Abstract
Neuronal voltage-gated sodium channels (Nav) are major targets for the neurophysiological actions of general anesthetics. In the adult brain, cell type-specific effects on synaptic transmission are attributed to the differential sensitivity to volatile anesthetics of specific Nav subtypes preferentially expressed in mature neurons (Nav1.1, Nav1.2, Nav1.6). Comparatively, developing neurons are more excitable than mature neurons. We determined volatile anesthetic effects on Na+ currents mediated by Nav1.3, the principal Nav subtype expressed in developing neurons. Sevoflurane at clinical concentrations inhibited peak Na+ current of human Nav1.3 heterologously expressed in HEK293T cells in a voltage-dependent manner, induced a − 6.1 mV hyperpolarizing shift in the voltage dependence of steady-state inactivation, and slowed recovery from fast inactivation. Nav1.3-mediated Na+ currents also exhibited distinct activation properties associated with hyperexcitability, including prominent persistent currents and ramp currents, both of which were significantly reduced by sevoflurane. Nav1.3 showed a more depolarized voltage dependence of steady-state inactivation than Nav1.2, consistent with its higher propensity for sustained repetitive firing. Nav1.2 exhibited minimal persistent and ramp currents, and these were unaffected by sevoflurane. These findings identify subtype-specific effects of sevoflurane on neuronal Nav subtype electrophysiological properties, and suggest a mechanistic basis for increased anesthetic sensitivity and toxicity in early neuronal differentiation and maturation.
Supplementary Information
The online version contains supplementary material available at 10.1038/s41598-025-15280-6.
Keywords: Voltage-gated sodium channel, Persistent current, Ramp current, Anesthetic mechanism, Volatile anesthetic
Subject terms: Biophysics, Neuroscience, Physiology
Introduction
Voltage-gated Na+ channels (Nav) regulate neuronal excitability1–4 action potential (AP)-driven Ca2+ influx, and Ca2+-dependent neurotransmitter release3,5–8 and are important targets for general anesthetics9–12. Upon neuronal depolarization, Nav rapidly activate then inactivate generating an AP, followed by a return to resting membrane potential13. Volatile anesthetics (VA) directly inhibit heterologously expressed Nav11 by reducing AP amplitude and nerve terminal depolarization in various neuronal preparations14–17 including dissociated neurons from the hippocampus18 a major site of anesthetic actions19.
Four major Nav subtypes (Nav1.1, Nav1.2, Nav1.3, Nav1.6) are expressed in the mammalian brain20,21, with Nav1.3 preferentially expressed during early development22,23. In the hippocampus, variable expression of Nav subtypes in specific neuronal populations influences cellular and network excitability due to differential presynaptic and postsynaptic localizations8. Mature inhibitory hippocampal interneurons express more Nav1.1 compared to excitatory pyramidal neurons that express relatively more Nav1.2 and Nav1.624–26. These Nav subtypes are inhibited by VAs in two ways: (1) stabilization of the inactivated state resulting in a shift of steady-state inactivation to 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 block9,27,28. Voltage- and frequency-dependent block of Nav17,28,29 reduces AP amplitude15,30 and neurotransmitter release15,31 with Nav subtype and neurotransmitter selectivity. For example, isoflurane decreases glutamate release more potently than GABA release from dissociated hippocampal neurons32. This is likely mediated through greater inhibition of Nav1.2 and Nav1.6 in glutamatergic boutons compared with lesser inhibition of Nav1.1 in GABAergic boutons18,33.
Heterogeneity in Nav expression also occurs during neuronal differentiation and development. Expression of Nav subtypes with distinct activation properties renders developing neurons hyperexcitable compared to mature neurons22. Compared to other subtypes, Nav1.3 recovers from inactivation four-fold faster and has slower closed-state inactivation, reducing the threshold for activation in neurons that express Nav1.3, leading to sustained repetition firing34–37 even with unchanged or absent stimuli. Slower inactivation or resurgent Na+ currents enhance repetitive firing and modulate overall neuronal excitability as opposed to AP initiation and propagation38,39. Although persistent Na+ currents in acute hippocampal slices are inhibited by isoflurane40, subtype-selective effects of anesthetic sensitivity on developmentally expressed Nav have not been investigated. We examined the effects of sevoflurane on the function of Nav1.3, a major neuronal subtype developmentally expressed in the immature mammalian brain. Comparative studies with Nav1.2 provide a neurophysiological basis for developmental stage-specific impact on neuronal excitability.
Results
Effects of sevoflurane on peak Na+ current
The effects of sevoflurane on Nav1.3 currents were tested using a periodic stimulation protocol that allowed us to determine voltage-dependent drug effects10,41,42. To assess peak Na+ current (INa) inhibition, a depolarizing pulse to 0 mV was applied every 5 s, following a 100 ms prepulse to either V0 (–120 mV), where most channels remained in an available resting state, or to a potential at which ~ 50% of channels were in an inactivated state V½−inact (–47 ± 5 mV). Figure 1b shows the time course of INa inhibition during sevoflurane wash-in. The slow onset likely reflects drug partitioning into intracellular compartments rather than an intrinsically slow binding rate to the channel43. Results are expressed as the ratio of peak INa in the presence of sevoflurane to control INa (Fig. 1a-c, Supplemental Table S1). Sevoflurane significantly inhibited peak INa following a prepulse to V½−inact in a concentration-dependent manner. Time control (sham) experiments using external solution without sevoflurane showed no time-dependent inhibition (Supplemental Fig. S2a, b; Supplemental Table S3). In addition to reducing peak INa, sevoflurane also accelerated macroscopic current decay. For Nav1.3, current decay of inactivation was best fit with a bi-exponential function, consistent with two components of inactivation, and sevoflurane significantly reduced the fast time constant τfast without affecting τslow (Supplemental Fig. S1a-c). In contrast, Nav1.2 current decay was adequately fit with a mono-exponential function, and τfast was similarly reduced by sevoflurane (Supplemental Fig. S1d-f).
Fig. 1.
Inhibition of peak Nav1.3 current by sevoflurane. Sevoflurane inhibited peak Na+ current (INa) in a voltage-dependent manner at concentrations equivalent to 0.5 MAC, 1 MAC, or 2 MAC (0.14 mM, 0.28 mM, or 0.57 mM). a, Macroscopic Na+ currents from a transfected HEK293T cell expressing Nav1.3 were evoked using a periodic alternating stimulation protocol (see inset) in the absence (CTL, black traces) or presence (SEVO, orange traces) of 1 MAC sevoflurane (0.28 mM). Sevoflurane did not inhibit peak INa when the test pulse followed a prepulse to V0, a potential at which all channels were in the resting state. However, sevoflurane significantly inhibited peak INa with a prepulse to V½−inact, a potential at which half of the channels were in a fast-inactivated state. b, The time course of the effects of wash–in and wash–out of 1 MAC sevoflurane. c, Concentration-dependent inhibition of Nav1.3 current by sevoflurane at 0.5, 1, and 2 MAC (for values, see Supplemental Table S1) following prepulse to V0 or V½−inact. All data are mean ± SD, n = 6–7, **P < 0.01 vs. control by paired two-tailed Student t-test (P = 0.0086 for 0.5 MAC; P = 0.0053 for 1 MAC; P = 0.0037 for 2 MAC).
Effects of sevoflurane on activation and inactivation
Sevoflurane caused a small but significant hyperpolarizing shift in the voltage dependence of channel activation, indicating that channels opened at slightly more negative potentials (Fig. 2a, b). This suggests that sevoflurane facilitates activation by lowering the voltage threshold for channel opening. Steady-state fast inactivation, also referred to as channel availability (or h∞ from the Hodgkin-Huxley model), was tested using a double-pulse protocol (Fig. 2a, inset). Sevoflurane significantly shifted the voltage dependence of steady-state inactivation by − 6.1 ± 2.7 mV toward more hyperpolarized potentials (Fig. 2c, Supplemental Table S1). No such shift was seen in time-control (sham) experiments using external solution without sevoflurane, confirming no significant time-dependent shift in V½−inact (Supplemental Fig. S2c, d; Supplemental Table S3). This indicates that a greater fraction of channels either transitioned into and/or remained in the inactivated state at any given potential, reducing the pool of available channels for subsequent activation. The shift in V½−inact was concentration-dependent, with the largest effect observed at 2 MAC (Fig. 2d). These findings are consistent with sevoflurane stabilization of the inactivated state, thereby limiting Nav availability, while also slightly modulating activation properties.
Fig. 2.
Effects of sevoflurane on Nav1.3 activation and inactivation. a, Representative families of whole-cell inward INa evoked by depolarisation (inset) from a holding potential (Vh) of − 80 mV in the absence (CTL; left, black traces) or presence of sevoflurane (1 MAC; 0.28 mM) (SEVO; right, orange traces). b, c, Current-voltage relationship of channel activation displayed as normalized conductance (G/Gmax), and of inactivation (INa/INamax). Sevoflurane shifted the voltage dependence of activation by − 6.2 ± 3.2 mV (V½−activ−17.6 ± 6.0 mV for CTL [white circles] vs. −23.7 ± 6.0 mV for SEVO [orange circles], P = 0.0026, n = 7). Sevoflurane also shifted the voltage for half-maximal inactivation (V½−inact) by − 6.1 ± 2.7 mV toward hyperpolarized potentials (V½−inact −44.8 ± 7.0 mV for CTL [white circles] vs. −50.9 ± 7.3 mV for SEVO [orange circles], P < 0.0001, n = 11). d, This shift in V½−inact was concentration-dependent, with the strongest effect of a − 10.4 ± 5.3 mV shift seen at 2 MAC sevoflurane. All data are mean ± SD; **P < 0.01, ****P < 0.0001 vs. control by paired two-tailed Student t-test.
Sevoflurane slows recovery from fast inactivation
We measured the effects of sevoflurane on recovery from inactivation using a holding potential of − 120 mV, at which the majority of channels were in a resting, closed state (Fig. 3a-c). Sevoflurane (1 MAC; 0.28 mM) significantly slowed the fast component of recovery, as reflected by an increase in the fast time constant τfast, while the slow component τslow was unaffected (Supplementary Table S1). This indicates that sevoflurane selectively impairs the rapid recovery of Nav1.3 from fast inactivation, potentially prolonging the refractory period without altering slower recovery kinetics.
Fig. 3.
Effects of sevoflurane on Nav1.3 recovery from fast inactivation. We used a two-pulse protocol from a holding potential (Vh) of − 120 mV, with two 10 ms test pulses to 0 mV separated by an inter-pulse interval of 1–200 ms (inset in (a) shows stimulation protocol). a, Overlaid macroscopic INa traces for the two pulses with increasing inter-pulse intervals in the absence (CTL, black traces) or presence (SEVO, orange traces) of sevoflurane (1 MAC; 0.28 mM). Peak INa of the second test pulse slowly recovers to control values with increasing inter-pulse durations. b, Data fitted to a bi-exponential equation and plotted against inter-pulse-interval in the absence (white circles, dotted line) or presence (orange circles, solid orange line) of sevoflurane. c, Recovery time constants τfast and τslow in the absence (white circles) or presence (orange circles) of sevoflurane. Sevoflurane increased τfast from 1.55 ± 0.4 ms to 2.07 ± 0.5 ms (P = 0.0073), thereby slowing recovery from fast inactivation; τslow was not affected with 55 ± 30 ms for CTL and 40 ± 30 ms for SEVO; P = 0.3789. Data are mean ± SD, n = 7, **P < 0.01 vs. control by paired two-tailed Student t-test.
Nav1.3 and Nav1.2 have distinct electrophysiological properties
We compared the effects of sevoflurane on Nav1.3 and Nav1.2 to investigate possible subtype-selective actions on developmentally distinct Nav subtypes. Under control conditions, the voltage-dependent properties of Nav1.3 were more depolarized compared to Nav1.2, with the voltage for half-maximal activation (V½−activ) and inactivation (V½−inact) significantly shifted by + 10 mV (Figs. 2 and 4a-c, Supplemental Table S1,S2). These differences indicate that Nav1.3 channels exhibit depolarized gating, enhancing their likelihood of opening during action potentials and maintaining availability over a wider voltage range, potentially influencing neuronal excitability.
Fig. 4.
Effects of sevoflurane on Nav1.2 peak INa, steady-state fast inactivation, and recovery from fast inactivation. a, Representative families of whole-cell inward Na+ currents evoked by depolarization (inset) from a holding potential (Vh) of − 80 mV in the absence (CTL; left, black traces) or presence of sevoflurane (1 MAC; 0.28 mM) (SEVO; right, orange traces). b, c, Current-voltage relationship of channel activation displayed as normalized conductance (G/Gmax) and inactivation (INa/INamax). Sevoflurane did not affect the voltage dependence of activation (V½−activ −27.3 ± 5.0 mV for CTL [white diamonds] vs. −29.4 ± 4.2 mV for SEVO [orange diamonds], P = 0.0544, n = 6). However, sevoflurane shifted the voltage for half-maximal inactivation (V½−inact) by − 6.1 ± 1.8 mV toward hyperpolarized potentials (V½−inact −54.2 ± 5.0 mV for CTL [white diamonds] vs. −60.3 ± 6.2 mV for SEVO [orange diamonds], P < 0.0005, n = 6). d, Effect of sevoflurane on peak INa (stimulation protocol see Fig. 1a). Sevoflurane did not inhibit peak INa when the test pulse followed a prepulse to V0, a potential at which all channels were in the resting state. However, with a prepulse to V½−inact, a potential at which half of the channels were in the fast-inactivated state, sevoflurane inhibited peak INa (normalized peak INa 0.97 ± 0.05 for V0, P = 0.2238 and 0.73 ± 0.08 for V½−inact, P = 0.0005, n = 6). e, f, Recovery from fast inactivation was tested using the same protocol as in Fig. 3b. e, Recovery data (Pulse2/Pulse1) were fitted to a bi-exponential equation and plotted against inter-pulse-interval in the absence (white diamonds, dotted line) or presence (orange diamonds, solid orange line) of sevoflurane. f, Recovery time constants τfast and τslow in the absence (white diamonds) or presence (orange diamonds) of sevoflurane. Sevoflurane increased τfast from 1.7 ± 0.4 ms to 2.8 ± 0.7 ms (P = 0.0039), thereby slowing recovery from fast inactivation (τslow was not affected with 40 ± 30 ms for CTL and 70.2 ± 70 ms for SEVO; n = 6, P = 0.3257). See Supplemental Table S2 for all values. Data are mean ± SD, **P < 0.01, ***P < 0.001 vs. control by paired two-tailed Student t-test.
Sevoflurane at a clinically relevant concentration of 1 MAC (0.28 mM) produced similar effects on Nav1.2 (Fig. 4a-f) as observed with Nav1.3. Specifically, sevoflurane significantly inhibited peak INa of Nav1.2 when the test pulse followed a prepulse to V½−inact (− 56 ± 7 mV), a potential at which ~ 50% of channels are in a fast-inactivated state, but not with a prepulse to V0 (− 120 mV) (Fig. 4d). The voltage dependence of steady-state fast inactivation was shifted by ~ − 6 mV toward more hyperpolarized potentials (Fig. 4b, c). Sevoflurane also slowed recovery from fast inactivation, with a significant increase in the fast time constant τfast (Fig. 4e, f). Sevoflurane did not significantly alter the activation properties of Nav1.2, highlighting a more pronounced effect on Nav1.3, where the drug modulates both activation and inactivation gating. This action might be particularly relevant in neurons where Nav1.3 is upregulated, such as during development or following injury.
Effects of sevoflurane on persistent Na+ current
The distinct gating mechanisms of Nav1.3 result in persistent Na+ current (INaP), which enhances repetitive firing and hyperexcitability during neuronal development34,36. INaP allows subthreshold depolarizations to generate APs, but it is unclear whether and how this is modulated by VAs. A large INaP was generated by Nav1.3 which was significantly inhibited by 1 MAC sevoflurane (Fig. 5a-c). INaP generated by Nav1.3 was ~ 6-fold greater than for Nav1.2, which was minimal and not further reduced by sevoflurane (–146 ± 161 pA for Nav1.3 vs. − 23 ± 19 pA for Nav1.2; Fig. 5d-f). To further validate this persistent current component of Nav1.3, we conducted additional experiments using a prolonged depolarizing step (120 ms), which confirmed the presence of a sustained non-inactivating current throughout the pulse (Supplemental Fig. S3). This current was similarly reduced by sevoflurane, and full block by tetrodotoxin (TTX) confirmed that the residual current was not due to leak or non-specific conductance.
Fig. 5.
Effects of sevoflurane on Nav1.3 and Nav1.2 persistent Na+ current (INaP). INaP was tested using the depicted stimulation protocol (a, see inset). a, d, Overlaid averaged macroscopic Na+ currents in the absence (black traces) or presence (orange traces) of sevoflurane (1 MAC; 0.28 mM). Right panels show a zoomed view of the final 5 ms of the 20 ms depolarization, highlighting INaP (CTL black traces, SEVO orange traces, shaded areas represent SD). b, e, Absolute INaP values in pA, calculated as the mean current of the final 5 ms. Sevoflurane reduced INaP in Nav1.3 (from − 146 ± 161 pA for CTL [white circles] to − 65 ± 83 pA for SEVO [orange circles], P = 0.0144, n = 10) but had no effect on Nav1.2 (from − 23 ± 19 pA for CTL [white diamonds] to − 25 ± 23 pA for SEVO [orange diamonds], n = 9). c, f, INaP expressed as a percentage of peak INa, to account for variability in cell size and peak INa amplitude across recordings. Sevoflurane reduced INaP in Nav1.3 from 2.9 ± 1.2% [CTL, white circles] to 1.3 ± 0.9% [SEVO, orange circles], P < 0.0011, n = 9), but not in Nav1.2 (from 1.0 ± 0.8% for CTL [white diamonds] to 0.7 ± 0.6% for SEVO [orange diamonds], n = 9). Data are mean ± SD, *P < 0.05, **P < 0.01 vs. control by paired two-tailed Student t-test.
Effects of sevoflurane on ramp currents
Inward Na+ currents during slow depolarizing ramps contribute to neuronal hyperexcitability by promoting repetitive firing. We examined the effects of sevoflurane on ramp-induced currents (INaR) using a depolarizing voltage ramp from − 120 mV to + 40 mV over 800 ms (0.2 mV/ms). INaR was quantified by integrating the inward INa over the voltage range from the activation threshold to the reversal potential, yielding the total charge transfer. To account for variability in channel expression levels among cells, this charge transfer was normalized to the peak INa elicited by a standard step depolarization, resulting in a normalized charge transfer value (pC/nA). Under control conditions, Nav1.3 mediated a greater normalized charge transfer during the ramp protocol compared to Nav1.2 (9.6 ± 2.0 pC/nA for Nav1.3 vs. 3.5 ± 1.0 pC/nA for Nav1.2; data from Fig. 6b). Sevoflurane at a clinically relevant concentration of 1 MAC (0.28 mM) significantly reduced the normalized charge transfer associated with INaR in Nav1.3-expressing cells (Fig. 6a, b). In contrast, sevoflurane did not alter the normalized charge transfer in Nav1.2 (Fig. 6b), suggesting a subtype-specific effect on Na+ channel ramp currents.
Fig. 6.
Sevoflurane selectively inhibits ramp-evoked Na+ currents (INaR) in Nav1.3 but not Nav1.2. a, Mean of all ramp currents recorded in response to a slow depolarizing ramp from − 120 mV to + 40 mV over 800 ms (0.2 mV/ms) for Nav1.3 (left panel) or Nav1.2 (right panel). Top panels show stimulation protocol. Bottom panel shows mean ramp current trace in the absence (CTL, solid black trace) or presence (SEVO, solid orange trace) of sevoflurane (1 MAC; 0.28 mM). The shaded areas show SEM for control (CTL, grey area) or sevoflurane (SEVO, orange area). Currents are normalized to peak INa (evoked by a depolarizing step to 0 mV) and expressed as a percentage. b, Normalized charge transfer (pC/nA) was calculated by integrating inward INa from the voltage at which currents activate to the voltage at which currents reverse polarity, followed by normalization to peak INa. Under control conditions, Nav1.3 showed greater normalized charge transfer than Nav1.2 (9.6 ± 5.7 pC/nA for Nav1.3 vs. 3.5 ± 2.7 pC/nA for Nav1.2, P = 0.0226 by unpaired two-tailed Student t-test). Sevoflurane reduced charge transfer in Nav1.3 (9.6 ± 5.7 to 5.6 ± 3.1 pC/nA, P = 0.0203, n = 8), but not in Nav1.2 (3.5 ± 2.7 to 1.7 ± 1.3 pC/nA, P = 0.1680 n = 7). Data are presented as mean ± SD. *P < 0.05 vs. control by paired two-tailed Student t-test.
Discussion
The developmentally expressed Na+ channel Nav1.3 is sensitive to modulation by the volatile anesthetic sevoflurane. While the effects of volatile anesthetics on Nav1.2 have been characterized previously, their impact on developmentally expressed Nav1.3 were previously unknown. We compared the effects of sevoflurane on the channel gating properties of human Nav1.3 and Nav1.2, the predominant Na+ channel subtypes expressed in immature and mature neurons, respectively. Under control conditions, Nav1.3 exhibited gating at more depolarized potentials compared to Nav1.2: both the voltage of half-maximal activation (V½−activ) and inactivation (V½−inact) were ~ 10 mV more positive in Nav1.3. Despite these differences, sevoflurane produced a comparable hyperpolarizing shift in steady-state fast inactivation for both subtypes. The degree of voltage-dependent inhibition of peak INa and the slowing of recovery from fast inactivation were also similar between Nav1.3 and Nav1.2, suggesting that baseline differences in gating do not substantially alter the overall pattern of modulation by sevoflurane.
Persistent Na+ current (INaP) can drive intrinsic neuronal hyperexcitability44 and has been implicated in various physiological and pathological processes, including pacemaking, neuronal injury, chronic pain, and epilepsy. We observed large INaP in cells expressing Nav1.3, consistent with enhanced excitability of immature neurons. Sevoflurane reduced INaP in Nav1.3 expressing cells, while Nav1.2 had much smaller persistent currents than Nav1.3, and were not further reduced by sevoflurane.
Ramp currents (INaR), which can enhance responses to subthreshold depolarizations by reducing action potential firing threshold45were more pronounced in Nav1.3 expressing cells. This observation is consistent with studies indicating that mutations in Nav1.3 linked to neuronal hyperexcitability increase ramp and persistent currents. Selective inhibition of these currents could mediate developmental stage-specific effects of sevoflurane and possibly other anesthetics on neuronal excitability.
Multiple volatile anesthetics have been shown to inhibit all neuronal Nav subtypes, including Nav1.1, Nav1.2, and Nav1.6, leading to reduced AP amplitude in mature neurons at clinically relevant concentrations11,15,31,33. In addition to shifting the voltage dependence of steady-state fast inactivation and slowing channel recovery, our findings show that sevoflurane also reduces INaP and INaR in Nav1.3-expressing cells, providing a basis for previous observations that volatile anesthetics suppress transient Na+ current (INaT) and INaP, thereby reducing hippocampal excitability40. Thus, during brain maturation, multiple Nav1.3 gating mechanisms are likely targeted by volatile anesthetics to inhibit Nav-activated Ca2+-influx and neurotransmitter release that contribute to the neurophysiological actions of anesthetics.
Modulation of neuronal activity by general anesthetics during early development can lead to persistent alterations in neuronal survival and function46,47. The neurotoxic effects of general anesthetics administered at critical periods early in development in the mammalian brain48–50 are associated with structural and functional changes, including alterations in dendritic spine dynamics, dendritic arborization, and synaptic protein expression48–50. General anesthetics have selective effects on neurotransmission by depressing excitatory and enhancing inhibitory synaptic transmission, which potentially alters the balance of excitotoxicity from glutamatergic transmission in immature neuronal networks51,52. The relative differences in anesthetic effects on excitatory or inhibitory neurons vary between anesthetic agents with synaptic inhibition by modulation of GABAA receptors essential for intravenous agents such as propofol, while reduction of glutamatergic transmission both presynaptically and postsynaptically contributes to the neurodepressant effects of volatile agents like sevoflurane53–55. These distinctions influence early deleterious actions as administration of drugs that attenuate GABAA receptor activity or potentiate AMPA signaling during emergence from neonatal general anesthesia can ameliorate neuronal circuit dysfunction and learning deficits under propofol and sevoflurane, respectively46,47. As Nav1.3 is predominantly expressed in excitatory neurons56 and drives persistent excitability, our findings suggest a potential role for Nav1.3 in the effects of volatile anesthetic exposure to alter activity of immature, excitatory brain networks. Further investigation is needed to understand compensatory actions on inhibitory networks as well as regulation of Nav1.3 by intravenous agents.
In native neurons, Na+ channels are composed of a pore-forming α-subunit and one or more auxiliary β-subunits (β1-β4). Auxiliary subunits are critical in regulating channel function, cell surface expression, and downstream signaling. Thus, co-expression with β-subunit(s) more accurately mimics endogenous neuronal physiology, particularly in heterologous expression systems such as HEK293 cells. Out of the four β-subunits, Navβ1 is ubiquitously expressed in the CNS and its structure-function effects are well-characterized compared to other subunits. It has strong modulatory effects on channel properties including voltage dependence, channel kinetics, recovery, and surface expression57–59. Specifically, β1 subunit accelerates both activation and inactivation of Nav1.3 currents, shifts steady-state inactivation to more hyperpolarized potentials, and speeds up recovery of Nav1.3 from inactivation59,60. In addition, β1 can enhance membrane trafficking and aid stabilization of Nav1.3 at the cell surface. These properties lead to sharper Nav1.3 Na+ currents and supports higher firing frequencies in cells59,60. Although in slightly reduced magnitude, co-expression of the β1 subunit similarly modulates Nav1.2 compared with Nav1.361,62 allowing for reproducible and reliable data collection.
Our baseline Nav1.3 activation properties (V½−activ) including INaP are consistent with previous reports45,63–65. However, our V½−inact values were notably more depolarized at ∼–45 mV compared to ∼–70 mV in those studies. Additionally, we observed larger ramp currents and normalized charge transfer in Nav1.3-expressing cells than those reported by Vanoye et al.64,65. These differences might stem from alterations in holding potentials (Vh) (we chose a more physiological Vh for certain experiments) or differences in auxiliary subunit expression, as we co-expressed only the β1-subunit, whereas Vanoye et al. included both β1 and β2. The β2 subunit is known to modulate inactivation gating and reduce persistent current amplitude60,66,67which might account for the more depolarized inactivation and enhanced ramp currents observed in our study.
Multiple binding sites have been proposed for volatile anesthetic interactions with Nav68–70; further studies are required to determine whether differences in anesthetic binding to specific sites contribute to these subtype-selective effects of sevoflurane. Recent structure-function studies in bacterial Na+ channels has revealed that sevoflurane can displace membrane lipids and bind within a conserved hydrophobic pocket formed between the S1–S4 helices of one subunit and the S5 helix of an adjacent subunit71. Whether a homologous site exists, or is functionally relevant, within the pseudotetrameric domains of mammalian Nav remains an important question for future work. Various volatile anesthetics might exhibit different affinities for Nav subtypes72,73with sevoflurane showing stronger interactions with Nav1.3 compared with Nav1.2. Moreover, Nav subtypes not only show cell-type specific expression7 but also distinct subcellular and cellular localizations8. Parvalbumin-expressing interneurons are enriched in Nav1.1, while excitatory pyramidal neurons express more Nav1.2, Nav1.3, and Nav1.67,56,74,75. While all Nav subtypes are found at the axon initial segment (AIS) to some degree, Nav1.1 and Nav1.2 are preferentially expressed in the proximal AIS76potentially allowing direct targeting of AP initiation by anesthetics.
The halogenated ethers isoflurane, sevoflurane, and desflurane represent the principal volatile anesthetics employed clinically in modern anesthesia. However, isoflurane and sevoflurane are more commonly used due to their safety profiles and cost-effectiveness. Considerable evidence from our group10,11,33,70,77–80 and others81–84 show inhibition of neuronal Nav by isoflurane with subtype-specificity. For example, Nav1.1 is less sensitive to isoflurane compared to Nav1.2 and 1.6 due to its distinct gating properties33. Although similar, sevoflurane has distinct differences in pharmacological actions and clinical use, gaining broader acceptance and use in modern anesthesia practice. For instance, sevoflurane’s rapid onset and offset leads to faster emergence and recovery and causes less airway irritation than isoflurane85,86. These characteristics make sevoflurane more suitable for various types of surgery, including in pediatric populations with increased Nav1.3 expression84. Although sensitivity to sevoflurane for neuronal Nav has been reported81,84,87, systematic effects of sevoflurane on selective neuronal Nav subtypes or comparative effects between Nav subtypes have not been reported. Here, we report agent-specific differences in efficacy for Na+ channel inhibition that could underlie differences in modulation of immature and mature brain networks.
Further studies of the selective effects of sevoflurane on Nav subtypes could lead to the development of more selective and potentially safer anesthetic agents that selectively target specific Nav subtypes based on neuronal stage or neuropathological state by reducing deleterious off target effects or specifically targeting desired effects.
Materials and methods
Materials
Sevoflurane was obtained from Henry Schein Medical (Melville, NY). HEK293T cells were from ATCC (Manassas, VA). All other chemicals were from Sigma-Aldrich (St. Louis, MO).
cDNA constructs
Human Nav1.2-pIR-CMV-IRES-mScarlet (Accession number NM_021007) and human Nav1.3-pIR (Accession number NM_001081676) cDNA were kindly provided by Professor A.L. George Jr. (Northwestern University, Evanston, IL) via AddGene (Watertown, MA) and Professor J. A. Kearney (Northwestern University, Evanston, IL), respectively.
Cell culture and electrophysiological recordings of Na+ currents
Human HEK293T cells (#CRL-3216, ATCC, Manassas, VA) were seeded into 35-mm dishes in Dulbecco’s Modified Eagle′s Medium (Sigma-Aldrich) supplemented with 10% (v/v) heat-inactivated fetal bovine serum, 100 U/ml penicillin, and 100 mg/ml streptomycin and incubated in a humidified atmosphere of 5% CO2 and 95% air at 37 °C. HEK293T cells were cotransfected with cDNA for hNav1.2 or hNav1.3 (3–4 µg) and human auxiliary hβ1 subunit (0.8–1 µg) using Lipofectamine LTX (Invitrogen, Carlsbad, CA). Additionally, hNav1.3 was cotransfected with 0.8 µg pEGFP-N1 (Clontech, Mountain View, CA) as a reporter to allow identification of EGFP-transfected cells by fluorescence microscopy. Cells were re-seeded at lower density onto 12-mm glass coverslips 20–24 h post-transfection, and electrophysiological studies were conducted at least 3 h later to allow cells to adhere.
For electrophysiological recordings, coverslips were transferred into a small volume laminar-flow perfusion chamber (Warner Instruments, Hamden, CT) and continuously perfused at ~ 2 ml/min with extracellular solution containing (in mM): 130 mM NaCl, 10 mM HEPES, 3.25 mM KCl, 2 mM MgCl2, 2 mM CaCl2, 0.1 mM CdCl2, and 20 mM TEA-Cl, adjusted to pH 7.4 (with NaOH) and to 310 mOsm/kg. Na+ currents were recorded in voltage-clamp mode at room temperature (23–24 °C) using an Axopatch 200B patch-clamp amplifier (Molecular Devices, San Jose, CA) with a 5 kHz low-pass filter at a sampling rate of 20 kHz. Recording pipettes were pulled from borosilicate glass capillaries (Sutter Instruments, Novato, CA) on a P-1000 horizontal puller (Sutter Instruments) and fire-polished before use. Pipette resistance was ~ 1.5–2.5 MΩ when filled with internal solution containing (in mM): 120 mM CsF, 10 mM NaCl, 10 mM HEPES, 10 mM EGTA, 10 mM TEA-Cl, 1 mM CaCl2, and 1 mM MgCl2 adjusted to pH 7.3 (with CsOH) and to 312 mOsm/kg (with sucrose). Access resistance (~ 2–4 MΩ) was reduced using 75–85% series resistance correction. To minimize space-clamp and series resistance errors, only cells expressing 1–8 nA peak current were analyzed. Initial whole cell seal resistance was 1–3 GΩ, and recordings were discarded if resistance dropped below 1 GΩ. Liquid–junction potentials were not corrected. Capacitive current transients were electronically canceled with the internal amplifier circuitry, and leak currents were digitally subtracted using the P/4 protocol88 Recordings began 5 min after attaining the whole-cell patch configuration to allow equilibration of pipette solution and cytosol. Stimulation protocols were applied in control external solution, and again after a 3-minute superperfusion with external solution containing sevoflurane.
Sevoflurane superfusion
A saturated stock solution of sevoflurane in external solution (~ 5.2 mM at 23 °C) was diluted to the desired final concentration in a gas-tight glass syringe. Sevoflurane solutions were superfused using a pressure-driven microperfusion system (ALA Scientific, Westbury, NY) with a 200 μm diameter perfusion pipette tip positioned ~ 200 μm from the recorded cell. Sevoflurane concentrations sampled at the perfusion pipette tip were determined by gas chromatography using a Shimadzu GC-2010 Plus gas chromatograph (Shimadzu, Tokyo, Japan) after extraction into octane89. An aqueous sevoflurane solution of 0.28 mM was used as the predicted minimum alveolar concentration (MAC; equivalent to the EC50 for immobilization) in humans after temperature adjustment to 24°C30,90. To ensure reproducibility and minimize variability, the time from break-in to initial control recordings was standardized across all experiments, as was the duration of sevoflurane exposure before drug measurements. This protocol was based on previously validated methods shown to maintain stable gating properties over time in similar experimental conditions10,42,70. In addition, sham (time control) recordings were performed in a subset of cells that underwent identical timing and perfusion procedures using external solution without sevoflurane. These experiments confirmed the absence of time-dependent shifts in the voltage-dependence of inactivation or peak current amplitude (see Supplemental Fig. S2 and Supplemental Table S3), thereby supporting the specificity of drug-induced effects reported in the main text.
Stimulation protocols and data analysis
The holding potential (Vh) was − 80 mV unless otherwise stated. Voltage-dependent inhibition of peak inward Na+ current (INa) was analyzed using a 20 ms test pulse to 0 mV preceded by a 100 ms prepulse alternating between either − 120 mV (V0) or the voltage of half-maximal inactivation (V½−inact = − 47 ± 5 mV) applied every 5 s10,41,42. V½−inact was determined for each individual cell using the double-pulse protocol for steady-state inactivation described below. For voltage-dependence of activation, conductance (G) values were derived from the current-voltage (I-V) relationship using:
![]() |
where I is the peak Na+ current (INa) at a given command voltage (V), and Vrev is the measured Na+ reversal potential. Conductance values were normalized to maximal conductance (G/Gmax), and the activation curve was fitted using a Boltzmann function:
![]() |
where V½-act is the voltage at which activation is half-maximal, and k is the slope factor describing the voltage sensitivity of activation.
Steady-state fast inactivation (h∞), was measured using a double-pulse protocol with a 100 ms prepulse ranging from − 110 to + 30 mV in 10 mV steps, followed by a 20 ms test pulse to 0 mV. Peak currents during the test pulse were normalized to the maximal current (INa/INamax, where INamax is the maximal current that is elicited at the test potential), plotted against the prepulse potential (Vm), and fitted with a Boltzmann function:
![]() |
where z1/2 and V½−inact denote the apparent gating valence and potential for half-maximal inactivation, respectively.
Recovery from fast inactivation was tested from a holding potential of − 130 mV using a two-pulse protocol with two identical 10 ms pulses to 0 mV separated by an inter-pulse interval with increasing durations from 1 to 200 ms. Current amplitudes were normalized (Pulse2/Pulse1), plotted against inter-pulse interval. The recovery time course was best fitted with a bi-exponential equation to obtain the recovery time constants τ:
![]() |
where Yt denotes the normalized current at time t, Y0 denotes the normalized current at zero time, Afast and Aslow represent the amplitudes of the fast and slow components, t is the inter-pulse interval, and τfast and τslow denote the fast and slow recovery time constants, respectively.
Statistical significance was assessed either by one-way ANOVA or by two-tailed paired Student t test. P < 0.05 was considered statistically significant. Programs used for data acquisition and analysis were pClamp 10 (Molecular Devices), Excel (Microsoft, Redmond, WA), and Prism 10 (GraphPad Software Inc., La Jolla, CA). Values are reported as mean ± SD unless otherwise stated.
Supplementary Information
Below is the link to the electronic supplementary material.
Acknowledgements
We thank Prof. Jennifer Kearney (Northwestern University) for generously providing the human Nav1.3 clone and for her valuable assistance with expression of Nav1.3 in our cell lines. We also thank Nicolas Lewis for his help with plasmid amplification.
Abbreviations
- Nav
Voltage-gated sodium channel
- SEVO
Sevoflurane
- VA
Volatile anesthetic
- AP
Action potential
- TEA-Cl
Tetraethylammonium chloride
Author contributions
KFH, HCH, and JP contributed to study conception and design. Material preparation, data collection, and analysis were performed by JX and KFH. Figures were prepared by KFH. The first draft of the manuscript was written by KH and JP and all authors reviewed previous versions of the the manuscript. All authors approved the final manuscript.
Funding
US National Institutes of Health grant R01 GM58055 to HCH, and R01 GM130722 to JP.
Data availability
The datasets generated and analyzed during the current study are included in this published article and its supplementary information files.
Declarations
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Jiaxin Xiang and Karl F. Herold contributed equally to this work.
Contributor Information
Jimcy Platholi, Email: jip2003@med.cornell.edu.
Hugh C. Hemmings, Jr., Email: hchemmi@med.cornell.edu
References
- 1.Hodgkin, A. L. & Huxley, A. F. A quantitative description of membrane current and its application to conduction and excitation in nerve 1952. Bull. Math. Biol.52, 25–71. 10.1007/bf02459568 (1990) (discussion 25-23). [DOI] [PubMed] [Google Scholar]
- 2.Catterall, W. A., Goldin, A. L. & Waxman, S. G. International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol. Rev.57, 397–409. 10.1124/pr.57.4.4 (2005). [DOI] [PubMed] [Google Scholar]
- 3.Hu, W. et al. Distinct contributions of Na(v)1.6 and Na(v)1.2 in action potential initiation and backpropagation. Nat. Neurosci.12, 996–1002. 10.1038/nn.2359 (2009). [DOI] [PubMed] [Google Scholar]
- 4.Clay, J. R. A comparative analysis of models of Na+ channel gating for mammalian and invertebrate nonmyelinated axons: Relationship to energy efficient action potentials. Prog. Biophys. Mol. Biol.111, 1–7. 10.1016/j.pbiomolbio.2012.08.005 (2013). [DOI] [PubMed] [Google Scholar]
- 5.Boiko, T. et al. Functional specialization of the axon initial segment by isoform-specific sodium channel targeting. J. Neurosci.23, 2306–2313. 10.1523/jneurosci.23-06-02306.2003 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Van Wart, A., Trimmer, J. S. & Matthews, G. Polarized distribution of ion channels within microdomains of the axon initial segment. J. Comput. Neurol.500, 339–352. 10.1002/cne.21173 (2007). [DOI] [PubMed] [Google Scholar]
- 7.Lorincz, A. & Nusser, Z. Cell-type-dependent molecular composition of the axon initial segment. J. Neurosci.28, 14329–14340. 10.1523/jneurosci.4833-08.2008 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.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. Comput. Neurol.525, 3563–3578. 10.1002/cne.24291 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Rehberg, B., Xiao, Y. H. & Duch, D. S. Central nervous system sodium channels are significantly suppressed at clinical concentrations of volatile anesthetics. Anesthesiology84, 1223–1233. 10.1097/00000542-199605000-00025 (1996) (discussion 1227A). [DOI] [PubMed] [Google Scholar]
- 10.Herold, K. F. et al. Volatile anesthetics inhibit sodium channels without altering bulk lipid bilayer properties. J. Gen. Physiol.144, 545–560. 10.1085/jgp.201411172 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Herold, K. F. & Hemmings, H. C. Jr. Sodium channels as targets for volatile anesthetics. Front. Pharmacol.3, 50. 10.3389/fphar.2012.00050 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.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.109, 2003–2011. 10.1016/j.bpj.2015.09.032 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Hodgkin, A. L. & Huxley, A. F. Currents carried by sodium and potassium ions through the membrane of the giant axon of Loligo. J. Physiol.116, 449–472. 10.1113/jphysiol.1952.sp004717 (1952). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Ratnakumari, L. & Hemmings, H. C. Jr. Inhibition of presynaptic sodium channels by halothane. Anesthesiology88, 1043–1054. 10.1097/00000542-199804000-00025 (1998). [DOI] [PubMed] [Google Scholar]
- 15.Wu, X. S., Sun, J. Y., Evers, A. S., Crowder, M. & Wu, L. G. Isoflurane inhibits transmitter release and the presynaptic action potential. Anesthesiology100, 663–670. 10.1097/00000542-200403000-00029 (2004). [DOI] [PubMed] [Google Scholar]
- 16.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.312, 801–808. 10.1124/jpet.104.074609 (2005). [DOI] [PubMed] [Google Scholar]
- 17.Ouyang, W., Wang, G. & Hemmings, H. C. Jr. Isoflurane and propofol inhibit voltage-gated sodium channels in isolated rat neurohypophysial nerve terminals. Mol. Pharmacol.64, 373–381. 10.1124/mol.64.2.373 (2003). [DOI] [PubMed] [Google Scholar]
- 18.Speigel, I. A. & Hemmings, H. C. Jr. Selective inhibition of gamma aminobutyric acid release from mouse hippocampal interneurone subtypes by the volatile anaesthetic isoflurane. Br. J. Anaesth.127, 587–599. 10.1016/j.bja.2021.06.042 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Hemmings, H. C. Jr. et al. Towards a comprehensive understanding of anesthetic mechanisms of action: A decade of discovery. Trends Pharmacol. Sci.40, 464–481. 10.1016/j.tips.2019.05.001 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Catterall, W. A. The molecular basis of neuronal excitability. Science223, 653–661. 10.1126/science.6320365 (1984). [DOI] [PubMed] [Google Scholar]
- 21.Yu, F. H. & Catterall, W. A. Overview of the voltage-gated sodium channel family. Genome Biol.4, 207. 10.1186/gb-2003-4-3-207 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Felts, P. A., Yokoyama, S., Dib-Hajj, S., Black, J. A. & Waxman, S. G. Sodium channel alpha-subunit mRNAs I, II, III, NaG, Na6 and hNE (PN1): Different expression patterns in developing rat nervous system. Brain Res. Mol. Brain Res.45, 71–82. 10.1016/s0169-328x(96)00241-0 (1997). [DOI] [PubMed] [Google Scholar]
- 23.Cummins, T. R. et al. Nav1.3 sodium channels: Rapid repriming and slow closed-state inactivation display quantitative differences after expression in a mammalian cell line and in spinal sensory neurons. J. Neurosci.21, 5952–5961. 10.1523/jneurosci.21-16-05952.2001 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Murakawa, R. & Kosaka, T. Structural features of mossy cells in the hamster dentate gyrus, with special reference to somatic thorny excrescences. J. Comput. Neurol.429, 113–126. 10.1002/1096-9861(20000101)429:1%3c113::aid-cne9%3e3.0.co;2-d (2001). [DOI] [PubMed] [Google Scholar]
- 25.Corbett, B. F. et al. Sodium channel cleavage is associated with aberrant neuronal activity and cognitive deficits in a mouse model of Alzheimer’s disease. J. Neurosci.33, 7020–7026. 10.1523/jneurosci.2325-12.2013 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Acsády, L., Kamondi, A., Sík, A., Freund, T. & Buzsáki, G. GABAergic cells are the major postsynaptic targets of mossy fibers in the rat hippocampus. J. Neurosci.18, 3386–3403. 10.1523/jneurosci.18-09-03386.1998 (1998). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Ragsdale, D. S., McPhee, J. C., Scheuer, T. & Catterall, W. A. Molecular determinants of state-dependent block of Na+ channels by local anesthetics. Science265, 1724–1728. 10.1126/science.8085162 (1994). [DOI] [PubMed] [Google Scholar]
- 28.OuYang, W. & Hemmings, H. C. Jr. Isoform-selective effects of isoflurane on voltage-gated Na+ channels. Anesthesiology107, 91–98. 10.1097/01.anes.0000268390.28362.4a (2007). [DOI] [PubMed] [Google Scholar]
- 29.Shiraishi, M. & Harris, R. A. Effects of alcohols and anesthetics on recombinant voltage-gated Na+ channels. J. Pharmacol. Exp. Ther.309, 987–994. 10.1124/jpet.103.064063 (2004). [DOI] [PubMed] [Google Scholar]
- 30.Ouyang, W., Herold, K. F. & Hemmings, H. C. Jr. Comparative effects of halogenated inhaled anesthetics on voltage-gated Na+ channel function. Anesthesiology110, 582–590. 10.1097/ALN.0b013e318197941e (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Westphalen, R. I. & Hemmings, H. C. Jr. Volatile anesthetic effects on glutamate versus GABA release from isolated rat cortical nerve terminals: 4-aminopyridine-evoked release. J. Pharmacol. Exp. Ther.316, 216–223. 10.1124/jpet.105.090662 (2006). [DOI] [PubMed] [Google Scholar]
- 32.Baumgart, J. P. et al. Isoflurane inhibits synaptic vesicle exocytosis through reduced Ca2+ influx, not Ca2+-exocytosis coupling. Proc. Natl. Acad. Sci. U. S. A.112, 11959–11964 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.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.369, 200–211. 10.1124/jpet.118.254938 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Chahine, M. et al. Sodium channel mutations in paramyotonia congenita uncouple inactivation from activation. Neuron12, 281–294. 10.1016/0896-6273(94)90271-2 (1994). [DOI] [PubMed] [Google Scholar]
- 35.Yang, N. et al. Sodium channel mutations in paramyotonia congenita exhibit similar biophysical phenotypes in vitro. Proc. Natl. Acad. Sci. U. S. A.91, 12785–12789. 10.1073/pnas.91.26.12785 (1994). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.O’Leary, M. E. Characterization of the isoform-specific differences in the gating of neuronal and muscle sodium channels. Can. J. Physiol. Pharmacol.76, 1041–1050. 10.1139/cjpp-76-10-11-1041 (1998). [DOI] [PubMed] [Google Scholar]
- 37.Waxman, S. G., Kocsis, J. D. & Black, J. A. Type III sodium channel mRNA is expressed in embryonic but not adult spinal sensory neurons, and is reexpressed following axotomy. J. Neurophysiol.72, 466–470. 10.1152/jn.1994.72.1.466 (1994). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Lewis, A. H. & Raman, I. M. Resurgent current of voltage-gated Na(+) channels. J. Physiol.592, 4825–4838. 10.1113/jphysiol.2014.277582 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Stafstrom, C. E. Persistent sodium current and its role in epilepsy. Epilepsy Curr.7, 15–22. 10.1111/j.1535-7511.2007.00156.x (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Zhao, W. et al. Isoflurane modulates hippocampal cornu ammonis pyramidal neuron excitability by inhibition of both transient and persistent sodium currents in mice. Anesthesiology131, 94–104. 10.1097/aln.0000000000002753 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Lundbaek, J. A. et al. Regulation of sodium channel function by bilayer elasticity: The importance of hydrophobic coupling. Effects of Micelle-forming amphiphiles and cholesterol. J. Gen. Physiol.123, 599–621. 10.1085/jgp.200308996 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Rusinova, R. et al. Thiazolidinedione insulin sensitizers alter lipid bilayer properties and voltage-dependent sodium channel function: Implications for drug discovery. J. Gen. Physiol.138, 249–270. 10.1085/jgp.201010529 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Gingrich, K. J., Burkat, P. M. & Roberts, W. A. Pentobarbital produces activation and block of {alpha}1{beta}2{gamma}2S GABAA receptors in rapidly perfused whole cells and membrane patches: divergent results can be explained by pharmacokinetics. J. Gen. Physiol.133, 171–188. 10.1085/jgp.200810081 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Baker, M. D., Chandra, S. Y., Ding, Y., Waxman, S. G. & Wood, J. N. GTP-induced tetrodotoxin-resistant Na+ current regulates excitability in mouse and rat small diameter sensory neurones. J. Physiol.548, 373–382. 10.1113/jphysiol.2003.039131 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Lampert, A., Hains, B. C. & Waxman, S. G. Upregulation of persistent and ramp sodium current in dorsal horn neurons after spinal cord injury. Exp. Brain Res.174, 660–666. 10.1007/s00221-006-0511-x (2006). [DOI] [PubMed] [Google Scholar]
- 46.Zhou, H., Xie, Z., Brambrink, A. M. & Yang, G. Behavioural impairments after exposure of neonatal mice to propofol are accompanied by reductions in neuronal activity in cortical circuitry. Br. J. Anaesth.126, 1141–1156. 10.1016/j.bja.2021.01.017 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Yang, L. et al. Sevoflurane induces neuronal activation and behavioral hyperactivity in young mice. Sci. Rep.10, 11226. 10.1038/s41598-020-66959-x (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Briner, A. et al. Developmental Stage-dependent persistent impact of propofol anesthesia on dendritic spines in the rat medial prefrontal cortex. Anesthesiology115, 282–293. 10.1097/ALN.0b013e318221fbbd (2011). [DOI] [PubMed] [Google Scholar]
- 49.De Roo, M. et al. Anesthetics rapidly promote synaptogenesis during a critical period of brain development. PLoS ONE4, e7043. 10.1371/journal.pone.0007043 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Huang, W. C. et al. Lateral mammillary body neurons in mouse brain are disproportionately vulnerable in Alzheimer’s disease. Sci. Transl. Med.15, eabq1019. 10.1126/scitranslmed.abq1019 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Vutskits, L., Gascon, E., Tassonyi, E. & Kiss, J. Z. Effect of ketamine on dendritic arbor development and survival of immature GABAergic neurons in vitro. Toxicol. Sci.91, 540–549. 10.1093/toxsci/kfj180 (2006). [DOI] [PubMed] [Google Scholar]
- 52.Vutskits, L., Gascon, E., Potter, G., Tassonyi, E. & Kiss, J. Z. Low concentrations of ketamine initiate dendritic atrophy of differentiated GABAergic neurons in culture. Toxicology234, 216–226. 10.1016/j.tox.2007.03.004 (2007). [DOI] [PubMed] [Google Scholar]
- 53.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. Anesthesiology83, 109–119. 10.1097/00000542-199507000-00014 (1995). [DOI] [PubMed] [Google Scholar]
- 54.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.88, 676–680. 10.1097/00000539-199903000-00039 (1999). [DOI] [PubMed] [Google Scholar]
- 55.MacIver, M. B. Anesthetic agent-specific effects on synaptic inhibition. Anesth. Analg.119, 558–569. 10.1213/ane.0000000000000321 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Lindia, J. A., Köhler, M. G., Martin, W. J. & Abbadie, C. Relationship between sodium channel NaV1.3 expression and neuropathic pain behavior in rats. Pain117, 145–153. 10.1016/j.pain.2005.05.027 (2005). [DOI] [PubMed] [Google Scholar]
- 57.Isom, L. L. et al. Functional co-expression of the beta 1 and type IIA alpha subunits of sodium channels in a mammalian cell line. J. Biol. Chem.270, 3306–3312. 10.1074/jbc.270.7.3306 (1995). [DOI] [PubMed] [Google Scholar]
- 58.Kazen-Gillespie, K. A. et al. Cloning, localization, and functional expression of sodium channel beta1A subunits. J. Biol. Chem.275, 1079–1088. 10.1074/jbc.275.2.1079 (2000). [DOI] [PubMed] [Google Scholar]
- 59.Isom, L. L. et al. Primary structure and functional expression of the beta 1 subunit of the rat brain sodium channel. Science256, 839–842. 10.1126/science.1375395 (1992). [DOI] [PubMed] [Google Scholar]
- 60.Meadows, L. S., Chen, Y. H., Powell, A. J., Clare, J. J. & Ragsdale, D. S. Functional modulation of human brain Nav1.3 sodium channels, expressed in mammalian cells, by auxiliary beta 1, beta 2 and beta 3 subunits. Neuroscience114, 745–753. 10.1016/s0306-4522(02)00242-7 (2002). [DOI] [PubMed] [Google Scholar]
- 61.McEwen, D. P., Meadows, L. S., Chen, C., Thyagarajan, V. & Isom, L. L. Sodium channel beta1 subunit-mediated modulation of Nav1.2 currents and cell surface density is dependent on interactions with contactin and ankyrin. J. Biol. Chem.279, 16044–16049. 10.1074/jbc.M400856200 (2004). [DOI] [PubMed] [Google Scholar]
- 62.McCormick, K. A. et al. Molecular determinants of Na+ channel function in the extracellular domain of the beta1 subunit. J. Biol. Chem.273, 3954–3962. 10.1074/jbc.273.7.3954 (1998). [DOI] [PubMed] [Google Scholar]
- 63.Estacion, M. & Waxman, S. G. The response of Na(V)1.3 sodium channels to ramp stimuli: multiple components and mechanisms. J. Neurophysiol.109, 306–314. 10.1152/jn.00438.2012 (2013). [DOI] [PubMed] [Google Scholar]
- 64.Vanoye, C. G., Gurnett, C. A., Holland, K. D., George, A. L. Jr. & Kearney, J. A. Novel SCN3A variants associated with focal epilepsy in children. Neurobiol. Dis.62, 313–322. 10.1016/j.nbd.2013.10.015 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Köster, P. A. et al. Nociceptor sodium channels shape subthreshold phase, upstroke, and shoulder of action potentials. J. Gen. Physiol.10.1085/jgp.202313526 (2025). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Chen, C. et al. Reduced sodium channel density, altered voltage dependence of inactivation, and increased susceptibility to seizures in mice lacking sodium channel beta 2-subunits. Proc. Natl. Acad. Sci. U. S. A.99, 17072–17077. 10.1073/pnas.212638099 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.O’Malley, H. A. & Isom, L. L. Sodium channel β subunits: Emerging targets in channelopathies. Annu. Rev. Physiol.77, 481–504. 10.1146/annurev-physiol-021014-071846 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Raju, S. G., Barber, A. F., LeBard, D. N., Klein, M. L. & Carnevale, V. Exploring volatile general anesthetic binding to a closed membrane-bound bacterial voltage-gated sodium channel via computation. PLoS Comput. Biol.9, e1003090. 10.1371/journal.pcbi.1003090 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Spurny, R. et al. Multisite binding of a general anesthetic to the prokaryotic pentameric Erwinia chrysanthemi ligand-gated ion channel (ELIC). J. Biol. Chem.288, 8355–8364. 10.1074/jbc.M112.424507 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Sand, R. M., Gingrich, K. J., Macharadze, T., Herold, K. F. & Hemmings, H. C. Jr. Isoflurane modulates activation and inactivation gating of the prokaryotic Na(+) channel NaChBac. J. Gen. Physiol.149, 623–638. 10.1085/jgp.201611600 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Hollingworth, D. et al. Functionally important binding site for a volatile anesthetic in a voltage-gated sodium channel identified by X-ray crystallography. 10.1101/2024.11.04.621342 (2024).
- 72.Hemmings, H. C. Jr. Sodium channels and the synaptic mechanisms of inhaled anaesthetics. Br. J. Anaesth.103, 61–69. 10.1093/bja/aep144 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Qiu, J. et al. The volatile anesthetic isoflurane differentially inhibits voltage-gated sodium channel currents between pyramidal and parvalbumin neurons in the prefrontal cortex. Front Neural Circuits17, 1185095. 10.3389/fncir.2023.1185095 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Li, T. et al. Action potential initiation in neocortical inhibitory interneurons. PLoS Biol.12, e1001944. 10.1371/journal.pbio.1001944 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Ogiwara, I. et al. 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.27, 5903–5914. 10.1523/jneurosci.5270-06.2007 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Wang, J., Ou, S. W. & Wang, Y. J. Distribution and function of voltage-gated sodium channels in the nervous system. Channels (Austin)11, 534–554. 10.1080/19336950.2017.1380758 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Ouyang, W. & Hemmings, H. Isoform-selective effects of isoflurane on voltage-gated Na+ channels. Anesthesiology107, 91–98. 10.1097/01.anes.0000268390.28362.4a00000542-200707000-00017[pii] (2007). [DOI] [PubMed] [Google Scholar]
- 78.Ouyang, W., Herold, K. F. & Hemmings, H. C. Comparative effects of halogenated inhaled anesthetics on voltage-gated Na+ channel function. Anesthesiology110, 582–590. 10.1097/ALN.0b013e318197941e (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Ouyang, W., Jih, T., Zhang, T., Correa, A. & Hemmings, H. Isoflurane inhibits NaChBac, a prokaryotic voltage-gated sodium channel. J. Pharmacol. Exp. Ther.322, 1076–1083. 10.1124/jpet.107.122929 (2007). [DOI] [PubMed] [Google Scholar]
- 80.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.115, 112–121. 10.1093/bja/aev203 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Barber, A. F., Carnevale, V., Klein, M. L., Eckenhoff, R. G. & Covarrubias, M. Modulation of a voltage-gated Na+ channel by sevoflurane involves multiple sites and distinct mechanisms. Proc. Natl. Acad. Sci. U. S. A.111, 6726–6731. 10.1073/pnas.1405768111 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Zhang, Y. et al. Bidirectional modulation of isoflurane potency by intrathecal tetrodotoxin and veratridine in rats. Br. J. Pharmacol.159, 872–878. 10.1111/j.1476-5381.2009.00583.x (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Zhang, Y. et al. Intrathecal veratridine administration increases minimum alveolar concentration in rats. Anesth. Analg.107, 875–878. 10.1213/ane.0b013e3181815fbc (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Denomme, N., Hull, J. M. & Mashour, G. A. Role of voltage-gated sodium channels in the mechanism of ether-induced unconsciousness. Pharmacol. Rev.71, 450–466. 10.1124/pr.118.016592 (2019). [DOI] [PubMed] [Google Scholar]
- 85.Eger, E. I. Characteristics of anesthetic agents used for induction and maintenance of general anesthesia. Am. J. Health Syst. Pharm.61(Suppl 4), S3-10. 10.1093/ajhp/61.suppl_4.S3 (2004). [DOI] [PubMed] [Google Scholar]
- 86.Frink, E. J. Jr. et al. Clinical comparison of sevoflurane and isoflurane in healthy patients. Anesth Analg74, 241–245. 10.1213/00000539-199202000-00012 (1992). [DOI] [PubMed] [Google Scholar]
- 87.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 ONE10, e0134960. 10.1371/journal.pone.0134960 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Bezanilla, F. & Armstrong, C. M. Inactivation of the sodium channel. I. Sodium current experiments. J. Gen. Physiol.70, 549–566. 10.1085/jgp.70.5.549 (1977). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Herold, K. F., Nau, C., Ouyang, W. & Hemmings, H. C. Jr. Isoflurane inhibits the tetrodotoxin-resistant voltage-gated sodium channel Nav1.8. Anesthesiology111, 591–599. 10.1097/ALN.0b013e3181af64d4 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Taheri, S. et al. What solvent best represents the site of action of inhaled anesthetics in humans, rats, and dogs?. Anesth. Analg.72, 627–634. 10.1213/00000539-199105000-00010 (1991). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The datasets generated and analyzed during the current study are included in this published article and its supplementary information files.










