., Jevtovic-Todorovic V., Talley E.M., Chen X., Bayliss D.A. Motoneuronal TASK channels contribute to immobilizing effects of inhalational general anesthetics. J. Neurosci. 2010;30(22):7691–7704. doi: 10.1523/JNEUROSCI.1655-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Tibbs G.R., Rowley T.J., Sanford R.L., Herold K.F., Proekt A., Hemmings H.C., Jr, Andersen O.S., Goldstein P.A., Flood P.D. HCN1 channels as targets for anesthetic and nonanesthetic propofol analogs in the amelioration of mechanical and thermal hyperalgesia in a mouse model of neuropathic pain. J. Pharmacol. Exp. Ther. 2013;345(3):363–373. doi: 10.1124/jpet.113.203620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Zhou C., Liang P., Liu J., Ke B., Wang X., Li F., Li T., Bayliss D.A., Chen X. HCN1 channels contribute to the effects of amnesia and hypnosis but not immobility of volatile anesthetics. Anesth. Analg. 2015;121(3):661–666. doi: 10.1213/ANE.0000000000000830. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Gao J., Hu Z., Shi L., Li N., Ouyang Y., Shu S., Yao S., Chen X. HCN channels contribute to the sensitivity of intravenous anesthetics in developmental mice. Oncotarget. 2018;9(16):12907–12917. doi: 10.18632/oncotarget.24408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Antkowiak B. In vitro networks: cortical mechanisms of anaesthetic action. Br. J. Anaesth. 2002;89(1):102–111. doi: 10.1093/bja/aef154. [DOI] [PubMed] [Google Scholar]
- 163.Linden A.M., Aller M.I., Leppä E., Rosenberg P.H., Wisden W., Korpi E.R.K. + channel TASK-1 knockout mice show enhanced sensitivities to ataxic and hypnotic effects of GABA(A) receptor ligands. J. Pharmacol. Exp. Ther. 2008;327(1):277–286. doi: 10.1124/jpet.108.142083. [DOI] [PubMed] [Google Scholar]
- 164.Chen X., Shu S., Bayliss D.A. HCN1 channel subunits are a molecular substrate for hypnotic actions of ketamine. J. Neurosci. 2009;29(3):600–609. doi: 10.1523/JNEUROSCI.3481-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Zhou C., Douglas J.E., Kumar N.N., Shu S., Bayliss D.A., Chen X. Forebrain HCN1 channels contribute to hypnotic actions of ketamine. Anesthesiology. 2013;118(4):785–795. doi: 10.1097/ALN.0b013e318287b7c8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Chen Y.A., Scheller R.H. SNARE-mediated membrane fusion. Nat. Rev. Mol. Cell Biol. 2001;2(2):98–106. doi: 10.1038/35052017. [DOI] [PubMed] [Google Scholar]
- 167.Chen Y.A., Scales S.J., Scheller R.H. Sequential SNARE assembly underlies priming and triggering of exocytosis. Neuron. 2001;30(1):161–170. doi: 10.1016/S0896-6273(01)00270-7. [DOI] [PubMed] [Google Scholar]
- 168.Söllner T., Bennett M.K., Whiteheart S.W., Scheller R.H., Rothman J.E. A protein assembly-disassembly pathway in vitro that may correspond to sequential steps of synaptic vesicle docking, activation, and fusion. Cell. 1993;75(3):409–418. doi: 10.1016/0092-8674(93)90376-2. [DOI] [PubMed] [Google Scholar]
- 169.Hawasli A.H., Saifee O., Liu C., Nonet M.L., Crowder C.M. Resistance to volatile anesthetics by mutations enhancing excitatory neurotransmitter release in Caenorhabditis elegans. Genetics. 2004;168(2):831–843. doi: 10.1534/genetics.104.030502. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Zalucki O.H., Menon H., Kottler B., Faville R., Day R., Bademosi A.T., Lavidis N., Karunanithi S., van Swinderen B. Syntaxin1A-mediated resistance and hypersensitivity to isoflurane in drosophila melanogaster. Anesthesiology. 2015;122(5):1060–1074. doi: 10.1097/ALN.0000000000000629. [DOI] [PubMed] [Google Scholar]
- 171.Nagele P., Mendel J.B., Placzek W.J., Scott B.A., D’Avignon D.A., Crowder C.M. Volatile anesthetics bind rat synaptic snare proteins. Anesthesiology. 2005;103(4):768–778. doi: 10.1097/00000542-200510000-00015. [DOI] [PubMed] [Google Scholar]
- 172.Herring B.E., Xie Z., Marks J., Fox A.P. Isoflurane inhibits the neurotransmitter release machinery. J. Neurophysiol. 2009;102(2):1265–1273. doi: 10.1152/jn.00252.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Xie Z., McMillan K., Pike C.M., Cahill A.L., Herring B.E., Wang Q., Fox A.P. Interaction of anesthetics with neurotransmitter release machinery proteins. J. Neurophysiol. 2013;109(3):758–767. doi: 10.1152/jn.00666.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Bademosi A.T., Steeves J., Karunanithi S., Zalucki O.H., Gormal R.S., Liu S., Lauwers E., Verstreken P., Anggono V., Meunier F.A., van Swinderen B. Trapping of syntaxin1a in presynaptic nanoclusters by a clinically relevant general anesthetic. Cell Rep. 2018;22(2):427–440. doi: 10.1016/j.celrep.2017.12.054. [DOI] [PubMed] [Google Scholar]
- 175.Herring B.E., McMillan K., Pike C.M., Marks J., Fox A.P., Xie Z. Etomidate and propofol inhibit the neurotransmitter release machinery at different sites. J. Physiol. 2011;589(Pt 5):1103–1115. doi: 10.1113/jphysiol.2010.200964. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Spray D.C., Duffy H.S., Scemes E. Gap junctions in glia. Types, roles, and plasticity. Adv. Exp. Med. Biol. 1999;468:339–359. doi: 10.1007/978-1-4615-4685-6_27. [DOI] [PubMed] [Google Scholar]
- 177.Johnston M.F., Simon S.A., Ramón F. Interaction of anaesthetics with electrical synapses. Nature. 1980;286(5772):498–500. doi: 10.1038/286498a0. [DOI] [PubMed] [Google Scholar]
- 178.Mantz J., Cordier J., Giaume C. Effects of general anesthetics on intercellular communications mediated by gap junctions between astrocytes in primary culture. Anesthesiology. 1993;78(5):892–901. doi: 10.1097/00000542-199305000-00014. [DOI] [PubMed] [Google Scholar]
- 179.Wentlandt K., Carlen P.L., Kushnir M., Naus C.C., El-Beheiry H. General anesthetics attenuate gap junction coupling in P19 cell line. J. Neurosci. Res. 2005;81(5):746–752. doi: 10.1002/jnr.20577. [DOI] [PubMed] [Google Scholar]
- 180.Masaki E., Kawamura M., Kato F. Attenuation of gap-junction-mediated signaling facilitated anesthetic effect of sevoflurane in the central nervous system of rats. Anesth. Analg. 2004;98(3):647–652. doi: 10.1213/01.ANE.0000103259.72635.72. [DOI] [PubMed] [Google Scholar]
- 181.Wentlandt K., Samoilova M., Carlen P.L., El Beheiry H. General anesthetics inhibit gap junction communication in cultured organotypic hippocampal slices. Anesth. Analg. 2006;102(6):1692–1698. doi: 10.1213/01.ane.0000202472.41103.78. [DOI] [PubMed] [Google Scholar]
- 182.Jacobson G.M., Voss L.J., Melin S.M., Cursons R.T., Sleigh J.W. The role of connexin36 gap junctions in modulating the hypnotic effects of isoflurane and propofol in mice. Anaesthesia. 2011;66(5):361–367. doi: 10.1111/j.1365-2044.2011.06658.x. [DOI] [PubMed] [Google Scholar]
- 183.Rudolph U., Antkowiak B. Molecular and neuronal substrates for general anaesthetics. Nat. Rev. Neurosci. 2004;5(9):709–720. doi: 10.1038/nrn1496. [DOI] [PubMed] [Google Scholar]
- 184.Allison D.W., Gelfand V.I., Spector I., Craig A.M. Role of actin in anchoring postsynaptic receptors in cultured hippocampal neurons: differential attachment of NMDA versus AMPA receptors. J. Neurosci. 1998;18(7):2423–2436. doi: 10.1523/JNEUROSCI.18-07-02423.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.van Rossum D., Hanisch U.K. Cytoskeletal dynamics in dendritic spines: direct modulation by glutamate receptors? Trends Neurosci. 1999;22(7):290–295. doi: 10.1016/S0166-2236(99)01404-6. [DOI] [PubMed] [Google Scholar]
- 186.Kaech S., Brinkhaus H., Matus A. Volatile anesthetics block actin-based motility in dendritic spines. Proc. Natl. Acad. Sci. USA. 1999;96(18):10433–10437. doi: 10.1073/pnas.96.18.10433. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Hirota K., Roth S.H. Sevoflurane modulates both GABAA and GABAB receptors in area CA1 of rat hippocampus. Br. J. Anaesth. 1997;78(1):60–65. doi: 10.1093/bja/78.1.60. [DOI] [PubMed] [Google Scholar]
- 188.Antkowiak B. Different actions of general anesthetics on the firing patterns of neocortical neurons mediated by the GABA(A) receptor. Anesthesiology. 1999;91(2):500–511. doi: 10.1097/00000542-199908000-00025. [DOI] [PubMed] [Google Scholar]
- 189.Nishikawa K., MacIver M.B. Agent-selective effects of volatile anesthetics on GABAA receptor-mediated synaptic inhibition in hippocampal interneurons. Anesthesiology. 2001;94(2):340–347. doi: 10.1097/00000542-200102000-00025. [DOI] [PubMed] [Google Scholar]
- 190.Kitamura A., Sato R., Marszalec W., Yeh J.Z., Ogawa R., Narahashi T. Halothane and propofol modulation of gamma-aminobutyric acidA receptor single-channel currents. Anesth. Analg. 2004;99(2):409–415. doi: 10.1213/01.ANE.0000131969.46439.71. [DOI] [PubMed] [Google Scholar]
- 191.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. 2005;21(1):93–102. doi: 10.1111/j.1460-9568.2004.03843.x. [DOI] [PubMed] [Google Scholar]
- 192.Olsen R.W., Tobin A.J. Molecular biology of GABAA receptors. FASEB J. 1990;4(5):1469–1480. doi: 10.1096/fasebj.4.5.2155149. [DOI] [PubMed] [Google Scholar]
- 193.Sieghart W. GABAA receptors: ligand-gated Cl- ion channels modulated by multiple drug-binding sites. Trends Pharmacol. Sci. 1992;13(12):446–450. doi: 10.1016/0165-6147(92)90142-S. [DOI] [PubMed] [Google Scholar]
- 194.Garcia P.S., Kolesky S.E., Jenkins A. General anesthetic actions on GABA(A) receptors. Curr. Neuropharmacol. 2010;8(1):2–9. doi: 10.2174/157015910790909502. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Nakahiro M., Yeh J.Z., Brunner E., Narahashi T. General anesthetics modulate GABA receptor channel complex in rat dorsal root ganglion neurons. FASEB J. 1989;3(7):1850–1854. doi: 10.1096/fasebj.3.7.2541038. [DOI] [PubMed] [Google Scholar]
- 196.Wakamori M., Ikemoto Y., Akaike N. Effects of two volatile anesthetics and a volatile convulsant on the excitatory and inhibitory amino acid responses in dissociated CNS neurons of the rat. J. Neurophysiol. 1991;66(6):2014–2021. doi: 10.1152/jn.1991.66.6.2014. [DOI] [PubMed] [Google Scholar]
- 197.Jones M.V., Brooks P.A., Harrison N.L. Enhancement of gamma-aminobutyric acid-activated Cl- currents in cultured rat hippocampal neurones by three volatile anaesthetics. J. Physiol. 1992;449:279–293. doi: 10.1113/jphysiol.1992.sp019086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Li X., Czajkowski C., Pearce R.A. Rapid and direct modulation of GABAA receptors by halothane. Anesthesiology. 2000;92(5):1366–1375. doi: 10.1097/00000542-200005000-00027. [DOI] [PubMed] [Google Scholar]
- 199.Hales T.G., Lambert J.J. The actions of propofol on inhibitory amino acid receptors of bovine adrenomedullary chromaffin cells and rodent central neurones. Br. J. Pharmacol. 1991;104(3):619–628. doi: 10.1111/j.1476-5381.1991.tb12479.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Uchida I., Kamatchi G., Burt D., Yang J. Etomidate potentiation of GABAA receptor gated current depends on the subunit composition. Neurosci. Lett. 1995;185(3):203–206. doi: 10.1016/0304-3940(95)11263-V. [DOI] [PubMed] [Google Scholar]
- 201.Olsen R.W., Yang J., King R.G., Dilber A., Stauber G.B., Ransom R.W. Barbiturate and benzodiazepine modulation of GABA receptor binding and function. Life Sci. 1986;39(21):1969–1976. doi: 10.1016/0024-3205(86)90320-6. [DOI] [PubMed] [Google Scholar]
- 202.Olsen R.W., Sapp D.M., Bureau M.H., Turner D.M., Kokka N. Allosteric actions of central nervous system depressants including anesthetics on subtypes of the inhibitory gamma-aminobutyric acidA receptor-chloride channel complex. Ann. N. Y. Acad. Sci. 1991;625:145–154. doi: 10.1111/j.1749-6632.1991.tb33838.x. [DOI] [PubMed] [Google Scholar]
- 203.Zhang Y., Stabernack C., Sonner J., Dutton R., Eger E.I. II Both cerebral GABA(A) receptors and spinal GABA(A) receptors modulate the capacity of isoflurane to produce immobility. Anesth. Analg. 2001;92(6):1585–1589. doi: 10.1097/00000539-200106000-00047. [DOI] [PubMed] [Google Scholar]
- 204.Lam D.W., Reynolds J.N. Modulatory and direct effects of propofol on recombinant GABAA receptors expressed in xenopus oocytes: influence of alpha- and gamma2-subunits. Brain Res. 1998;784(1-2):179–187. doi: 10.1016/S0006-8993(97)01334-6. [DOI] [PubMed] [Google Scholar]
- 205.Hara M., Kai Y., Ikemoto Y. Propofol activates GABAA receptor-chloride ionophore complex in dissociated hippocampal pyramidal neurons of the rat. Anesthesiology. 1993;79(4):781–788. doi: 10.1097/00000542-199310000-00021. [DOI] [PubMed] [Google Scholar]
- 206.Hara M., Kai Y., Ikemoto Y. Enhancement by propofol of the gamma-aminobutyric acidA response in dissociated hippocampal pyramidal neurons of the rat. Anesthesiology. 1994;81(4):988–994. doi: 10.1097/00000542-199410000-00026. [DOI] [PubMed] [Google Scholar]
- 207.Orser B.A., Wang L.Y., Pennefather P.S., MacDonald J.F. Propofol modulates activation and desensitization of GABAA receptors in cultured murine hippocampal neurons. J. Neurosci. 1994;14(12):7747–7760. doi: 10.1523/JNEUROSCI.14-12-07747.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Bai D., Pennefather P.S., MacDonald J.F., Orser B.A. The general anesthetic propofol slows deactivation and desensitization of GABA(A) receptors. J. Neurosci. 1999;19(24):10635–10646. doi: 10.1523/JNEUROSCI.19-24-10635.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.Adodra S., Hales T.G. Potentiation, activation and blockade of GABAA receptors of clonal murine hypothalamic GT1-7 neurones by propofol. Br. J. Pharmacol. 1995;115(6):953–960. doi: 10.1111/j.1476-5381.1995.tb15903.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Thyagarajan R., Ramanjaneyulu R., Ticku M.K. Enhancement of diazepam and gamma-aminobutyric acid binding by (+)etomidate and pentobarbital. J. Neurochem. 1983;41(2):578–585. doi: 10.1111/j.1471-4159.1983.tb04778.x. [DOI] [PubMed] [Google Scholar]
- 211.Ashton D., Wauquier A. Modulation of a GABA-ergic inhibitory circuit in the in vitro hippocampus by etomidate isomers. Anesth. Analg. 1985;64(10):975–980. doi: 10.1213/00000539-198510000-00006. [DOI] [PubMed] [Google Scholar]
- 212.Proctor W.R., Mynlieff M., Dunwiddie T.V. Facilitatory action of etomidate and pentobarbital on recurrent inhibition in rat hippocampal pyramidal neurons. J. Neurosci. 1986;6(11):3161–3168. doi: 10.1523/JNEUROSCI.06-11-03161.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Yang J., Uchida I. Mechanisms of etomidate potentiation of GABAA receptor-gated currents in cultured postnatal hippocampal neurons. Neuroscience. 1996;73(1):69–78. doi: 10.1016/0306-4522(96)00018-8. [DOI] [PubMed] [Google Scholar]
- 214.Delgado-Lezama R., Loeza-Alcocer E., Andrés C., Aguilar J., Guertin P.A., Felix R. Extrasynaptic GABA(A) receptors in the brainstem and spinal cord: structure and function. Curr. Pharm. Des. 2013;19(24):4485–4497. doi: 10.2174/1381612811319240013. [DOI] [PubMed] [Google Scholar]
- 215.Kotani N., Akaike N. The effects of volatile anesthetics on synaptic and extrasynaptic GABA-induced neurotransmission. Brain Res. Bull. 2013;93:69–79. doi: 10.1016/j.brainresbull.2012.08.001. [DOI] [PubMed] [Google Scholar]
- 216.Topf N., Jenkins A., Baron N., Harrison N.L. Effects of isoflurane on gamma-aminobutyric acid type A receptors activated by full and partial agonists. Anesthesiology. 2003;98(2):306–311. doi: 10.1097/00000542-200302000-00007. [DOI] [PubMed] [Google Scholar]
- 217.Olsen R.W., Li G.D. GABA(A) receptors as molecular targets of general anesthetics: identification of binding sites provides clues to allosteric modulation. Can. J. Anaesth. 2011;58(2):206–215. doi: 10.1007/s12630-010-9429-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218.Bai D., Zhu G., Pennefather P., Jackson M.F., MacDonald J.F., Orser B.A. Distinct functional and pharmacological properties of tonic and quantal inhibitory postsynaptic currents mediated by gamma-aminobutyric acid(A) receptors in hippocampal neurons. Mol. Pharmacol. 2001;59(4):814–824. doi: 10.1124/mol.59.4.814. [DOI] [PubMed] [Google Scholar]
- 219.Burt D.R., Kamatchi G.L. GABAA receptor subtypes: from pharmacology to molecular biology. FASEB J. 1991;5(14):2916–2923. doi: 10.1096/fasebj.5.14.1661244. [DOI] [PubMed] [Google Scholar]
- 220.Macdonald R.L., Olsen R.W. GABAA receptor channels. Annu. Rev. Neurosci. 1994;17:569–602. doi: 10.1146/annurev.ne.17.030194.003033. [DOI] [PubMed] [Google Scholar]
- 221.Ogurusu T., Shingai R. Cloning of a putative gamma-aminobutyric acid (GABA) receptor subunit rho 3 cDNA. Biochim. Biophys. Acta. 1996;1305(1-2):15–18. doi: 10.1016/0167-4781(95)00205-7. [DOI] [PubMed] [Google Scholar]
- 222.Horne A.L., Harkness P.C., Hadingham K.L., Whiting P., Kemp J.A. The influence of the gamma 2L subunit on the modulation of responses to GABAA receptor activation. Br. J. Pharmacol. 1993;108(3):711–716. doi: 10.1111/j.1476-5381.1993.tb12866.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223.Blair L.A., Levitan E.S., Marshall J., Dionne V.E., Barnard E.A. Single subunits of the GABAA receptor form ion channels with properties of the native receptor. Science. 1988;242(4878):577–579. doi: 10.1126/science.2845583. [DOI] [PubMed] [Google Scholar]
- 224.Sigel E., Baur R., Trube G., Möhler H., Malherbe P. The effect of subunit composition of rat brain GABAA receptors on channel function. Neuron. 1990;5(5):703–711. doi: 10.1016/0896-6273(90)90224-4. [DOI] [PubMed] [Google Scholar]
- 225.Angelotti T.P., Macdonald R.L. Assembly of GABAA receptor subunits: alpha 1 beta 1 and alpha 1 beta 1 gamma 2S subunits produce unique ion channels with dissimilar single-channel properties. J. Neurosci. 1993;13(4):1429–1440. doi: 10.1523/JNEUROSCI.13-04-01429.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Im H.K., Im W.B., Carter D.B., McKinley D.D. Interaction of beta-carboline inverse agonists for the benzodiazepine site with another site on GABAA receptors. Br. J. Pharmacol. 1995;114(5):1040–1044. doi: 10.1111/j.1476-5381.1995.tb13310.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Chang Y., Wang R., Barot S., Weiss D.S. Stoichiometry of a recombinant GABAA receptor. J. Neurosci. 1996;16(17):5415–5424. doi: 10.1523/JNEUROSCI.16-17-05415.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Lolait S.J., O’Carroll A.M., Kusano K., Mahan L.C. Pharmacological characterization and region-specific expression in brain of the beta 2- and beta 3-subunits of the rat GABAA receptor. FEBS Lett. 1989;258(1):17–21. doi: 10.1016/0014-5793(89)81605-9. [DOI] [PubMed] [Google Scholar]
- 229.Ymer S., Schofield P.R., Draguhn A., Werner P., Köhler M., Seeburg P.H. GABAA receptor beta subunit heterogeneity: functional expression of cloned cDNAs. EMBO J. 1989;8(6):1665–1670. doi: 10.1002/j.1460-2075.1989.tb03557.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Valeyev A.Y., Barker J.L., Cruciani R.A., Lange G.D., Smallwood V.V., Mahan L.C. Characterization of the gamma-aminobutyric acidA receptor-channel complex composed of alpha 1 beta 2 and alpha 1 beta 3 subunits from rat brain. J. Pharmacol. Exp. Ther. 1993;265(2):985–991. [PubMed] [Google Scholar]
- 231.Liu K., Jounaidi Y., Forman S.A., Feng H.J. Etomidate uniquely modulates the desensitization of recombinant α1β3δ GABA(A) receptors. Neuroscience. 2015;300:307–313. doi: 10.1016/j.neuroscience.2015.05.051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Pritchett D.B., Sontheimer H., Shivers B.D., Ymer S., Kettenmann H., Schofield P.R., Seeburg P.H. Importance of a novel GABAA receptor subunit for benzodiazepine pharmacology. Nature. 1989;338(6216):582–585. doi: 10.1038/338582a0. [DOI] [PubMed] [Google Scholar]
- 233.Jones M.V., Harrison N.L., Pritchett D.B., Hales T.G. Modulation of the GABAA receptor by propofol is independent of the gamma subunit. J. Pharmacol. Exp. Ther. 1995;274(2):962–968. [PubMed] [Google Scholar]
- 234.Sigel E., Baur R., Malherbe P., Möhler H. The rat beta 1-subunit of the GABAA receptor forms a picrotoxin-sensitive anion channel open in the absence of GABA. FEBS Lett. 1989;257(2):377–379. doi: 10.1016/0014-5793(89)81576-5. [DOI] [PubMed] [Google Scholar]
- 235.Krishek B.J., Moss S.J., Smart T.G. Homomeric beta 1 gamma-aminobutyric acid A receptor-ion channels: evaluation of pharmacological and physiological properties. Mol. Pharmacol. 1996;49(3):494–504. [PubMed] [Google Scholar]
- 236.Cestari I.N., Uchida I., Li L., Burt D., Yang J. The agonistic action of pentobarbital on GABAA beta-subunit homomeric receptors. Neuroreport. 1996;7(4):943–947. doi: 10.1097/00001756-199603220-00023. [DOI] [PubMed] [Google Scholar]
- 237.Davies P.A., Hanna M.C., Hales T.G., Kirkness E.F. Insensitivity to anaesthetic agents conferred by a class of GABA(A) receptor subunit. Nature. 1997;385(6619):820–823. doi: 10.1038/385820a0. [DOI] [PubMed] [Google Scholar]
- 238.Sanna E., Garau F., Harris R.A. Novel properties of homomeric beta 1 gamma-aminobutyric acid type A receptors: actions of the anesthetics propofol and pentobarbital. Mol. Pharmacol. 1995;47(2):213–217. [PubMed] [Google Scholar]
- 239.Williams D.B., Akabas M.H. Structural evidence that propofol stabilizes different GABA(A) receptor states at potentiating and activating concentrations. J. Neurosci. 2002;22(17):7417–7424. doi: 10.1523/JNEUROSCI.22-17-07417.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Krasowski M.D., Koltchine V.V., Rick C.E., Ye Q., Finn S.E., Harrison N.L. Propofol and other intravenous anesthetics have sites of action on the gamma-aminobutyric acid type A receptor distinct from that for isoflurane. Mol. Pharmacol. 1998;53(3):530–538. doi: 10.1124/mol.53.3.530. [DOI] [PubMed] [Google Scholar]
- 241.Siegwart R., Jurd R., Rudolph U. Molecular determinants for the action of general anesthetics at recombinant alpha(2)beta(3)gamma(2)gamma-aminobutyric acid(A) receptors. J. Neurochem. 2002;80(1):140–148. doi: 10.1046/j.0022-3042.2001.00682.x. [DOI] [PubMed] [Google Scholar]
- 242.Lor C., Perouansky M., Pearce R.A. Isoflurane potentiation of GABAA receptors is reduced but not eliminated by the β3(n265m) mutation. Int. J. Mol. Sci. 2020;21(24):E9534. doi: 10.3390/ijms21249534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243.Drexler B., Jurd R., Rudolph U., Antkowiak B. Distinct actions of etomidate and propofol at beta3-containing gamma-aminobutyric acid type A receptors. Neuropharmacology. 2009;57(4):446–455. doi: 10.1016/j.neuropharm.2009.06.014. [DOI] [PubMed] [Google Scholar]
- 244.Jurd R., Arras M., Lambert S., Drexler B., Siegwart R., Crestani F., Zaugg M., Vogt K.E., Ledermann B., Antkowiak B., Rudolph U. General anesthetic actions in vivo strongly attenuated by a point mutation in the GABA(A) receptor beta3 subunit. FASEB J. 2003;17(2):250–252. doi: 10.1096/fj.02-0611fje. [DOI] [PubMed] [Google Scholar]
- 245.Reynolds D.S., Rosahl T.W., Cirone J., O’Meara G.F., Haythornthwaite A., Newman R.J., Myers J., Sur C., Howell O., Rutter A.R., Atack J., Macaulay A.J., Hadingham K.L., Hutson P.H., Belelli D., Lambert J.J., Dawson G.R., McKernan R., Whiting P.J., Wafford K.A. Sedation and anesthesia mediated by distinct GABA(A) receptor isoforms. J. Neurosci. 2003;23(24):8608–8617. doi: 10.1523/JNEUROSCI.23-24-08608.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.O’Meara G.F., Newman R.J., Fradley R.L., Dawson G.R., Reynolds D.S. The GABA-A beta3 subunit mediates anaesthesia induced by etomidate. Neuroreport. 2004;15(10):1653–1656. doi: 10.1097/01.wnr.0000134842.56131.fe. [DOI] [PubMed] [Google Scholar]
- 247.Carlson B.X., Belhage B., Hansen G.H., Elster L., Olsen R.W., Schousboe A. Expression of the GABA(A) receptor alpha6 subunit in cultured cerebellar granule cells is developmentally regulated by activation of GABA(A) receptors. J. Neurosci. Res. 1997;50(6):1053–1062. doi: 10.1002/(SICI)1097-4547(19971215)50:6<1053:AID-JNR17>3.0.CO;2-5. [DOI] [PubMed] [Google Scholar]
- 248.Quinlan J.J., Homanics G.E., Firestone L.L. Anesthesia sensitivity in mice that lack the beta3 subunit of the gamma-aminobutyric acid type A receptor. Anesthesiology. 1998;88(3):775–780. doi: 10.1097/00000542-199803000-00030. [DOI] [PubMed] [Google Scholar]
- 249.Lambert S., Arras M., Vogt K.E., Rudolph U. Isoflurane-induced surgical tolerance mediated only in part by beta3-containing GABA(A) receptors. Eur. J. Pharmacol. 2005;516(1):23–27. doi: 10.1016/j.ejphar.2005.04.030. [DOI] [PubMed] [Google Scholar]
- 250.Liao M., Sonner J.M., Husain S.S., Miller K.W., Jurd R., Rudolph U., Eger E.I. II R (+) etomidate and the photoactivable R (+) azietomidate have comparable anesthetic activity in wild-type mice and comparably decreased activity in mice with a N265M point mutation in the gamma-aminobutyric acid receptor beta3 subunit. Anesth. Analg. 2005;101(1):131–135. doi: 10.1213/01.ANE.0000153011.64764.6F. [DOI] [PubMed] [Google Scholar]
- 251.Akeju O., Hamilos A.E., Song A.H., Pavone K.J., Purdon P.L., Brown E.N. GABAA circuit mechanisms are associated with ether anesthesia-induced unconsciousness. Clin. Neurophysiol. 2016;127(6):2472–2481. doi: 10.1016/j.clinph.2016.02.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Rajendra S., Vandenberg R.J., Pierce K.D., Cunningham A.M., French P.W., Barry P.H., Schofield P.R. The unique extracellular disulfide loop of the glycine receptor is a principal ligand binding element. EMBO J. 1995;14(13):2987–2998. doi: 10.1002/j.1460-2075.1995.tb07301.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Birnir B., Tierney M.L., Lim M., Cox G.B., Gage P.W. Nature of the 5′ residue in the M2 domain affects function of the human alpha 1 beta 1 GABAA receptor. Synapse. 1997;26(3):324–327. doi: 10.1002/(SICI)1098-2396(199707)26:3<324:AID-SYN13>3.0.CO;2-V. [DOI] [PubMed] [Google Scholar]
- 254.Krasowski M.D., Nishikawa K., Nikolaeva N., Lin A., Harrison N.L. Methionine 286 in transmembrane domain 3 of the GABAA receptor beta subunit controls a binding cavity for propofol and other alkylphenol general anesthetics. Neuropharmacology. 2001;41(8):952–964. doi: 10.1016/S0028-3908(01)00141-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255.Bali M., Akabas M.H. Defining the propofol binding site location on the GABAA receptor. Mol. Pharmacol. 2004;65(1):68–76. doi: 10.1124/mol.65.1.68. [DOI] [PubMed] [Google Scholar]
- 256.Chang C.S., Olcese R., Olsen R.W. A single M1 residue in the beta2 subunit alters channel gating of GABAA receptor in anesthetic modulation and direct activation. J. Biol. Chem. 2003;278(44):42821–42828. doi: 10.1074/jbc.M306978200. [DOI] [PubMed] [Google Scholar]
- 257.Mihic S.J., Ye Q., Wick M.J., Koltchine V.V., Krasowski M.D., Finn S.E., Mascia M.P., Valenzuela C.F., Hanson K.K., Greenblatt E.P., Harris R.A., Harrison N.L. Sites of alcohol and volatile anaesthetic action on GABA(A) and glycine receptors. Nature. 1997;389(6649):385–389. doi: 10.1038/38738. [DOI] [PubMed] [Google Scholar]
- 258.Werner D.F., Swihart A., Rau V., Jia F., Borghese C.M., McCracken M.L., Iyer S., Fanselow M.S., Oh I., Sonner J.M., Eger E.I., II, Harrison N.L., Harris R.A., Homanics G.E. Inhaled anesthetic responses of recombinant receptors and knockin mice harboring α2(S270H/L277A) GABA(A) receptor subunits that are resistant to isoflurane. J. Pharmacol. Exp. Ther. 2011;336(1):134–144. doi: 10.1124/jpet.110.170431. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Nishikawa K., Jenkins A., Paraskevakis I., Harrison N.L. Volatile anesthetic actions on the GABAA receptors: contrasting effects of alpha 1(S270) and beta 2(N265) point mutations. Neuropharmacology. 2002;42(3):337–345. doi: 10.1016/S0028-3908(01)00189-7. [DOI] [PubMed] [Google Scholar]
- 260.Sonner J.M., Werner D.F., Elsen F.P., Xing Y., Liao M., Harris R.A., Harrison N.L., Fanselow M.S., Eger E.I., II, Homanics G.E. Effect of isoflurane and other potent inhaled anesthetics on minimum alveolar concentration, learning, and the righting reflex in mice engineered to express alpha1 gamma-aminobutyric acid type A receptors unresponsive to isoflurane. Anesthesiology. 2007;106(1):107–113. doi: 10.1097/00000542-200701000-00019. [DOI] [PubMed] [Google Scholar]
- 261.Ying S.W., Werner D.F., Homanics G.E., Harrison N.L., Goldstein P.A. Isoflurane modulates excitability in the mouse thalamus via GABA-dependent and GABA-independent mechanisms. Neuropharmacology. 2009;56(2):438–447. doi: 10.1016/j.neuropharm.2008.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262.Rau V., Iyer S.V., Oh I., Chandra D., Harrison N., Eger E.I., II, Fanselow M.S., Homanics G.E., Sonner J.M. Gamma-aminobutyric acid type A receptor alpha 4 subunit knockout mice are resistant to the amnestic effect of isoflurane. Anesth. Analg. 2009;109(6):1816–1822. doi: 10.1213/ANE.0b013e3181bf6ae6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Sun C., Sieghart W., Kapur J. Distribution of alpha1, alpha4, gamma2, and delta subunits of GABAA receptors in hippocampal granule cells. Brain Res. 2004;1029(2):207–216. doi: 10.1016/j.brainres.2004.09.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264.Benkwitz C., Banks M.I., Pearce R.A. Influence of GABAA receptor gamma2 splice variants on receptor kinetics and isoflurane modulation. Anesthesiology. 2004;101(4):924–936. doi: 10.1097/00000542-200410000-00018. [DOI] [PubMed] [Google Scholar]
- 265.Caraiscos V.B., Newell J.G., You-Ten K.E., Elliott E.M., Rosahl T.W., Wafford K.A., MacDonald J.F., Orser B.A. Selective enhancement of tonic GABAergic inhibition in murine hippocampal neurons by low concentrations of the volatile anesthetic isoflurane. J. Neurosci. 2004;24(39):8454–8458. doi: 10.1523/JNEUROSCI.2063-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Cheng V.Y., Martin L.J., Elliott E.M., Kim J.H., Mount H.T., Taverna F.A., Roder J.C., Macdonald J.F., Bhambri A., Collinson N., Wafford K.A., Orser B.A. Alpha5GABAA receptors mediate the amnestic but not sedative-hypnotic effects of the general anesthetic etomidate. J. Neurosci. 2006;26(14):3713–3720. doi: 10.1523/JNEUROSCI.5024-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267.Bieda M.C., Su H., Maciver M.B. Anesthetics discriminate between tonic and phasic gamma-aminobutyric acid receptors on hippocampal CA1 neurons. Anesth. Analg. 2009;108(2):484–490. doi: 10.1213/ane.0b013e3181904571. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 268.Ogawa S.K., Tanaka E., Shin M.C., Kotani N., Akaike N. Volatile anesthetic effects on isolated GABA synapses and extrasynaptic receptors. Neuropharmacology. 2011;60(4):701–710. doi: 10.1016/j.neuropharm.2010.11.016. [DOI] [PubMed] [Google Scholar]
- 269.Dai S., Perouansky M., Pearce R.A. Isoflurane enhances both fast and slow synaptic inhibition in the hippocampus at amnestic concentrations. Anesthesiology. 2012;116(4):816–823. doi: 10.1097/ALN.0b013e31824be0e3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270.Collinson N., Kuenzi F.M., Jarolimek W., Maubach K.A., Cothliff R., Sur C., Smith A., Otu F.M., Howell O., Atack J.R., McKernan R.M., Seabrook G.R., Dawson G.R., Whiting P.J., Rosahl T.W. Enhanced learning and memory and altered GABAergic synaptic transmission in mice lacking the alpha 5 subunit of the GABAA receptor. J. Neurosci. 2002;22(13):5572–5580. doi: 10.1523/JNEUROSCI.22-13-05572.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271.Ortells M.O., Lunt G.G. Evolutionary history of the ligand-gated ion-channel superfamily of receptors. Trends Neurosci. 1995;18(3):121–127. doi: 10.1016/0166-2236(95)93887-4. [DOI] [PubMed] [Google Scholar]
- 272.Tassonyi E., Charpantier E., Muller D., Dumont L., Bertrand D. The role of nicotinic acetylcholine receptors in the mechanisms of anesthesia. Brain Res. Bull. 2002;57(2):133–150. doi: 10.1016/S0361-9230(01)00740-7. [DOI] [PubMed] [Google Scholar]
- 273.Violet J.M., Downie D.L., Nakisa R.C., Lieb W.R., Franks N.P. Differential sensitivities of mammalian neuronal and muscle nicotinic acetylcholine receptors to general anesthetics. Anesthesiology. 1997;86(4):866–874. doi: 10.1097/00000542-199704000-00017. [DOI] [PubMed] [Google Scholar]
- 274.Cooper E., Couturier S., Ballivet M. Pentameric structure and subunit stoichiometry of a neuronal nicotinic acetylcholine receptor. Nature. 1991;350(6315):235–238. doi: 10.1038/350235a0. [DOI] [PubMed] [Google Scholar]
- 275.Role L.W., Berg D.K. Nicotinic receptors in the development and modulation of CNS synapses. Neuron. 1996;16(6):1077–1085. doi: 10.1016/S0896-6273(00)80134-8. [DOI] [PubMed] [Google Scholar]
- 276.McGehee D.S., Heath M.J., Gelber S., Devay P., Role L.W. Nicotine enhancement of fast excitatory synaptic transmission in CNS by presynaptic receptors. Science. 1995;269(5231):1692–1696. doi: 10.1126/science.7569895. [DOI] [PubMed] [Google Scholar]
- 277.McGehee D.S., Role L.W. Physiological diversity of nicotinic acetylcholine receptors expressed by vertebrate neurons. Annu. Rev. Physiol. 1995;57:521–546. doi: 10.1146/annurev.ph.57.030195.002513. [DOI] [PubMed] [Google Scholar]
- 278.Arimura H., Ikemoto Y. Action of enflurane on cholinergic transmission in identified Aplysia neurones. Br. J. Pharmacol. 1986;89(3):573–582. doi: 10.1111/j.1476-5381.1986.tb11158.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 279.McKenzie D., Franks N.P., Lieb W.R. Actions of general anaesthetics on a neuronal nicotinic acetylcholine receptor in isolated identified neurones of Lymnaea stagnalis. Br. J. Pharmacol. 1995;115(2):275–282. doi: 10.1111/j.1476-5381.1995.tb15874.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 280.Yashima N., Wada A., Izumi F. Halothane inhibits the cholinergic-receptor-mediated influx of calcium in primary culture of bovine adrenal medulla cells. Anesthesiology. 1986;64(4):466–472. doi: 10.1097/00000542-198604000-00009. [DOI] [PubMed] [Google Scholar]
- 281.Pocock G., Richards C.D. The action of volatile anaesthetics on stimulus-secretion coupling in bovine adrenal chromaffin cells. Br. J. Pharmacol. 1988;95(1):209–217. doi: 10.1111/j.1476-5381.1988.tb16566.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282.Cardoso R.A., Yamakura T., Brozowski S.J., Chavez-Noriega L.E., Harris R.A. Human neuronal nicotinic acetylcholine receptors expressed in Xenopus oocytes predict efficacy of halogenated compounds that disobey the Meyer-Overton rule. Anesthesiology. 1999;91(5):1370–1377. doi: 10.1097/00000542-199911000-00029. [DOI] [PubMed] [Google Scholar]
- 283.Yamakura T., Chavez-Noriega L.E., Harris R.A. Subunit-dependent inhibition of human neuronal nicotinic acetylcholine receptors and other ligand-gated ion channels by dissociative anesthetics ketamine and dizocilpine. Anesthesiology. 2000;92(4):1144–1153. doi: 10.1097/00000542-200004000-00033. [DOI] [PubMed] [Google Scholar]
- 284.Yamashita M., Mori T., Nagata K., Yeh J.Z., Narahashi T. Isoflurane modulation of neuronal nicotinic acetylcholine receptors expressed in human embryonic kidney cells. Anesthesiology. 2005;102(1):76–84. doi: 10.1097/00000542-200501000-00015. [DOI] [PubMed] [Google Scholar]
- 285.Yamakura T., Borghese C., Harris R.A. A transmembrane site determines sensitivity of neuronal nicotinic acetylcholine receptors to general anesthetics. J. Biol. Chem. 2000;275(52):40879–40886. doi: 10.1074/jbc.M005771200. [DOI] [PubMed] [Google Scholar]
- 286.Mowrey D.D., Liu Q., Bondarenko V., Chen Q., Seyoum E., Xu Y., Wu J., Tang P. Insights into distinct modulation of α7 and α7β2 nicotinic acetylcholine receptors by the volatile anesthetic isoflurane. J. Biol. Chem. 2013;288(50):35793–35800. doi: 10.1074/jbc.M113.508333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 287.Bondarenko V., Mowrey D.D., Tillman T.S., Seyoum E., Xu Y., Tang P. NMR structures of the human α7 nAChR transmembrane domain and associated anesthetic binding sites. Biochim. Biophys. Acta. 2014;1838(5):1389–1395. doi: 10.1016/j.bbamem.2013.12.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288.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. 1997;86(4):859–865. doi: 10.1097/00000542-199704000-00016. [DOI] [PubMed] [Google Scholar]
- 289.Zhang L., Oz M., Stewart R.R., Peoples R.W., Weight F.F. Volatile general anaesthetic actions on recombinant nACh alpha 7, 5-HT3 and chimeric nACh alpha 7-5-HT3 receptors expressed in Xenopus oocytes. Br. J. Pharmacol. 1997;120(3):353–355. doi: 10.1038/sj.bjp.0700934. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 290.Mori T., Zhao X., Zuo Y., Aistrup G.L., Nishikawa K., Marszalec W., Yeh J.Z., Narahashi T. Modulation of neuronal nicotinic acetylcholine receptors by halothane in rat cortical neurons. Mol. Pharmacol. 2001;59(4):732–743. doi: 10.1124/mol.59.4.732. [DOI] [PubMed] [Google Scholar]
- 291.Liu L.T., Willenbring D., Xu Y., Tang P. General anesthetic binding to neuronal alpha4beta2 nicotinic acetylcholine receptor and its effects on global dynamics. J. Phys. Chem. B. 2009;113(37):12581–12589. doi: 10.1021/jp9039513. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 292.Eger E.I., II, Zhang Y., Laster M., Flood P., Kendig J.J., Sonner J.M. Acetylcholine receptors do not mediate the immobilization produced by inhaled anesthetics. Anesth. Analg. 2002;94(6):1500–1504. doi: 10.1097/00000539-200206000-00023. [DOI] [PubMed] [Google Scholar]
- 293.Flood P., Sonner J.M., Gong D., Coates K.M. Heteromeric nicotinic inhibition by isoflurane does not mediate MAC or loss of righting reflex. Anesthesiology. 2002;97(4):902–905. doi: 10.1097/00000542-200210000-00023. [DOI] [PubMed] [Google Scholar]
- 294.Zhang Y., Laster M.J., Eger E.I., II, Sharma M., Sonner J.M. Blockade of acetylcholine receptors does not change the dose of etomidate required to produce immobility in rats. Anesth. Analg. 2007;104(4):850–852. doi: 10.1213/01.ane.0000258018.82583.0b. [DOI] [PubMed] [Google Scholar]
- 295.Leung L.S., Petropoulos S., Shen B., Luo T., Herrick I., Rajakumar N., Ma J. Lesion of cholinergic neurons in nucleus basalis enhances response to general anesthetics. Exp. Neurol. 2011;228(2):259–269. doi: 10.1016/j.expneurol.2011.01.019. [DOI] [PubMed] [Google Scholar]
- 296.Mori H., Mishina M. Structure and function of the NMDA receptor channel. Neuropharmacology. 1995;34(10):1219–1237. doi: 10.1016/0028-3908(95)00109-J. [DOI] [PubMed] [Google Scholar]
- 297.Johnson J.W., Ascher P. Glycine potentiates the NMDA response in cultured mouse brain neurons. Nature. 1987;325(6104):529–531. doi: 10.1038/325529a0. [DOI] [PubMed] [Google Scholar]
- 298.Scheller M.S., Zornow M.H., Fleischer J.E., Shearman G.T., Greber T.F. The noncompetitive N-methyl-D-aspartate receptor antagonist, MK-801 profoundly reduces volatile anesthetic requirements in rabbits. Neuropharmacology. 1989;28(7):677–681. doi: 10.1016/0028-3908(89)90150-0. [DOI] [PubMed] [Google Scholar]
- 299.Daniell L.C. The noncompetitive N-methyl-D-aspartate antagonists, MK-801, phencyclidine and ketamine, increase the potency of general anesthetics. Pharmacol. Biochem. Behav. 1990;36(1):111–115. doi: 10.1016/0091-3057(90)90134-4. [DOI] [PubMed] [Google Scholar]
- 300.Yang J., Zorumski C.F. Effects of isoflurane on N-methyl-D-aspartate gated ion channels in cultured rat hippocampal neurons. Ann. N. Y. Acad. Sci. 1991;625:287–289. doi: 10.1111/j.1749-6632.1991.tb33851.x. [DOI] [PubMed] [Google Scholar]
- 301.Martin D.C., Abraham J.E., Plagenhoef M., Aronstam R.S. Volatile anesthetics and NMDA receptors. Enflurane inhibition of glutamate-stimulated [3H]MK-801 binding and reversal by glycine. Neurosci. Lett. 1991;132(1):73–76. doi: 10.1016/0304-3940(91)90436-W. [DOI] [PubMed] [Google Scholar]
- 302.Martin D.C., Aronstam R.S. Spermidine attenuation of volatile anesthetic inhibition of glutamate-stimulated [3H](5D,10S)-(+)-methyl-10,11-dihydro-5H- dibenzo[a,d]cyclohepten-5,10-imine ([3H]MK-801) binding to N-methyl-D-aspartate (NMDA) receptors in rat brain. Biochem. Pharmacol. 1995;50(9):1373–1377. doi: 10.1016/0006-2952(95)02017-9. [DOI] [PubMed] [Google Scholar]
- 303.Ishizaki K., Yoshida N., Yoon D.M., Yoon M.H., Sudoh M., Fujita T. Intrathecally administered NMDA receptor antagonists reduce the MAC of isoflurane in rats. Can. J. Anaesth. 1996;43(7):724–730. doi: 10.1007/BF03017958. [DOI] [PubMed] [Google Scholar]
- 304.Ishizaki K., Sasaki M., Karasawa S., Obata H., Nara T., Goto F. Intrathecal co-administration of NMDA antagonist and NK-1 antagonist reduces MAC of isoflurane in rats. Acta Anaesthesiol. Scand. 1999;43(7):753–759. doi: 10.1034/j.1399-6576.1999.430711.x. [DOI] [PubMed] [Google Scholar]
- 305.Carlà V., Moroni F. General anaesthetics inhibit the responses induced by glutamate receptor agonists in the mouse cortex. Neurosci. Lett. 1992;146(1):21–24. doi: 10.1016/0304-3940(92)90162-Z. [DOI] [PubMed] [Google Scholar]
- 306.Perouansky M., Kirson E.D., Yaari Y. Mechanism of action of volatile anesthetics: effects of halothane on glutamate receptors in vitro. Toxicol. Lett. 1998;100-101:65–69. doi: 10.1016/S0378-4274(98)00166-0. [DOI] [PubMed] [Google Scholar]
- 307.MacDonald J.F., Bartlett M.C., Mody I., Pahapill P., Reynolds J.N., Salter M.W., Schneiderman J.H., Pennefather P.S. Actions of ketamine, phencyclidine and MK-801 on NMDA receptor currents in cultured mouse hippocampal neurones. J. Physiol. 1991;432:483–508. doi: 10.1113/jphysiol.1991.sp018396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 308.Mayer M.L., Westbrook G.L., Vyklický L., Jr Sites of antagonist action on N-methyl-D-aspartic acid receptors studied using fluctuation analysis and a rapid perfusion technique. J. Neurophysiol. 1988;60(2):645–663. doi: 10.1152/jn.1988.60.2.645. [DOI] [PubMed] [Google Scholar]
- 309.Irifune M., Shimizu T., Nomoto M., Fukuda T. Ketamine-induced anesthesia involves the N-methyl-D-aspartate receptor-channel complex in mice. Brain Res. 1992;596(1-2):1–9. doi: 10.1016/0006-8993(92)91525-J. [DOI] [PubMed] [Google Scholar]
- 310.Orser B.A., Pennefather P.S., MacDonald J.F. Multiple mechanisms of ketamine blockade of N-methyl-D-aspartate receptors. Anesthesiology. 1997;86(4):903–917. doi: 10.1097/00000542-199704000-00021. [DOI] [PubMed] [Google Scholar]
- 311.Yamakura T., Sakimura K., Shimoji K., Mishina M. Effects of propofol on various AMPA-, kainate- and NMDA-selective glutamate receptor channels expressed in Xenopus oocytes. Neurosci. Lett. 1995;188(3):187–190. doi: 10.1016/0304-3940(95)11431-U. [DOI] [PubMed] [Google Scholar]
- 312.Bianchi M., Battistin T., Galzigna L. 2,6-diisopropylphenol, a general anesthetic, inhibits glutamate action on rat synaptosomes. Neurochem. Res. 1991;16(4):443–446. doi: 10.1007/BF00965564. [DOI] [PubMed] [Google Scholar]
- 313.Orser B.A., Bertlik M., Wang L.Y., MacDonald J.F. Inhibition by propofol (2,6 di-isopropylphenol) of the N-methyl-D-aspartate subtype of glutamate receptor in cultured hippocampal neurones. Br. J. Pharmacol. 1995;116(2):1761–1768. doi: 10.1111/j.1476-5381.1995.tb16660.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 314.Laube B., Kuhse J., Betz H. Evidence for a tetrameric structure of recombinant NMDA receptors. J. Neurosci. 1998;18(8):2954–2961. doi: 10.1523/JNEUROSCI.18-08-02954.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 315.Gan Q., Salussolia C.L., Wollmuth L.P. Assembly of AMPA receptors: mechanisms and regulation. J. Physiol. 2015;593(1):39–48. doi: 10.1113/jphysiol.2014.273755. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 316.Hollmann M.W., Liu H.T., Hoenemann C.W., Liu W.H., Durieux M.E. Modulation of NMDA receptor function by ketamine and magnesium. Part II: interactions with volatile anesthetics. Anesth. Analg. 2001;92(5):1182–1191. doi: 10.1097/00000539-200105000-00020. [DOI] [PubMed] [Google Scholar]
- 317.Ogata J., Shiraishi M., Namba T., Smothers C.T., Woodward J.J., Harris R.A. Effects of anesthetics on mutant N-methyl-D-aspartate receptors expressed in Xenopus oocytes. J. Pharmacol. Exp. Ther. 2006;318(1):434–443. doi: 10.1124/jpet.106.101691. [DOI] [PubMed] [Google Scholar]
- 318.Solt K., Eger E.I., II, Raines D.E. Differential modulation of human N-methyl-D-aspartate receptors by structurally diverse general anesthetics. Anesth. Analg. 2006;102(5):1407–1411. doi: 10.1213/01.ane.0000204252.07406.9f. [DOI] [PubMed] [Google Scholar]
- 319.Petrenko A.B., Yamakura T., Fujiwara N., Askalany A.R., Baba H., Sakimura K. Reduced sensitivity to ketamine and pentobarbital in mice lacking the N-methyl-D-aspartate receptor GluRepsilon1 subunit. Anesth. Analg. 2004;99(4):1136–1140. doi: 10.1213/01.ANE.0000131729.54986.30. [DOI] [PubMed] [Google Scholar]
- 320.Petrenko A.B., Yamakura T., Kohno T., Sakimura K., Baba H. Reduced immobilizing properties of isoflurane and nitrous oxide in mutant mice lacking the N-methyl-D-aspartate receptor GluR(epsilon)1 subunit are caused by the secondary effects of gene knockout. Anesth. Analg. 2010;110(2):461–465. doi: 10.1213/ANE.0b013e3181c76e73. [DOI] [PubMed] [Google Scholar]
- 321.Wang J.Q., Liu X., Zhang G., Parelkar N.K., Arora A., Haines M., Fibuch E.E., Mao L. Phosphorylation of glutamate receptors: a potential mechanism for the regulation of receptor function and psychostimulant action. J. Neurosci. Res. 2006;84(8):1621–1629. doi: 10.1002/jnr.21050. [DOI] [PubMed] [Google Scholar]
- 322.Tingley W.G., Ehlers M.D., Kameyama K., Doherty C., Ptak J.B., Riley C.T., Huganir R.L. Characterization of protein kinase A and protein kinase C phosphorylation of the N-methyl-D-aspartate receptor NR1 subunit using phosphorylation site-specific antibodies. J. Biol. Chem. 1997;272(8):5157–5166. doi: 10.1074/jbc.272.8.5157. [DOI] [PubMed] [Google Scholar]
- 323.Dudman J.T., Eaton M.E., Rajadhyaksha A., Macías W., Taher M., Barczak A., Kameyama K., Huganir R., Konradi C. Dopamine D1 receptors mediate CREB phosphorylation via phosphorylation of the NMDA receptor at Ser897-NR1. J. Neurochem. 2003;87(4):922–934. doi: 10.1046/j.1471-4159.2003.02067.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 324.Kingston S., Mao L., Yang L., Arora A., Fibuch E.E., Wang J.Q. Propofol inhibits phosphorylation of N-methyl-D-aspartate receptor NR1 subunits in neurons. Anesthesiology. 2006;104(4):763–769. doi: 10.1097/00000542-200604000-00021. [DOI] [PubMed] [Google Scholar]
- 325.Kozinn J., Mao L., Arora A., Yang L., Fibuch E.E., Wang J.Q. Inhibition of glutamatergic activation of extracellular signal-regulated protein kinases in hippocampal neurons by the intravenous anesthetic propofol. Anesthesiology. 2006;105(6):1182–1191. doi: 10.1097/00000542-200612000-00018. [DOI] [PubMed] [Google Scholar]
- 326.Haines M., Mao L.M., Yang L., Arora A., Fibuch E.E., Wang J.Q. Modulation of AMPA receptor GluR1 subunit phosphorylation in neurons by the intravenous anaesthetic propofol. Br. J. Anaesth. 2008;100(5):676–682. doi: 10.1093/bja/aen051. [DOI] [PubMed] [Google Scholar]
- 327.Shi S.H., Hayashi Y., Petralia R.S., Zaman S.H., Wenthold R.J., Svoboda K., Malinow R. Rapid spine delivery and redistribution of AMPA receptors after synaptic NMDA receptor activation. Science. 1999;284(5421):1811–1816. doi: 10.1126/science.284.5421.1811. [DOI] [PubMed] [Google Scholar]
- 328.Adams J.P., Sweatt J.D. Molecular psychology: roles for the ERK MAP kinase cascade in memory. Annu. Rev. Pharmacol. Toxicol. 2002;42:135–163. doi: 10.1146/annurev.pharmtox.42.082701.145401. [DOI] [PubMed] [Google Scholar]
- 329.Snyder G.L., Galdi S., Hendrick J.P., Hemmings H.C., Jr General anesthetics selectively modulate glutamatergic and dopaminergic signaling via site-specific phosphorylation in vivo. Neuropharmacology. 2007;53(5):619–630. doi: 10.1016/j.neuropharm.2007.07.008. [DOI] [PubMed] [Google Scholar]
- 330.Hang L., Shao D., Yang Y., Sun W., Dai T., Zeng Y. Alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptors participate in the analgesic but not hypnotic effects of emulsified halogenated anaesthetics. Basic Clin. Pharmacol. Toxicol. 2008;103(1):31–35. doi: 10.1111/j.1742-7843.2008.00270.x. [DOI] [PubMed] [Google Scholar]
- 331.Jevtovic-Todorovic V., Hartman R.E., Izumi Y., Benshoff N.D., Dikranian K., Zorumski C.F., Olney J.W., Wozniak D.F. Early exposure to common anesthetic agents causes widespread neurodegeneration in the developing rat brain and persistent learning deficits. J. Neurosci. 2003;23(3):876–882. doi: 10.1523/JNEUROSCI.23-03-00876.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 332.Perouansky M., Pearce R.A. How we recall (or don’t): the hippocampal memory machine and anesthetic amnesia. Can. J. Anaesth. 2011;58(2):157–166. doi: 10.1007/s12630-010-9417-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 333.Sanders R.D., Hassell J., Davidson A.J., Robertson N.J., Ma D. Impact of anaesthetics and surgery on neurodevelopment: an update. Br. J. Anaesth. 2013;110(Suppl. 1):53–72. doi: 10.1093/bja/aet054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 334.Blanpied T.A., Ehlers M.D. Microanatomy of dendritic spines: emerging principles of synaptic pathology in psychiatric and neurological disease. Biol. Psychiatry. 2004;55(12):1121–1127. doi: 10.1016/j.biopsych.2003.10.006. [DOI] [PubMed] [Google Scholar]
- 335.De Roo M., Klauser P., Briner A., Nikonenko I., Mendez P., Dayer A., Kiss J.Z., Muller D., Vutskits L. Anesthetics rapidly promote synaptogenesis during a critical period of brain development. PLoS One. 2009;4(9):e7043. doi: 10.1371/journal.pone.0007043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 336.Briner A., De Roo M., Dayer A., Muller D., Habre W., Vutskits L. Volatile anesthetics rapidly increase dendritic spine density in the rat medial prefrontal cortex during synaptogenesis. Anesthesiology. 2010;112(3):546–556. doi: 10.1097/ALN.0b013e3181cd7942. [DOI] [PubMed] [Google Scholar]
- 337.Briner A., Nikonenko I., De Roo M., Dayer A., Muller D., Vutskits L. Developmental Stage-dependent persistent impact of propofol anesthesia on dendritic spines in the rat medial prefrontal cortex. Anesthesiology. 2011;115(2):282–293. doi: 10.1097/ALN.0b013e318221fbbd. [DOI] [PubMed] [Google Scholar]
- 338.Qiu L., Zhu C., Bodogan T., Gómez-Galán M., Zhang Y., Zhou K., Li T., Xu G., Blomgren K., Eriksson L.I., Vutskits L., Terrando N. Acute and long-term effects of brief sevoflurane anesthesia during the early postnatal period in rats. Toxicol. Sci. 2016;149(1):121–133. doi: 10.1093/toxsci/kfv219. [DOI] [PubMed] [Google Scholar]
- 339.Hensch T.K. Critical period regulation. Annu. Rev. Neurosci. 2004;27:549–579. doi: 10.1146/annurev.neuro.27.070203.144327. [DOI] [PubMed] [Google Scholar]
- 340.Zhang Z., Zhang J., Li J., Zhang J., Chen L., Li Y., Guo G. Ketamine regulates phosphorylation of CRMP2 to mediate dendritic spine plasticity. J. Mol. Neurosci. 2020;70(3):353–364. doi: 10.1007/s12031-019-01419-4. [DOI] [PubMed] [Google Scholar]
- 341.Krucker T., Siggins G.R., Halpain S. Dynamic actin filaments are required for stable long-term potentiation (LTP) in area CA1 of the hippocampus. Proc. Natl. Acad. Sci. USA. 2000;97(12):6856–6861. doi: 10.1073/pnas.100139797. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 342.Lynch G., Rex C.S., Gall C.M. LTP consolidation: substrates, explanatory power, and functional significance. Neuropharmacology. 2007;52(1):12–23. doi: 10.1016/j.neuropharm.2006.07.027. [DOI] [PubMed] [Google Scholar]
- 343.Kasai H., Hayama T., Ishikawa M., Watanabe S., Yagishita S., Noguchi J. Learning rules and persistence of dendritic spines. Eur. J. Neurosci. 2010;32(2):241–249. doi: 10.1111/j.1460-9568.2010.07344.x. [DOI] [PubMed] [Google Scholar]
- 344.Head B.P., Patel H.H., Niesman I.R., Drummond J.C., Roth D.M., Patel P.M. Inhibition of p75 neurotrophin receptor attenuates isoflurane-mediated neuronal apoptosis in the neonatal central nervous system. Anesthesiology. 2009;110(4):813–825. doi: 10.1097/ALN.0b013e31819b602b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 345.Lemkuil B.P., Head B.P., Pearn M.L., Patel H.H., Drummond J.C., Patel P.M. Isoflurane neurotoxicity is mediated by p75NTR-RhoA activation and actin depolymerization. Anesthesiology. 2011;114(1):49–57. doi: 10.1097/ALN.0b013e318201dcb3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 346.Zimering J.H., Dong Y., Fang F., Huang L., Zhang Y., Xie Z. Anesthetic Sevoflurane Causes Rho-dependent filopodial shortening in mouse neurons. PLoS One. 2016;11(7):e0159637. doi: 10.1371/journal.pone.0159637. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 347.Jiang S., Hao Z., Li X., Bo L., Zhang R., Wang Y., Duan X., Kang R., Huang L. Ketamine destabilizes growth of dendritic spines in developing hippocampal neurons in vitro via a Rho dependent mechanism. Mol. Med. Rep. 2018;18(6):5037–5043. doi: 10.3892/mmr.2018.9531. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 348.Poo M.M. Neurotrophins as synaptic modulators. Nat. Rev. Neurosci. 2001;2(1):24–32. doi: 10.1038/35049004. [DOI] [PubMed] [Google Scholar]
- 349.Chapleau C.A., Larimore J.L., Theibert A., Pozzo-Miller L. Modulation of dendritic spine development and plasticity by BDNF and vesicular trafficking: fundamental roles in neurodevelopmental disorders associated with mental retardation and autism. J. Neurodev. Disord. 2009;1(3):185–196. doi: 10.1007/s11689-009-9027-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 350.Teng H.K., Teng K.K., Lee R., Wright S., Tevar S., Almeida R.D., Kermani P., Torkin R., Chen Z.Y., Lee F.S., Kraemer R.T., Nykjaer A., Hempstead B.L. ProBDNF induces neuronal apoptosis via activation of a receptor complex of p75NTR and sortilin. J. Neurosci. 2005;25(22):5455–5463. doi: 10.1523/JNEUROSCI.5123-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 351.Cowansage K.K., LeDoux J.E., Monfils M.H. Brain-derived neurotrophic factor: a dynamic gatekeeper of neural plasticity. Curr. Mol. Pharmacol. 2010;3(1):12–29. doi: 10.2174/1874467211003010012. [DOI] [PubMed] [Google Scholar]
- 352.Woo N.H., Teng H.K., Siao C.J., Chiaruttini C., Pang P.T., Milner T.A., Hempstead B.L., Lu B. Activation of p75NTR by proBDNF facilitates hippocampal long-term depression. Nat. Neurosci. 2005;8(8):1069–1077. doi: 10.1038/nn1510. [DOI] [PubMed] [Google Scholar]
- 353.Yang J., Siao C.J., Nagappan G., Marinic T., Jing D., McGrath K., Chen Z.Y., Mark W., Tessarollo L., Lee F.S., Lu B., Hempstead B.L. Neuronal release of proBDNF. Nat. Neurosci. 2009;12(2):113–115. doi: 10.1038/nn.2244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 354.Pearn M.L., Hu Y., Niesman I.R., Patel H.H., Drummond J.C., Roth D.M., Akassoglou K., Patel P.M., Head B.P. Propofol neurotoxicity is mediated by p75 neurotrophin receptor activation. Anesthesiology. 2012;116(2):352–361. doi: 10.1097/ALN.0b013e318242a48c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 355.Lu L.X., Yon J.H., Carter L.B., Jevtovic-Todorovic V. General anesthesia activates BDNF-dependent neuroapoptosis in the developing rat brain. Apoptosis. 2006;11(9):1603–1615. doi: 10.1007/s10495-006-8762-3. [DOI] [PubMed] [Google Scholar]
- 356.Soppet D., Escandon E., Maragos J., Middlemas D.S., Reid S.W., Blair J., Burton L.E., Stanton B.R., Kaplan D.R., Hunter T., Nikolics K., Parada L.F. The neurotrophic factors brain-derived neurotrophic factor and neurotrophin-3 are ligands for the trkB tyrosine kinase receptor. Cell. 1991;65(5):895–903. doi: 10.1016/0092-8674(91)90396-G. [DOI] [PubMed] [Google Scholar]
- 357.Mizuno M., Yamada K., He J., Nakajima A., Nabeshima T. Involvement of BDNF receptor TrkB in spatial memory formation. Learn. Mem. 2003;10(2):108–115. doi: 10.1101/lm.56003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 358.Yang T. A small molecule TrkB/TrkC neurotrophin receptor coactivator with distinctive effects on neuronal survival and process outgrowth. Neuropharmacology, 2016;110(Pt A):343–361. doi: 10.1016/j.neuropharm.2016.06.015. [DOI] [PubMed] [Google Scholar]
- 359.Vutskits L., Lysakowski C., Czarnetzki C., Jenny B., Copin J.C., Tramèr M.R. Plasma concentrations of brain-derived neurotrophic factor in patients undergoing minor surgery: a randomized controlled trial. Neurochem. Res. 2008;33(7):1325–1331. doi: 10.1007/s11064-007-9586-4. [DOI] [PubMed] [Google Scholar]
- 360.Ji M., Dong L., Jia M., Liu W., Zhang M., Ju L., Yang J., Xie Z., Yang J. Epigenetic enhancement of brain-derived neurotrophic factor signaling pathway improves cognitive impairments induced by isoflurane exposure in aged rats. Mol. Neurobiol. 2014;50(3):937–944. doi: 10.1007/s12035-014-8659-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 361.Citri A., Malenka R.C. Synaptic plasticity: multiple forms, functions, and mechanisms. Neuropsychopharmacology. 2008;33(1):18–41. doi: 10.1038/sj.npp.1301559. [DOI] [PubMed] [Google Scholar]
- 362.Simon W., Hapfelmeier G., Kochs E., Zieglgänsberger W., Rammes G. Isoflurane blocks synaptic plasticity in the mouse hippocampus. Anesthesiology. 2001;94(6):1058–1065. doi: 10.1097/00000542-200106000-00021. [DOI] [PubMed] [Google Scholar]
- 363.Chen B., Deng X., Wang B., Liu H. Persistent neuronal apoptosis and synaptic loss induced by multiple but not single exposure of propofol contribute to long-term cognitive dysfunction in neonatal rats. J. Toxicol. Sci. 2016;41(5):627–636. doi: 10.2131/jts.41.627. [DOI] [PubMed] [Google Scholar]
- 364.Perouansky M., Rau V., Ford T., Oh S.I., Perkins M., Eger E.I., II, Pearce R.A. Slowing of the hippocampal θ rhythm correlates with anesthetic-induced amnesia. Anesthesiology. 2010;113(6):1299–1309. doi: 10.1097/ALN.0b013e3181f90ccc. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 365.Peng S., Zhang Y., Li G.J., Zhang D.X., Sun D.P., Fang Q. The effect of sevoflurane on the expression of M1 acetylcholine receptor in the hippocampus and cognitive function of aged rats. Mol. Cell. Biochem. 2012;361(1-2):229–233. doi: 10.1007/s11010-011-1107-8. [DOI] [PubMed] [Google Scholar]
- 366.Wei H., Xiong W., Yang S., Zhou Q., Liang C., Zeng B.X., Xu L. Propofol facilitates the development of long-term depression (LTD) and impairs the maintenance of long-term potentiation (LTP) in the CA1 region of the hippocampus of anesthetized rats. Neurosci. Lett. 2002;324(3):181–184. doi: 10.1016/S0304-3940(02)00183-0. [DOI] [PubMed] [Google Scholar]
- 367.Lin D., Zuo Z. Isoflurane induces hippocampal cell injury and cognitive impairments in adult rats. Neuropharmacology. 2011;61(8):1354–1359. doi: 10.1016/j.neuropharm.2011.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 368.Yu D., Jiang Y., Gao J., Liu B., Chen P. Repeated exposure to propofol potentiates neuroapoptosis and long-term behavioral deficits in neonatal rats. Neurosci. Lett. 2013;534:41–46. doi: 10.1016/j.neulet.2012.12.033. [DOI] [PubMed] [Google Scholar]
- 369.Zhu C., Gao J., Karlsson N., Li Q., Zhang Y., Huang Z., Li H., Kuhn H.G., Blomgren K. Isoflurane anesthesia induced persistent, progressive memory impairment, caused a loss of neural stem cells, and reduced neurogenesis in young, but not adult, rodents. J. Cereb. Blood Flow Metab. 2010;30(5):1017–1030. doi: 10.1038/jcbfm.2009.274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 370.Peng S., Zhang Y., Sun D.P., Zhang D.X., Fang Q., Li G.J. The effect of sevoflurane anesthesia on cognitive function and the expression of Insulin-like Growth Factor-1 in CA1 region of hippocampus in old rats. Mol. Biol. Rep. 2011;38(2):1195–1199. doi: 10.1007/s11033-010-0217-9. [DOI] [PubMed] [Google Scholar]
- 371.MacIver M.B., Tauck D.L., Kendig J.J. General anaesthetic modification of synaptic facilitation and long-term potentiation in hippocampus. Br. J. Anaesth. 1989;62(3):301–310. doi: 10.1093/bja/62.3.301. [DOI] [PubMed] [Google Scholar]
- 372.Haseneder R., Kratzer S., von Meyer L., Eder M., Kochs E., Rammes G. Isoflurane and sevoflurane dose-dependently impair hippocampal long-term potentiation. Eur. J. Pharmacol. 2009;623(1-3):47–51. doi: 10.1016/j.ejphar.2009.09.022. [DOI] [PubMed] [Google Scholar]
- 373.Guo D., Gan J., Tan T., Tian X., Wang G., Ng K.T. Neonatal exposure of ketamine inhibited the induction of hippocampal long-term potentiation without impairing the spatial memory of adult rats. Cogn. Neurodynamics. 2018;12(4):377–383. doi: 10.1007/s11571-018-9474-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 374.Stringer J.L., Guyenet P.G. Elimination of long-term potentiation in the hippocampus by phencyclidine and ketamine. Brain Res. 1983;258(1):159–164. doi: 10.1016/0006-8993(83)91244-1. [DOI] [PubMed] [Google Scholar]
- 375.MacDonald J.F., Miljkovic Z., Pennefather P. Use-dependent block of excitatory amino acid currents in cultured neurons by ketamine. J. Neurophysiol. 1987;58(2):251–266. doi: 10.1152/jn.1987.58.2.251. [DOI] [PubMed] [Google Scholar]
- 376.Wang R.R., Jin J.H., Womack A.W., Lyu D., Kokane S.S., Tang N., Zou X., Lin Q., Chen J. Neonatal ketamine exposure causes impairment of long-term synaptic plasticity in the anterior cingulate cortex of rats. Neuroscience. 2014;268:309–317. doi: 10.1016/j.neuroscience.2014.03.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 377.Matsuura T., Kamiya Y., Itoh H., Higashi T., Yamada Y., Andoh T. Inhibitory effects of isoflurane and nonimmobilizing halogenated compounds on neuronal nicotinic acetylcholine receptors. Anesthesiology. 2002;97(6):1541–1549. doi: 10.1097/00000542-200212000-00029. [DOI] [PubMed] [Google Scholar]
- 378.Rada E.M., Tharakan E.C., Flood P. Volatile anesthetics reduce agonist affinity at nicotinic acetylcholine receptors in the brain. Anesth. Analg. 2003;96(1):108–111. doi: 10.1097/00000539-200301000-00023. [DOI] [PubMed] [Google Scholar]
- 379.Piao M.H., Liu Y., Wang Y.S., Qiu J.P., Feng C.S. Volatile anesthetic isoflurane inhibits LTP induction of hippocampal CA1 neurons through α4β2 nAChR subtype-mediated mechanisms. Ann. Fr. Anesth. Reanim. 2013;32(10):e135–e141. doi: 10.1016/j.annfar.2013.05.012. [DOI] [PubMed] [Google Scholar]
- 380.Mawhinney L.J., de Rivero Vaccari J.P., Alonso O.F., Jimenez C.A., Furones C., Moreno W.J., Lewis M.C., Dietrich W.D., Bramlett H.M. Isoflurane/nitrous oxide anesthesia induces increases in NMDA receptor subunit NR2B protein expression in the aged rat brain. Brain Res. 2012;1431:23–34. doi: 10.1016/j.brainres.2011.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 381.Uchimoto K., Miyazaki T., Kamiya Y., Mihara T., Koyama Y., Taguri M., Inagawa G., Takahashi T., Goto T. Isoflurane impairs learning and hippocampal long-term potentiation via the saturation of synaptic plasticity. Anesthesiology. 2014;121(2):302–310. doi: 10.1097/ALN.0000000000000269. [DOI] [PubMed] [Google Scholar]
- 382.Yamakura T., Bertaccini E., Trudell J.R., Harris R.A. Anesthetics and ion channels: molecular models and sites of action. Annu. Rev. Pharmacol. Toxicol. 2001;41:23–51. doi: 10.1146/annurev.pharmtox.41.1.23. [DOI] [PubMed] [Google Scholar]
- 383.Kato R., Tachibana K., Nishimoto N., Hashimoto T., Uchida Y., Ito R., Tsuruga K., Takita K., Morimoto Y. Neonatal exposure to sevoflurane causes significant suppression of hippocampal long-term potentiation in postgrowth rats. Anesth. Analg. 2013;117(6):1429–1435. doi: 10.1213/ANE.0b013e3182a8c709. [DOI] [PubMed] [Google Scholar]
- 384.Nagashima K., Zorumski C.F., Izumi Y. Propofol inhibits long-term potentiation but not long-term depression in rat hippocampal slices. Anesthesiology. 2005;103(2):318–326. doi: 10.1097/00000542-200508000-00015. [DOI] [PubMed] [Google Scholar]
- 385.Wang W., Wang H., Gong N., Xu T.L. Changes of K+ -Cl- cotransporter 2 (KCC2) and circuit activity in propofol-induced impairment of long-term potentiation in rat hippocampal slices. Brain Res. Bull. 2006;70(4-6):444–449. doi: 10.1016/j.brainresbull.2006.07.004. [DOI] [PubMed] [Google Scholar]
- 386.Takamatsu I., Sekiguchi M., Wada K., Sato T., Ozaki M. Propofol-mediated impairment of CA1 long-term potentiation in mouse hippocampal slices. Neurosci. Lett. 2005;389(3):129–132. doi: 10.1016/j.neulet.2005.07.043. [DOI] [PubMed] [Google Scholar]
- 387.Zarnowska E.D., Rodgers F.C., Oh I., Rau V., Lor C., Laha K.T., Jurd R., Rudolph U., Eger E.I.N., Pearce R.A. Etomidate blocks LTP and impairs learning but does not enhance tonic inhibition in mice carrying the N265M point mutation in the beta3 subunit of the GABA(A) receptor. Neuropharmacology. 2015;93:171–178. doi: 10.1016/j.neuropharm.2015.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 388.Ikonomidou C., Bosch F., Miksa M., Bittigau P., Vöckler J., Dikranian K., Tenkova T.I., Stefovska V., Turski L., Olney J.W. Blockade of NMDA receptors and apoptotic neurodegeneration in the developing brain. Science. 1999;283(5398):70–74. doi: 10.1126/science.283.5398.70. [DOI] [PubMed] [Google Scholar]
- 389.Fredriksson A., Pontén E., Gordh T., Eriksson P. Neonatal exposure to a combination of N-methyl-D-aspartate and gamma-aminobutyric acid type A receptor anesthetic agents potentiates apoptotic neurodegeneration and persistent behavioral deficits. Anesthesiology. 2007;107(3):427–436. doi: 10.1097/01.anes.0000278892.62305.9c. [DOI] [PubMed] [Google Scholar]
- 390.Slikker W., Jr, Zou X., Hotchkiss C.E., Divine R.L., Sadovova N., Twaddle N.C., Doerge D.R., Scallet A.C., Patterson T.A., Hanig J.P., Paule M.G., Wang C. Ketamine-induced neuronal cell death in the perinatal rhesus monkey. Toxicol. Sci. 2007;98(1):145–158. doi: 10.1093/toxsci/kfm084. [DOI] [PubMed] [Google Scholar]
- 391.Satomoto M., Satoh Y., Terui K., Miyao H., Takishima K., Ito M., Imaki J. Neonatal exposure to sevoflurane induces abnormal social behaviors and deficits in fear conditioning in mice. Anesthesiology. 2009;110(3):628–637. doi: 10.1097/ALN.0b013e3181974fa2. [DOI] [PubMed] [Google Scholar]
- 392.Zou X., Patterson T.A., Divine R.L., Sadovova N., Zhang X., Hanig J.P., Paule M.G., Slikker W., Jr, Wang C. Prolonged exposure to ketamine increases neurodegeneration in the developing monkey brain. Int. J. Dev. Neurosci. 2009;27(7):727–731. doi: 10.1016/j.ijdevneu.2009.06.010. [DOI] [PubMed] [Google Scholar]
- 393.Brambrink A.M., Evers A.S., Avidan M.S., Farber N.B., Smith D.J., Zhang X., Dissen G.A., Creeley C.E., Olney J.W. Isoflurane-induced neuroapoptosis in the neonatal rhesus macaque brain. Anesthesiology. 2010;112(4):834–841. doi: 10.1097/ALN.0b013e3181d049cd. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 394.Zou X., Liu F., Zhang X., Patterson T.A., Callicott R., Liu S., Hanig J.P., Paule M.G., Slikker W., Jr, Wang C. Inhalation anesthetic-induced neuronal damage in the developing rhesus monkey. Neurotoxicol. Teratol. 2011;33(5):592–597. doi: 10.1016/j.ntt.2011.06.003. [DOI] [PubMed] [Google Scholar]
- 395.Brambrink A.M., Evers A.S., Avidan M.S., Farber N.B., Smith D.J., Martin L.D., Dissen G.A., Creeley C.E., Olney J.W. Ketamine-induced neuroapoptosis in the fetal and neonatal rhesus macaque brain. Anesthesiology. 2012;116(2):372–384. doi: 10.1097/ALN.0b013e318242b2cd. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 396.Creeley C., Dikranian K., Dissen G., Martin L., Olney J., Brambrink A. Propofol-induced apoptosis of neurones and oligodendrocytes in fetal and neonatal rhesus macaque brain. Br. J. Anaesth. 2013;(1 Suppl. 1):29–38. doi: 10.1093/bja/aet173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 397.Noguchi K.K., Johnson S.A., Dissen G.A., Martin L.D., Manzella F.M., Schenning K.J., Olney J.W., Brambrink A.M. Isoflurane exposure for three hours triggers apoptotic cell death in neonatal macaque brain. Br. J. Anaesth. 2017;119(3):524–531. doi: 10.1093/bja/aex123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 398.Paule M.G., Li M., Allen R.R., Liu F., Zou X., Hotchkiss C., Hanig J.P., Patterson T.A., Slikker W., Jr, Wang C. Ketamine anesthesia during the first week of life can cause long-lasting cognitive deficits in rhesus monkeys. Neurotoxicol. Teratol. 2011;33(2):220–230. doi: 10.1016/j.ntt.2011.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 399.Neudecker V., Perez-Zoghbi J.F., Coleman K., Neuringer M., Robertson N., Bemis A., Glickman B., Schenning K.J., Fair D.A., Martin L.D., Dissen G.A., Brambrink A.M. Infant isoflurane exposure affects social behaviours, but does not impair specific cognitive domains in juvenile non-human primates. Br. J. Anaesth. 2021;126(2):486–499. doi: 10.1016/j.bja.2020.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 400.Coleman K., Robertson N.D., Dissen G.A., Neuringer M.D., Martin L.D., Cuzon Carlson V.C., Kroenke C., Fair D., Brambrink A.M. Isoflurane anesthesia has long-term consequences on motor and behavioral development in infant rhesus macaques. Anesthesiology. 2017;126(1):74–84. doi: 10.1097/ALN.0000000000001383. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 401.Xiao H., Liu B., Chen Y., Zhang J. Learning, memory and synaptic plasticity in hippocampus in rats exposed to sevoflurane. Int. J. Dev. Neurosci. 2016;48:38–49. doi: 10.1016/j.ijdevneu.2015.11.001. [DOI] [PubMed] [Google Scholar]
- 402.Guo S., Liu L., Wang C., Jiang Q., Dong Y., Tian Y. Repeated exposure to sevoflurane impairs the learning and memory of older male rats. Life Sci. 2018;192:75–83. doi: 10.1016/j.lfs.2017.11.025. [DOI] [PubMed] [Google Scholar]
- 403.Huang L., Yang G. Repeated exposure to ketamine-xylazine during early development impairs motor learning-dependent dendritic spine plasticity in adulthood. Anesthesiology. 2015;122(4):821–831. doi: 10.1097/ALN.0000000000000579. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 404.Liu J., Zhao Y., Yang J., Zhang X., Zhang W., Wang P. Neonatal repeated exposure to isoflurane not sevoflurane in mice reversibly impaired spatial cognition at juvenile-age. Neurochem. Res. 2017;42(2):595–605. doi: 10.1007/s11064-016-2114-7. [DOI] [PubMed] [Google Scholar]
- 405.Lee B.H., Chan J.T., Kraeva E., Peterson K., Sall J.W. Isoflurane exposure in newborn rats induces long-term cognitive dysfunction in males but not females. Neuropharmacology. 2014;83:9–17. doi: 10.1016/j.neuropharm.2014.03.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 406.Sasaki R.J.M., Hagelstein M., Lee B.H., Sall J.W. Anesthesia-induced recognition deficit is improved in postnatally gonadectomized male rats. J. Neurosurg. Anesthesiol. 2021;33(3):273–280. doi: 10.1097/ANA.0000000000000641. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 407.Yi X., Cai Y., Li W. Isoflurane damages the developing brain of mice and induces subsequent learning and memory deficits through FASL-FAS Signaling. BioMed Res. Int. 2015;2015:315872. doi: 10.1155/2015/315872. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 408.Li X., Wei K., Hu R., Zhang B., Li L., Wan L., Zhang C., Yao W. Upregulation of Cdh1 attenuates isoflurane-induced neuronal apoptosis and long-term cognitive impairments in developing rats. Front. Cell. Neurosci. 2017;11:368. doi: 10.3389/fncel.2017.00368. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 409.Peng S., Zhang Y., Zhang J., Wang H., Ren B. Effect of ketamine on ERK expression in hippocampal neural cell and the ability of learning behavior in minor rats. Mol. Biol. Rep. 2010;37(7):3137–3142. doi: 10.1007/s11033-009-9892-9. [DOI] [PubMed] [Google Scholar]
- 410.Huang L., Liu Y., Jin W., Ji X., Dong Z. Ketamine potentiates hippocampal neurodegeneration and persistent learning and memory impairment through the PKCγ-ERK signaling pathway in the developing brain. Brain Res. 2012;1476:164–171. doi: 10.1016/j.brainres.2012.07.059. [DOI] [PubMed] [Google Scholar]
- 411.Yu X., Liu Y., Bo S., Qinghua L. Effects of sevoflurane on learning, memory, and expression of pERK1/2 in hippocampus in neonatal rats. Acta Anaesthesiol. Scand. 2015;59(1):78–84. doi: 10.1111/aas.12433. [DOI] [PubMed] [Google Scholar]
- 412.Liang L., Xie R., Lu R., Ma R., Wang X., Wang F., Liu B., Wu S., Wang Y., Zhang H. Involvement of homodomain interacting protein kinase 2-c-Jun N-terminal kinase/c-Jun cascade in the long-term synaptic toxicity and cognition impairment induced by neonatal Sevoflurane exposure. J. Neurochem. 2020;154(4):372–388. doi: 10.1111/jnc.14910. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 413.Wang S.Q., Fang F., Xue Z.G., Cang J., Zhang X.G. Neonatal sevoflurane anesthesia induces long-term memory impairment and decreases hippocampal PSD-95 expression without neuronal loss. Eur. Rev. Med. Pharmacol. Sci. 2013;17(7):941–950. [PubMed] [Google Scholar]
- 414.Ling Y.Z., Ma W., Yu L., Zhang Y., Liang Q.S. Decreased PSD95 expression in medial prefrontal cortex (mPFC) was associated with cognitive impairment induced by sevoflurane anesthesia. J. Zhejiang Univ. Sci. B. 2015;16(9):763–771. doi: 10.1631/jzus.B1500006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 415.Schaefer M.L., Perez P.J., Wang M., Gray C., Krall C., Sun X., Hunter E., Skinner J., Johns R.A. Neonatal isoflurane anesthesia or disruption of postsynaptic density-95 protein interactions change dendritic spine densities and cognitive function in juvenile mice. Anesthesiology. 2020;133(4):812–823. doi: 10.1097/ALN.0000000000003482. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 416.Wiklund A., Granon S., Faure P., Sundman E., Changeux J.P., Eriksson L.I. Object memory in young and aged mice after sevoflurane anaesthesia. Neuroreport. 2009;20(16):1419–1423. doi: 10.1097/WNR.0b013e328330cd2b. [DOI] [PubMed] [Google Scholar]
- 417.Su D., Zhao Y., Wang B., Xu H., Li W., Chen J., Wang X. Isoflurane-induced spatial memory impairment in mice is prevented by the acetylcholinesterase inhibitor donepezil. PLoS One. 2011;6(11):e27632. doi: 10.1371/journal.pone.0027632. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 418.Wang H., Xu Z., Feng C., Wang Y., Jia X., Wu A., Yue Y. Changes of learning and memory in aged rats after isoflurane inhalational anaesthesia correlated with hippocampal acetylcholine level. Ann. Fr. Anesth. Reanim. 2012;31(3):e61–e66. doi: 10.1016/j.annfar.2011.02.005. [DOI] [PubMed] [Google Scholar]
- 419.Xiong L., Duan L., Xu W., Wang Z. Nerve growth factor metabolic dysfunction contributes to sevoflurane-induced cholinergic degeneration and cognitive impairments. Brain Res. 2019;1707:107–116. doi: 10.1016/j.brainres.2018.11.033. [DOI] [PubMed] [Google Scholar]
- 420.Kong F.J., Ma L.L., Zhang H.H., Zhou J.Q. Alpha 7 nicotinic acetylcholine receptor agonist GTS-21 mitigates isoflurane-induced cognitive impairment in aged rats. J. Surg. Res. 2015;194(1):255–261. doi: 10.1016/j.jss.2014.09.043. [DOI] [PubMed] [Google Scholar]
- 421.Tang X., Li Y., Ao J., Ding L., Liu Y., Yuan Y., Wang Z., Wang G. Role of α7nAChR-NMDAR in sevoflurane-induced memory deficits in the developing rat hippocampus. PLoS One. 2018;13(2):e0192498. doi: 10.1371/journal.pone.0192498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 422.Li Z., Ni C., Xia C., Jaw J., Wang Y., Cao Y., Xu M., Guo X. Calcineurin/nuclear factor-κB signaling mediates isoflurane-induced hippocampal neuroinflammation and subsequent cognitive impairment in aged rats. Mol. Med. Rep. 2017;15(1):201–209. doi: 10.3892/mmr.2016.5967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 423.Stratmann G., Sall J.W., May L.D., Bell J.S., Magnusson K.R., Rau V., Visrodia K.H., Alvi R.S., Ku B., Lee M.T., Dai R. Isoflurane differentially affects neurogenesis and long-term neurocognitive function in 60-day-old and 7-day-old rats. Anesthesiology. 2009;110(4):834–848. doi: 10.1097/ALN.0b013e31819c463d. [DOI] [PubMed] [Google Scholar]
- 424.Stratmann G., Sall J.W., Bell J.S., Alvi R.S., May Ld., Ku B., Dowlatshahi M., Dai R., Bickler P.E., Russell I., Lee M.T., Hrubos M.W., Chiu C. Isoflurane does not affect brain cell death, hippocampal neurogenesis, or long-term neurocognitive outcome in aged rats. Anesthesiology. 2010;112(2):305–315. doi: 10.1097/ALN.0b013e3181ca33a1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 425.Callaway J.K., Jones N.C., Royse A.G., Royse C.F. Memory impairment in rats after desflurane anesthesia is age and dose dependent. J. Alzheimers Dis. 2015;44(3):995–1005. doi: 10.3233/JAD-132444. [DOI] [PubMed] [Google Scholar]
- 426.Huang H., Liu C.M., Sun J., Jin W.J., Wu Y.Q., Chen J. Repeated 2% sevoflurane administration in 7 and 60-day-old rats: Neurotoxicity and neurocognitive dysfunction. Anaesthesist. 2017;66(11):850–857. doi: 10.1007/s00101-017-0359-4. [DOI] [PubMed] [Google Scholar]
- 427.Liang X., Zhang Y., Zhang C., Tang C., Wang Y., Ren J., Chen X., Zhang Y., Zhu Z. Effect of repeated neonatal sevoflurane exposure on the learning, memory and synaptic plasticity at juvenile and adult age. Am. J. Transl. Res. 2017;9(11):4974–4983. [PMC free article] [PubMed] [Google Scholar]
- 428.Callaway J.K., Jones N.C., Royse C.F. Isoflurane induces cognitive deficits in the Morris water maze task in rats. Eur. J. Anaesthesiol. 2012;29(5):239–245. doi: 10.1097/EJA.0b013e32835103c1. [DOI] [PubMed] [Google Scholar]
- 429.Martin L.J., Oh G.H., Orser B.A. Etomidate targets alpha5 gamma-aminobutyric acid subtype A receptors to regulate synaptic plasticity and memory blockade. Anesthesiology. 2009;111(5):1025–1035. doi: 10.1097/ALN.0b013e3181bbc961. [DOI] [PubMed] [Google Scholar]
- 430.Zurek A.A., Bridgwater E.M., Orser B.A. Inhibition of α5 γ-Aminobutyric acid type A receptors restores recognition memory after general anesthesia. Anesth. Analg. 2012;114(4):845–855. doi: 10.1213/ANE.0b013e31824720da. [DOI] [PubMed] [Google Scholar]
- 431.Zurek A.A., Yu J., Wang D.S., Haffey S.C., Bridgwater E.M., Penna A., Lecker I., Lei G., Chang T., Salter E.W., Orser B.A. Sustained increase in α5GABAA receptor function impairs memory after anesthesia. J. Clin. Invest. 2014;124(12):5437–5441. doi: 10.1172/JCI76669. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 432.Landin J.D., Palac M., Carter J.M., Dzumaga Y., Santerre-Anderson J.L., Fernandez G.M., Savage L.M., Varlinskaya E.I., Spear L.P., Moore S.D., Swartzwelder H.S., Fleming R.L., Werner D.F. General anesthetic exposure in adolescent rats causes persistent maladaptations in cognitive and affective behaviors and neuroplasticity. Neuropharmacology. 2019;150:153–163. doi: 10.1016/j.neuropharm.2019.03.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 433.Wu J., Bie B., Naguib M. Epigenetic manipulation of brain-derived neurotrophic factor improves memory deficiency induced by neonatal anesthesia in rats. Anesthesiology. 2016;124(3):624–640. doi: 10.1097/ALN.0000000000000981. [DOI] [PubMed] [Google Scholar]
- 434.Zhang F., Zhu Z.Q., Liu D.X., Zhang C., Gong Q.H., Zhu Y.H. Emulsified isoflurane anesthesia decreases brain-derived neurotrophic factor expression and induces cognitive dysfunction in adult rats. Exp. Ther. Med. 2014;8(2):471–477. doi: 10.3892/etm.2014.1769. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 435.Xu Z., Qian B. Sevoflurane anesthesia-mediated oxidative stress and cognitive impairment in hippocampal neurons of old rats can be ameliorated by expression of brain derived neurotrophic factor. Neurosci. Lett. 2020;721:134785. doi: 10.1016/j.neulet.2020.134785. [DOI] [PubMed] [Google Scholar]
- 436.Goulart B.K., de Lima M.N., de Farias C.B., Reolon G.K., Almeida V.R., Quevedo J., Kapczinski F., Schröder N., Roesler R. Ketamine impairs recognition memory consolidation and prevents learning-induced increase in hippocampal brain-derived neurotrophic factor levels. Neuroscience. 2010;167(4):969–973. doi: 10.1016/j.neuroscience.2010.03.032. [DOI] [PubMed] [Google Scholar]
- 437.Zhang G., Dong Y., Zhang B., Ichinose F., Wu X., Culley D.J., Crosby G., Tanzi R.E., Xie Z. Isoflurane-induced caspase-3 activation is dependent on cytosolic calcium and can be attenuated by memantine. J. Neurosci. 2008;28(17):4551–4560. doi: 10.1523/JNEUROSCI.5694-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 438.Zhao Y., Liang G., Chen Q., Joseph D.J., Meng Q., Eckenhoff R.G., Eckenhoff M.F., Wei H. Anesthetic-induced neurodegeneration mediated via inositol 1,4,5-trisphosphate receptors. J. Pharmacol. Exp. Ther. 2010;333(1):14–22. doi: 10.1124/jpet.109.161562. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 439.Wang H., Dong Y., Zhang J., Xu Z., Wang G., Swain C.A., Zhang Y., Xie Z. Isoflurane induces endoplasmic reticulum stress and caspase activation through ryanodine receptors. Br. J. Anaesth. 2014;113(4):695–707. doi: 10.1093/bja/aeu053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 440.Eckenhoff R.G., Johansson J.S., Wei H., Carnini A., Kang B., Wei W., Pidikiti R., Keller J.M., Eckenhoff M.F. Inhaled anesthetic enhancement of amyloid-beta oligomerization and cytotoxicity. Anesthesiology. 2004;101(3):703–709. doi: 10.1097/00000542-200409000-00019. [DOI] [PubMed] [Google Scholar]
- 441.Xie Z., Dong Y., Maeda U., Alfille P., Culley D.J., Crosby G., Tanzi R.E. The common inhalation anesthetic isoflurane induces apoptosis and increases amyloid beta protein levels. Anesthesiology. 2006;104(5):988–994. doi: 10.1097/00000542-200605000-00015. [DOI] [PubMed] [Google Scholar]
- 442.Bianchi S.L., Tran T., Liu C., Lin S., Li Y., Keller J.M., Eckenhoff R.G., Eckenhoff M.F. Brain and behavior changes in 12-month-old Tg2576 and nontransgenic mice exposed to anesthetics. Neurobiol. Aging. 2008;29(7):1002–1010. doi: 10.1016/j.neurobiolaging.2007.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 443.Zhang S., Hu X., Guan W., Luan L., Li B., Tang Q., Fan H. Isoflurane anesthesia promotes cognitive impairment by inducing expression of β-amyloid protein-related factors in the hippocampus of aged rats. PLoS One. 2017;12(4):e0175654. doi: 10.1371/journal.pone.0175654. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 444.Liu H., Weng H. Up-regulation of Alzheimer’s disease-associated proteins may cause enflurane anesthesia induced cognitive decline in aged rats. Neurol. Sci. 2014;35(2):185–189. doi: 10.1007/s10072-013-1474-x. [DOI] [PubMed] [Google Scholar]
- 445.Li C., Liu S., Xing Y., Tao F. The role of hippocampal tau protein phosphorylation in isoflurane-induced cognitive dysfunction in transgenic APP695 mice. Anesth. Analg. 2014;119(2):413–419. doi: 10.1213/ANE.0000000000000315. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 446.Tao G., Zhang J., Zhang L., Dong Y., Yu B., Crosby G., Culley D.J., Zhang Y., Xie Z. Sevoflurane induces tau phosphorylation and glycogen synthase kinase 3β activation in young mice. Anesthesiology. 2014;121(3):510–527. doi: 10.1097/ALN.0000000000000278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 447.Yu Y., Yang Y., Tan H., Boukhali M., Khatri A., Yu Y., Hua F., Liu L., Li M., Yang G., Dong Y., Zhang Y., Haas W., Xie Z. Tau Contributes to Sevoflurane-induced Neurocognitive Impairment in Neonatal Mice. Anesthesiology. 2020;133(3):595–610. doi: 10.1097/ALN.0000000000003452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 448.Le Freche H., Brouillette J., Fernandez-Gomez F.J., Patin P., Caillierez R., Zommer N., Sergeant N., Buée-Scherrer V., Lebuffe G., Blum D., Buée L. Tau phosphorylation and sevoflurane anesthesia: an association to postoperative cognitive impairment. Anesthesiology. 2012;116(4):779–787. doi: 10.1097/ALN.0b013e31824be8c7. [DOI] [PubMed] [Google Scholar]
- 449.Wang L., Zheng M., Wu S., Niu Z. MicroRNA-188-3p is involved in sevoflurane anesthesia-induced neuroapoptosis by targeting MDM2. Mol. Med. Rep. 2018;17(3):4229–4236. doi: 10.3892/mmr.2018.8437. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 450.Evered L., Silbert B., Knopman D.S., Scott D.A., DeKosky S.T., Rasmussen L.S., Oh E.S., Crosby G., Berger M., Eckenhoff R.G. Recommendations for the nomenclature of cognitive change associated with anaesthesia and surgery-2018. Br. J. Anaesth. 2018;121(5):1005–1012. doi: 10.1016/j.bja.2017.11.087. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 451.Demeure M.J., Fain M.J. The elderly surgical patient and postoperative delirium. J. Am. Coll. Surg. 2006;203(5):752–757. doi: 10.1016/j.jamcollsurg.2006.07.032. [DOI] [PubMed] [Google Scholar]
- 452.Rasmussen L.S. Postoperative cognitive dysfunction: incidence and prevention. Baillieres. Best Pract. Res. Clin. Anaesthesiol. 2006;20(2):315–330. doi: 10.1016/j.bpa.2005.10.011. [DOI] [PubMed] [Google Scholar]
- 453.Deiner S., Silverstein J.H. Postoperative delirium and cognitive dysfunction. Br. J. Anaesth. 2009;103(Suppl. 1):41–46. doi: 10.1093/bja/aep291. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 454.Morimoto Y., Yoshimura M., Utada K., Setoyama K., Matsumoto M., Sakabe T. Prediction of postoperative delirium after abdominal surgery in the elderly. J. Anesth. 2009;23(1):51–56. doi: 10.1007/s00540-008-0688-1. [DOI] [PubMed] [Google Scholar]
- 455.Schmitt E.M., Marcantonio E.R., Alsop D.C., Jones R.N., Rogers S.O., Jr, Fong T.G., Metzger E., Inouye S.K. Novel risk markers and long-term outcomes of delirium: the successful aging after elective surgery (SAGES) study design and methods. J. Am. Med. Dir. Assoc. 2012;13(9):818.e1–818.e10. doi: 10.1016/j.jamda.2012.08.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 456.Daiello L.A., Racine A.M., Yun Gou R., Marcantonio E.R., Xie Z., Kunze L.J., Vlassakov K.V., Inouye S.K., Jones R.N., Alsop D., Travison T., Arnold S., Cooper Z., Dickerson B., Fong T., Metzger E., Pascual-Leone A., Schmitt E.M., Shafi M., Cavallari M., Dai W., Dillon S.T., McElhaney J., Guttmann C., Hshieh T., Kuchel G., Libermann T., Ngo L., Press D., Saczynski J., Vasunilashorn S., O’Connor M., Kimchi E., Strauss J., Wong B., Belkin M., Ayres D., Callery M., Pomposelli F., Wright J., Schermerhorn M., Abrantes T., Albuquerque A., Bertrand S., Brown A., Callahan A., D’Aquila M., Dowal S., Fox M., Gallagher J., Anna Gersten R., Hodara A., Helfand B., Inloes J., Kettell J., Kuczmarska A., Nee J., Nemeth E., Ochsner L., Palihnich K., Parisi K., Puelle M., Rastegar S., Vella M., Xu G., Bryan M., Guess J., Enghorn D., Gross A., Gou Y., Habtemariam D., Isaza I., Kosar C., Rockett C., Tommet D., Gruen T., Ross M., Tasker K., Gee J., Kolanowski A., Pisani M., de Rooij S., Rogers S., Studenski S., Stern Y., Whittemore A., Gottlieb G., Orav J., Sperling R. Postoperative delirium and postoperative cognitive dysfunction: overlap and divergence. Anesthesiology. 2019;131(3):477–491. doi: 10.1097/ALN.0000000000002729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 457.Monk T.G., Weldon B.C., Garvan C.W., Dede D.E., van der Aa M.T., Heilman K.M., Gravenstein J.S. Predictors of cognitive dysfunction after major noncardiac surgery. Anesthesiology. 2008;108(1):18–30. doi: 10.1097/01.anes.0000296071.19434.1e. [DOI] [PubMed] [Google Scholar]
- 458.Berger M., Nadler J.W., Browndyke J., Terrando N., Ponnusamy V., Cohen H.J., Whitson H.E., Mathew J.P. Postoperative cognitive dysfunction: minding the gaps in our knowledge of a common postoperative complication in the elderly. Anesthesiol. Clin. 2015;33(3):517–550. doi: 10.1016/j.anclin.2015.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 459.Terrando N., Eriksson L.I., Eckenhoff R.G. Perioperative neurotoxicity in the elderly: summary of the 4th International Workshop. Anesth. Analg. 2015;120(3):649–652. doi: 10.1213/ANE.0000000000000624. [DOI] [PubMed] [Google Scholar]
- 460.Williams-Russo P., Sharrock N.E., Mattis S., Szatrowski T.P., Charlson M.E. Cognitive effects after epidural vs general anesthesia in older adults. A randomized trial. JAMA. 1995;274(1):44–50. doi: 10.1001/jama.1995.03530010058035. [DOI] [PubMed] [Google Scholar]
- 461.Mason S.E., Noel-Storr A., Ritchie C.W. The impact of general and regional anesthesia on the incidence of post-operative cognitive dysfunction and post-operative delirium: a systematic review with meta-analysis. J. Alzheimers Dis. 2010;22(Suppl. 3):67–79. doi: 10.3233/JAD-2010-101086. [DOI] [PubMed] [Google Scholar]
- 462.Silbert B., Evered L., Scott D.A., McMahon S., Choong P., Ames D., Maruff P., Jamrozik K. Preexisting cognitive impairment is associated with postoperative cognitive dysfunction after hip joint replacement surgery. Anesthesiology. 2015;122(6):1224–1234. doi: 10.1097/ALN.0000000000000671. [DOI] [PubMed] [Google Scholar]
- 463.Feinkohl I., Winterer G., Spies C.D., Pischon T. Cognitive reserve and the risk of postoperative cognitive dysfunction. Dtsch. Arztebl. Int. 2017;114(7):110–117. doi: 10.3238/arztebl.2017.0110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 464.Moller J.T., Cluitmans P., Rasmussen L.S., Houx P., Rasmussen H., Canet J., Rabbitt P., Jolles J., Larsen K., Hanning C.D., Langeron O., Johnson T., Lauven P.M., Kristensen P.A., Biedler A., van Beem H., Fraidakis O., Silverstein J.H., Beneken J.E., Gravenstein J.S. Long-term postoperative cognitive dysfunction in the elderly ISPOCD1 study. ISPOCD investigators. Lancet. 1998;351(9106):857–861. doi: 10.1016/S0140-6736(97)07382-0. [DOI] [PubMed] [Google Scholar]
- 465.Newman M.F., Grocott H.P., Mathew J.P., White W.D., Landolfo K., Reves J.G., Laskowitz D.T., Mark D.B., Blumenthal J.A. Report of the substudy assessing the impact of neurocognitive function on quality of life 5 years after cardiac surgery. Stroke. 2001;32(12):2874–2881. doi: 10.1161/hs1201.099803. [DOI] [PubMed] [Google Scholar]
- 466.Rohan D., Buggy D.J., Crowley S., Ling F.K., Gallagher H., Regan C., Moriarty D.C. Increased incidence of postoperative cognitive dysfunction 24 hr after minor surgery in the elderly. Can. J. Anaesth. 2005;52(2):137–142. doi: 10.1007/BF03027718. [DOI] [PubMed] [Google Scholar]
- 467.Rörtgen D., Kloos J., Fries M., Grottke O., Rex S., Rossaint R., Coburn M. Comparison of early cognitive function and recovery after desflurane or sevoflurane anaesthesia in the elderly: a double-blinded randomized controlled trial. Br. J. Anaesth. 2010;104(2):167–174. doi: 10.1093/bja/aep369. [DOI] [PubMed] [Google Scholar]
- 468.Royse C.F., Andrews D.T., Newman S.N., Stygall J., Williams Z., Pang J., Royse A.G. The influence of propofol or desflurane on postoperative cognitive dysfunction in patients undergoing coronary artery bypass surgery. Anaesthesia. 2011;66(6):455–464. doi: 10.1111/j.1365-2044.2011.06704.x. [DOI] [PubMed] [Google Scholar]
- 469.Zhang B., Tian M., Zhen Y., Yue Y., Sherman J., Zheng H., Li S., Tanzi R.E., Marcantonio E.R., Xie Z. The effects of isoflurane and desflurane on cognitive function in humans. Anesth. Analg. 2012;114(2):410–415. doi: 10.1213/ANE.0b013e31823b2602. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 470.Chen G., Zhou Y., Shi Q., Zhou H. Comparison of early recovery and cognitive function after desflurane and sevoflurane anaesthesia in elderly patients: A meta-analysis of randomized controlled trials. J. Int. Med. Res. 2015;43(5):619–628. doi: 10.1177/0300060515591064. [DOI] [PubMed] [Google Scholar]
- 471.Tachibana S., Hayase T., Osuda M., Kazuma S., Yamakage M. Recovery of postoperative cognitive function in elderly patients after a long duration of desflurane anesthesia: a pilot study. J. Anesth. 2015;29(4):627–630. doi: 10.1007/s00540-015-1979-y. [DOI] [PubMed] [Google Scholar]
- 472.Geng Y.J., Wu Q.H., Zhang R.Q. Effect of propofol, sevoflurane, and isoflurane on postoperative cognitive dysfunction following laparoscopic cholecystectomy in elderly patients: A randomized controlled trial. J. Clin. Anesth. 2017;38:165–171. doi: 10.1016/j.jclinane.2017.02.007. [DOI] [PubMed] [Google Scholar]
- 473.Miller D., Lewis S.R., Pritchard M.W., Schofield-Robinson O.J., Shelton C.L., Alderson P., Smith A.F. Intravenous versus inhalational maintenance of anaesthesia for postoperative cognitive outcomes in elderly people undergoing non-cardiac surgery. Cochrane Database Syst. Rev. 2018;8(8):CD012317. doi: 10.1002/14651858.CD012317.pub2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 474.Zhang Y., Shan G.J., Zhang Y.X., Cao S.J., Zhu S.N., Li H.J., Ma D., Wang D.X. Propofol compared with sevoflurane general anaesthesia is associated with decreased delayed neurocognitive recovery in older adults. Br. J. Anaesth. 2018;121(3):595–604. doi: 10.1016/j.bja.2018.05.059. [DOI] [PubMed] [Google Scholar]
- 475.Qiao Y., Feng H., Zhao T., Yan H., Zhang H., Zhao X. Postoperative cognitive dysfunction after inhalational anesthesia in elderly patients undergoing major surgery: the influence of anesthetic technique, cerebral injury and systemic inflammation. BMC Anesthesiol. 2015;15:154. doi: 10.1186/s12871-015-0130-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 476.Zhang Y.H., Guo X.H., Zhang Q.M., Yan G.T., Wang T.L. Serum CRP and urinary trypsin inhibitor implicate postoperative cognitive dysfunction especially in elderly patients. Int. J. Neurosci. 2015;125(7):501–506. doi: 10.3109/00207454.2014.949341. [DOI] [PubMed] [Google Scholar]
- 477.Mathew J.P., Podgoreanu M.V., Grocott H.P., White W.D., Morris R.W., Stafford-Smith M., Mackensen G.B., Rinder C.S., Blumenthal J.A., Schwinn D.A., Newman M.F. Genetic variants in P-selectin and C-reactive protein influence susceptibility to cognitive decline after cardiac surgery. J. Am. Coll. Cardiol. 2007;49(19):1934–1942. doi: 10.1016/j.jacc.2007.01.080. [DOI] [PubMed] [Google Scholar]
- 478.Mathew J.P., Rinder C.S., Howe J.G., Fontes M., Crouch J., Newman M.F., Phillips-Bute B., Smith B.R. Platelet PlA2 polymorphism enhances risk of neurocognitive decline after cardiopulmonary bypass. Multicenter Study of Perioperative Ischemia (McSPI) Research Group. Ann. Thorac. Surg. 2001;71(2):663–666. doi: 10.1016/S0003-4975(00)02335-3. [DOI] [PubMed] [Google Scholar]
- 479.Newman M.F., Croughwell N.D., Blumenthal J.A., Lowry E., White W.D., Spillane W., Davis R.D., Jr, Glower D.D., Smith L.R., Mahanna E.P. Predictors of cognitive decline after cardiac operation. Ann. Thorac. Surg. 1995;59(5):1326–1330. doi: 10.1016/0003-4975(95)00076-W. [DOI] [PubMed] [Google Scholar]
- 480.Roses A.D. A model for susceptibility polymorphisms for complex diseases: apolipoprotein E and Alzheimer disease. Neurogenetics. 1997;1(1):3–11. doi: 10.1007/s100480050001. [DOI] [PubMed] [Google Scholar]
- 481.Perry E., Walker M., Grace J., Perry R. Acetylcholine in mind: a neurotransmitter correlate of consciousness? Trends Neurosci. 1999;22(6):273–280. doi: 10.1016/S0166-2236(98)01361-7. [DOI] [PubMed] [Google Scholar]
- 482.Fodale V., Santamaria L.B. The inhibition of central nicotinic nAch receptors is the possible cause of prolonged cognitive impairment after anesthesia. Anesth. Analg. 2003;97(4):1207. doi: 10.1213/01.ANE.0000077658.77618.C1. [DOI] [PubMed] [Google Scholar]
- 483.Fodale V., Santamaria L.B. Drugs of anesthesia, central nicotinic receptors and post-operative cognitive dysfunction. Acta Anaesthesiol. Scand. 2003;47(9):1180. doi: 10.1034/j.1399-6576.2003.00226.x. [DOI] [PubMed] [Google Scholar]
- 484.Terry R.D., Masliah E., Salmon D.P., Butters N., DeTeresa R., Hill R., Hansen L.A., Katzman R. Physical basis of cognitive alterations in Alzheimer’s disease: synapse loss is the major correlate of cognitive impairment. Ann. Neurol. 1991;30(4):572–580. doi: 10.1002/ana.410300410. [DOI] [PubMed] [Google Scholar]
- 485.Zhang B., Tian M., Zheng H., Zhen Y., Yue Y., Li T., Li S., Marcantonio E.R., Xie Z. Effects of anesthetic isoflurane and desflurane on human cerebrospinal fluid Aβ and τ level. Anesthesiology. 2013;119(1):52–60. doi: 10.1097/ALN.0b013e31828ce55d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 486.Breteler M.M., van Duijn C.M., Chandra V., Fratiglioni L., Graves A.B., Heyman A., Jorm A.F., Kokmen E., Kondo K., Mortimer J.A. Medical history and the risk of Alzheimer’s disease: a collaborative re-analysis of case-control studies. Int. J. Epidemiol. 1991;20(Suppl. 2):S36–S42. doi: 10.1093/ije/20.Supplement_2.S36. [DOI] [PubMed] [Google Scholar]
- 487.Bohnen N.I., Warner M.A., Kokmen E., Beard C.M., Kurland L.T. Alzheimer’s disease and cumulative exposure to anesthesia: a case-control study. J. Am. Geriatr. Soc. 1994;42(2):198–201. doi: 10.1111/j.1532-5415.1994.tb04952.x. [DOI] [PubMed] [Google Scholar]
- 488.Gasparini M., Vanacore N., Schiaffini C., Brusa L., Panella M., Talarico G., Bruno G., Meco G., Lenzi G.L. A case-control study on Alzheimer’s disease and exposure to anesthesia. Neurol. Sci. 2002;23(1):11–14. doi: 10.1007/s100720200017. [DOI] [PubMed] [Google Scholar]
- 489.Seitz D.P., Reimer C.L., Siddiqui N. A review of epidemiological evidence for general anesthesia as a risk factor for Alzheimer’s disease. Prog. Neuropsychopharmacol. Biol. Psychiatry. 2013;47:122–127. doi: 10.1016/j.pnpbp.2012.06.022. [DOI] [PubMed] [Google Scholar]
- 490.Chen C.W., Lin C.C., Chen K.B., Kuo Y.C., Li C.Y., Chung C.J. Increased risk of dementia in people with previous exposure to general anesthesia: a nationwide population-based case-control study. Alzheimers Dement. 2014;10(2):196–204. doi: 10.1016/j.jalz.2013.05.1766. [DOI] [PubMed] [Google Scholar]
- 491.Jevtovic-Todorovic V., Absalom A.R., Blomgren K., Brambrink A., Crosby G., Culley D.J., Fiskum G., Giffard R.G., Herold K.F., Loepke A.W., Ma D., Orser B.A., Planel E., Slikker W., Jr, Soriano S.G., Stratmann G., Vutskits L., Xie Z., Hemmings H.C., Jr Anaesthetic neurotoxicity and neuroplasticity: an expert group report and statement based on the BJA Salzburg Seminar. Br. J. Anaesth. 2013;111(2):143–151. doi: 10.1093/bja/aet177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 492.DiMaggio C., Sun L.S., Kakavouli A., Byrne M.W., Li G. A retrospective cohort study of the association of anesthesia and hernia repair surgery with behavioral and developmental disorders in young children. J. Neurosurg. Anesthesiol. 2009;21(4):286–291. doi: 10.1097/ANA.0b013e3181a71f11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 493.Wilder R.T., Flick R.P., Sprung J., Katusic S.K., Barbaresi W.J., Mickelson C., Gleich S.J., Schroeder D.R., Weaver A.L., Warner D.O. Early exposure to anesthesia and learning disabilities in a population-based birth cohort. Anesthesiology. 2009;110(4):796–804. doi: 10.1097/01.anes.0000344728.34332.5d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 494.Flick R.P., Katusic S.K., Colligan R.C., Wilder R.T., Voigt R.G., Olson M.D., Sprung J., Weaver A.L., Schroeder D.R., Warner D.O. Cognitive and behavioral outcomes after early exposure to anesthesia and surgery. Pediatrics. 2011;128(5):e1053–e1061. doi: 10.1542/peds.2011-0351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 495.DiMaggio C., Sun L.S., Ing C., Li G. Pediatric anesthesia and neurodevelopmental impairments: a Bayesian meta-analysis. J. Neurosurg. Anesthesiol. 2012;24(4):376–381. doi: 10.1097/ANA.0b013e31826a038d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 496.Ing C.H., DiMaggio C.J., Whitehouse A.J., Hegarty M.K., Sun M., von Ungern-Sternberg B.S., Davidson A.J., Wall M.M., Li G., Sun L.S. Neurodevelopmental outcomes after initial childhood anesthetic exposure between ages 3 and 10 years. J. Neurosurg. Anesthesiol. 2014;26(4):377–386. doi: 10.1097/ANA.0000000000000121. [DOI] [PubMed] [Google Scholar]
- 497.Fan C.H., Peng B., Zhang F.C. The postoperative effect of sevoflurane inhalational anesthesia on cognitive function and inflammatory response of pediatric patients. Eur. Rev. Med. Pharmacol. Sci. 2018;22(12):3971–3975. doi: 10.26355/eurrev_201806_15281. [DOI] [PubMed] [Google Scholar]
- 498.Bartels M., Althoff R.R., Boomsma D.I. Anesthesia and cognitive performance in children: no evidence for a causal relationship. Twin Res. Hum. Genet. 2009;12(3):246–253. doi: 10.1375/twin.12.3.246. [DOI] [PubMed] [Google Scholar]
- 499.Hansen T.G., Pedersen J.K., Henneberg S.W., Pedersen D.A., Murray J.C., Morton N.S., Christensen K. Academic performance in adolescence after inguinal hernia repair in infancy: a nationwide cohort study. Anesthesiology. 2011;114(5):1076–1085. doi: 10.1097/ALN.0b013e31820e77a0. [DOI] [PubMed] [Google Scholar]
- 500.Hansen T.G., Pedersen J.K., Henneberg S.W., Morton N.S., Christensen K. Educational outcome in adolescence following pyloric stenosis repair before 3 months of age: a nationwide cohort study. Paediatr. Anaesth. 2013;23(10):883–890. doi: 10.1111/pan.12225. [DOI] [PubMed] [Google Scholar]
- 501.Davidson A.J., Disma N., de Graaff J.C., Withington D.E., Dorris L., Bell G., Stargatt R., Bellinger D.C., Schuster T., Arnup S.J., Hardy P., Hunt R.W., Takagi M.J., Giribaldi G., Hartmann P.L., Salvo I., Morton N.S., von Ungern Sternberg B.S., Locatelli B.G., Wilton N., Lynn A., Thomas J.J., Polaner D., Bagshaw O., Szmuk P., Absalom A.R., Frawley G., Berde C., Ormond G.D., Marmor J., McCann M.E. Neurodevelopmental outcome at 2 years of age after general anaesthesia and awake-regional anaesthesia in infancy (GAS): an international multicentre, randomised controlled trial. Lancet. 2016;387(10015):239–250. doi: 10.1016/S0140-6736(15)00608-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 502.McCann M.E., de Graaff J.C., Dorris L., Disma N., Withington D., Bell G., Grobler A., Stargatt R., Hunt R.W., Sheppard S.J., Marmor J., Giribaldi G., Bellinger D.C., Hartmann P.L., Hardy P., Frawley G., Izzo F., von Ungern Sternberg B.S., Lynn A., Wilton N., Mueller M., Polaner D.M., Absalom A.R., Szmuk P., Morton N., Berde C., Soriano S., Davidson A.J. Neurodevelopmental outcome at 5 years of age after general anaesthesia or awake-regional anaesthesia in infancy (GAS): an international, multicentre, randomised, controlled equivalence trial. Lancet. 2019;393(10172):664–677. doi: 10.1016/S0140-6736(18)32485-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 503.Sun L.S., Li G., Miller T.L., Salorio C., Byrne M.W., Bellinger D.C., Ing C., Park R., Radcliffe J., Hays S.R., DiMaggio C.J., Cooper T.J., Rauh V., Maxwell L.G., Youn A., McGowan F.X. Association between a single general anesthesia exposure before age 36 months and neurocognitive outcomes in later childhood. JAMA. 2016;315(21):2312–2320. doi: 10.1001/jama.2016.6967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 504.Warner D.O., Zaccariello M.J., Katusic S.K., Schroeder D.R., Hanson A.C., Schulte P.J., Buenvenida S.L., Gleich S.J., Wilder R.T., Sprung J., Hu D., Voigt R.G., Paule M.G., Chelonis J.J., Flick R.P. Neuropsychological and behavioral outcomes after exposure of young children to procedures requiring general anesthesia: the mayo anesthesia safety in kids (mask) study. Anesthesiology. 2018;129(1):89–105. doi: 10.1097/ALN.0000000000002232. [DOI] [PMC free article] [PubMed] [Google Scholar]
