Fact-checked by Grok 2 weeks ago

Action potential

An action potential is a rapid, transient reversal in the electrical potential across the plasma of an excitable , such as a or , enabling the transmission of signals over long distances without decrement. This electrochemical event, typically lasting about 1 millisecond, is characterized by a sequence of phases driven by the selective permeability of the to ions like sodium (Na⁺) and potassium (K⁺). In , action potentials originate at the axon hillock when the depolarizes to a of approximately -55 mV, triggering voltage-gated channels. The process begins with , where voltage-gated Na⁺ channels open, allowing Na⁺ influx that rapidly shifts the from a resting value of about -70 mV toward +30 mV or higher, creating the characteristic spike. This is followed by , as Na⁺ channels inactivate and voltage-gated K⁺ channels open, permitting K⁺ efflux that restores the negative . A brief hyperpolarization (afterhyperpolarization) often occurs due to delayed closure of K⁺ channels, temporarily making the membrane more negative than rest, which contributes to the refractory period that ensures unidirectional . The all-or-none nature of action potentials means they fire fully once threshold is reached, independent of stimulus strength, a principle first quantitatively modeled by Hodgkin and Huxley in their seminal 1952 study on squid giant axons. Action potentials are fundamental to physiological processes, serving as the primary mechanism for intercellular communication in the , where they propagate along axons at speeds up to 120 m/s in myelinated fibers via . In cardiac and , they trigger contractions by coupling to calcium release and excitation-contraction mechanisms. Disruptions in action potential generation or propagation underlie disorders like , arrhythmias, and channelopathies, highlighting their critical role in health and disease.

Fundamentals

Definition and overview

An action potential is a transient, rapid reversal in the of excitable cells, such as neurons and muscle cells, from a negative resting value to a positive peak, driven by selective fluxes of ions across the . This electrical event serves as the fundamental unit of in these cells. The concept of the action potential developed in the early through pioneering electrophysiological studies, with Keith Lucas demonstrating in 1909 that the electrical response in follows an all-or-none principle, laying groundwork for understanding its discrete nature. Action potentials play a critical role in biological communication: they propagate signals along neurons to enable information processing in the , trigger by linking motor neurons to myofibers, and facilitate sensory by converting physical or chemical stimuli into neural codes. In a typical action potential , the process starts with , a steep rise in membrane potential from the resting state (around -70 mV) to a positive overshoot (near +30 mV), followed by , a quick return toward the resting level, and often concluding with a short hyperpolarization phase that temporarily dips below baseline before stabilizing. This characteristic shape, lasting about 1-2 milliseconds in neurons, underscores the event's speed and precision.

Resting membrane potential

The resting membrane potential refers to the stable electrical voltage across the of excitable cells, such as neurons, in their non-stimulated state, typically measuring approximately -70 inside relative to outside. This negative potential arises from the unequal distribution of ions across the and the selective permeability of the to those ions. The ion concentration gradients essential for this potential are actively maintained by the sodium-potassium pump, an electrogenic transmembrane that hydrolyzes ATP to transport three sodium ions (Na⁺) out of the and two potassium ions (K⁺) into the per cycle. Discovered in crab nerve membranes, this pump counters the passive leakage of ions, establishing higher intracellular K⁺ concentrations (around 140 mM) and higher extracellular Na⁺ concentrations (around 145 mM), while also contributing a small direct electrogenic effect due to the net export of positive charge. At rest, the membrane exhibits high selective permeability to K⁺ through constitutively open leak channels, allowing K⁺ to diffuse out down its concentration gradient and thereby establishing a potential close to the K⁺ value. Permeability to Na⁺ is much lower (about 5% of K⁺ permeability), limiting Na⁺ influx and preventing significant depolarization. The potential for each , representing the voltage at which its is balanced, is calculated using the : E_{\text{ion}} = \frac{RT}{zF} \ln \left( \frac{[\text{ion}]_{\text{out}}}{[\text{ion}]_{\text{in}}} \right) where R is the gas constant, T is temperature in Kelvin, z is the ion's valence, and F is Faraday's constant. For typical neuronal concentrations, this yields an equilibrium potential of approximately +60 mV for Na⁺ and -90 mV for K⁺. The actual resting potential of -70 mV reflects a permeability-weighted average, dominated by K⁺ conductance as described in early squid axon models. Chloride ions (Cl⁻) and other anions play a supporting role in fine-tuning the , with moderate membrane permeability allowing Cl⁻ influx that adds to intracellular negativity, as its potential aligns near -70 due to higher extracellular concentrations (around 110 ). Impermeable intracellular anions, such as proteins, further contribute to the negative charge balance without direct flux.

Ion channels and pumps

Voltage-gated sodium channels (Nav) are integral membrane proteins essential for the initiation of action potentials in excitable cells. Each Nav channel consists of a single α-subunit forming the pore, comprising four homologous domains (I-IV), each with six transmembrane segments (S1-S6), and often associated with one or more β-subunits that modulate function. The voltage-sensing domain, formed by the S4 segment with positively charged residues, detects depolarization and triggers conformational changes that open the channel pore, allowing rapid Na⁺ influx. Inactivation occurs via a "ball-and-chain" mechanism involving the intracellular loop between domains III and IV, which plugs the pore shortly after activation to terminate Na⁺ flow. Voltage-gated potassium channels (Kv), particularly the delayed rectifier subtypes, play a crucial role in repolarizing the following . These channels are tetramers of α-subunits, each with six transmembrane segments, where the S5-S6 regions form the central -selective , and the S4 segment acts as the voltage sensor. Delayed rectifier Kv channels, such as family members, activate more slowly than channels in response to , permitting efflux that restores the negative . Their structure ensures high selectivity through a conserved in the pore loop, enabling rapid permeation near the limit. The Na⁺/K⁺-ATPase, an electrogenic pump, maintains ionic gradients by actively transporting 3 Na⁺ s out of the and 2 K⁺ s inward per ATP hydrolyzed, generating a net outward current that contributes to the resting . This consists of α, β, and γ subunits; the α-subunit contains the catalytic site and ion-binding domains, undergoing E1-E2 conformational cycles to alternately bind Na⁺ intracellularly and K⁺ extracellularly. The pump's energy requirement is approximately 1 ATP per cycle, consuming a significant portion of cellular energy in neurons to counteract passive ion leaks and support excitability. Leak channels, primarily from the two-pore (K2P) , provide a constitutive K⁺ conductance that stabilizes the resting near the K⁺ equilibrium potential. These dimeric channels, each subunit with four transmembrane segments and two pore loops forming a single wide pore, lack voltage sensitivity and are modulated by factors like , , and . In neurons, K2P channels such as TREK-1 and TASK-1 dominate resting K⁺ permeability, contributing up to 50-70% of the total conductance and preventing excessive . Voltage-gated calcium channels (Cav), though less prominent in the initial action potential phases in many neurons, support excitability by allowing Ca²⁺ entry during prolonged or high-frequency firing. In neuronal contexts, high-voltage-activated Cav1 and Cav2 subtypes, structured similarly to with four domains and voltage-sensing S4 segments, couple to Ca²⁺ influx that modulates release and .

Mechanism in Neurons

Initiation of action potentials

Action potentials in typically initiate at the axon hillock or the adjacent axon initial segment (AIS), regions characterized by a high density of voltage-gated sodium (Na⁺) channels, which lowers the threshold for compared to the or dendrites. This specialized integration site allows the to summate incoming signals efficiently, with Na⁺ channel densities often several times higher than in the (e.g., 3-16 fold in some models), facilitating rapid upon sufficient excitatory input. Initiation requires a stimulus that depolarizes the to the , approximately -55 mV, at which point the influx of Na⁺ through opening voltage-gated channels exceeds the outward leak of positive ions, creating a loop. Stimuli can be sensory, such as mechanical or chemical in peripheral neurons; synaptic, involving release that generates excitatory postsynaptic potentials (EPSPs); or artificial, like depolarizing current injection in electrophysiological experiments. In central neurons, synaptic inputs predominate, where multiple EPSPs from presynaptic neurons summate temporally (overlapping in time from the same ) or spatially (from different synapses) at the hillock to reach threshold. Concurrent inhibitory postsynaptic potentials (IPSPs), which hyperpolarize the , can counteract this summation, preventing initiation unless excitatory drive overcomes inhibition. The ease of action potential initiation is further modulated by the neuron's refractory state following a prior action potential. During the absolute refractory period, lasting about 1-2 ms, Na⁺ channels are inactivated, rendering the inexcitable regardless of stimulus strength. In contrast, the relative refractory period, which follows and can extend several milliseconds, involves partial hyperpolarization and some Na⁺ channel recovery, allowing but only with a suprathreshold stimulus stronger than normal due to reduced excitability. These periods ensure unidirectional propagation and limit firing rates, with the relative phase particularly influencing the timing and probability of subsequent initiations in response to ongoing synaptic barrage.

Phases of the action potential

The action potential in neurons unfolds as a series of distinct voltage phases driven by selective fluxes across the , primarily involving sodium (Na⁺) and potassium (K⁺) s through voltage-gated channels. This temporal sequence begins at the resting of approximately -70 mV and results in a stereotypical lasting about 1 ms in total. The phases are characterized by rapid changes in membrane conductance, as modeled quantitatively by Hodgkin and Huxley based on voltage-clamp experiments in giant axons. The , or depolarization, occurs when the membrane potential reaches threshold and voltage-gated Na⁺ channels open, allowing rapid Na⁺ influx that shifts the voltage from -70 mV to +30 mV within about 1 ms. This positive feedback loop amplifies the initial due to the electrochemical driving force favoring Na⁺ entry. At the , the membrane potential briefly plateaus at +30 to +40 mV, where Na⁺ conductance reaches its maximum before channels begin to inactivate, temporarily balancing inward Na⁺ current with emerging outward currents. This phase lasts less than 1 ms and marks the reversal of the net current direction. The falling phase, or , follows as Na⁺ channels inactivate and voltage-gated K⁺ channels—known as delayed rectifiers—open, permitting K⁺ efflux that restores the toward -70 mV over 1-2 ms. The delayed activation of these K⁺ channels ensures that occurs after the peak, preventing premature termination of the action potential. The , or undershoot, ensues when K⁺ conductance remains elevated longer than necessary, causing the potential to overshoot to about -80 mV before gradually returning to rest. This , lasting a few milliseconds, arises from the slower deactivation of K⁺ channels and contributes to spacing successive action potentials. In the Hodgkin-Huxley model, these phases are governed by probabilistic variables for channels: for ⁺ channels, the m controls fast (opening during rising ) and the h gate mediates inactivation (closing during falling ); for K⁺ channels, the n handles slower (delayed behavior during and afterhyperpolarization). These gates follow voltage- and time-dependent rate functions derived from experimental data, enabling predictive simulations of the action potential waveform.

All-or-none principle and refractory periods

The all-or-none principle states that once the for is reached in a , the action potential is generated with a fixed and , independent of the stimulus strength beyond that threshold; subthreshold stimuli produce no action potential, while suprathreshold stimuli elicit the full response without gradation. This nature ensures reliable signal transmission along the , as demonstrated in early experiments on isolated fibers where responses were either maximal or absent. Following an action potential, the enters an absolute refractory period during which no new action potential can be initiated, regardless of stimulus intensity, due to the inactivation of voltage-gated s that prevents further . This period typically lasts 1-2 ms in mammalian s and corresponds to the rising and falling phases of the action potential. The absolute refractory period is followed by a relative refractory period, where a new action potential can be elicited but only by a stronger-than-normal stimulus, as the membrane is hyperpolarized below the due to lingering potassium efflux, raising the for excitation. This phase generally endures 2-4 ms, allowing partial recovery of availability while the remains more negative than rest. These periods serve critical functional roles: the phase enforces unidirectional by preventing backward firing into recently activated segments, while both periods cap the maximum firing , typically to around 500 Hz in neurons, ensuring discrete signaling without overlap. Although action potentials adhere strictly to the all-or-none rule along axons, exceptions occur in graded potentials within dendrites and the cell body, where changes vary proportionally with stimulus intensity and can summate without limitations.

Propagation and Conduction

Continuous conduction

Continuous conduction refers to the mechanism by which action potentials propagate along unmyelinated through a continuous, wave-like spread without interruptions. In this process, the initial at one site on the triggers local currents, where ions flow intracellularly along the and extracellularly, creating an electrotonic spread of to adjacent segments. This passive electrotonic brings nearby regions of the to , prompting active regeneration of the action potential at each point via the sequential opening of voltage-gated sodium channels, which influx Na⁺ and further depolarize the . The voltage-gated channels ensure that the signal is amplified and regenerated continuously along the entire length of the , maintaining the action potential's amplitude and shape. The speed of continuous conduction in unmyelinated typically ranges from 0.5 to 10 m/s, depending on various physiological factors. Conduction velocity increases with axon because larger diameters reduce the axoplasmic , allowing more efficient flow of local currents and faster of adjacent segments. For example, thinner unmyelinated fibers, such as those in C-class nociceptive neurons with diameters around 0.2–1.2 μm, conduct at slower speeds near 1 m/s, while slightly larger ones achieve higher velocities within the typical range. This mode of incurs a significant energy cost due to the continuous influx of Na⁺ through voltage-gated channels along the entire length, which must be counteracted by the Na⁺/K⁺-ATPase pump to restore ionic gradients. Each action potential requires by the pump to extrude three Na⁺ ions and import two K⁺ ions, with the energy demand scaling with the frequency and extent of ; in unmyelinated axons, this can consume a substantial portion of the neuron's metabolic resources compared to more efficient myelinated conduction.

Saltatory conduction and myelin

Myelin sheaths are multilayered lipid-rich membranes that insulate , enabling efficient action potential propagation. In the peripheral nervous system (PNS), Schwann cells form these sheaths by spirally wrapping their plasma membrane around a single segment, producing concentric lamellae rich in and proteins such as myelin basic protein and peripheral myelin protein 22. In the (CNS), generate through a similar wrapping process but extend processes to myelinate up to 50 different , creating compact sheaths with approximately 12 nm periodicity between layers. These structures reduce axonal and prevent extracellular leakage, concentrating electrical current flow. Nodes of Ranvier are short, unmyelinated gaps, typically 0.8–1.1 μm in length, that interrupt the myelin sheath at regular intervals of 0.2–2 mm depending on diameter. These nodes feature a high density of voltage-gated sodium channels—up to 2000 per μm²—clustered via interactions with axoglial adhesion molecules like and neurofascin. The paranodal regions flanking the nodes form septate-like junctions that seal the periaxonal space, further insulating the internodal segments. Saltatory conduction occurs when action potentials regenerate actively only at the nodes of Ranvier, while the depolarizing current spreads passively and rapidly through the insulated internodal regions via electrotonic conduction. This "jumping" mechanism, first demonstrated experimentally in sciatic nerves, relies on the low and high axial of myelinated internodes to advance the signal temporally ahead of the nodal . Unlike continuous conduction along unmyelinated axons, saltatory propagation minimizes the spatial extent of flux by limiting activation to discrete nodal sites. This process dramatically increases conduction velocity to 70–150 m/s in large myelinated axons, compared to 0.5–10 m/s in unmyelinated fibers, while also conserving energy by reducing the number of ion channels that must cycle sodium and potassium. Optimal myelin thickness, around 16 lamellae, further maximizes speed by balancing insulation with current flow through the periaxonal space, a narrow (~12 nm) conductive layer between the axon and myelin. Demyelination disrupts this efficiency, as seen in where immune-mediated loss of sheaths exposes internodes, leading to current leakage, slowed conduction, or complete block at affected nodes. This results in neurological symptoms such as , sensory deficits, and , with conduction failure exacerbated by repetitive firing or temperature increases due to impaired redistribution. In chronic lesions, persistent axonal vulnerability can progress to degeneration, underscoring 's role in long-term neuronal integrity.

Cable theory basics

Cable theory models neuronal processes, such as and dendrites, as electrical cables to describe the passive spread of voltage signals along their length. In this framework, the is treated as a cylindrical core conductor surrounded by a , with key electrical parameters including the intracellular (axial) R_i (in \Omega \cdot \text{cm}), the extracellular resistance R_o (often negligible compared to R_i), the C_m (in \mu\text{F/cm}^2), the R_m (in \Omega \cdot \text{cm}^2), and the radius a (in cm). These parameters determine how voltage changes propagate without active regeneration, analogous to current flow in a leaky . The , denoted \lambda, quantifies the distance over which a steady-state voltage signal decays to $1/e (approximately 37%) of its initial value along the . It is given by the formula \lambda = \sqrt{\frac{a R_m}{2 R_i}}, where higher R_m (less leaky ), larger a (thicker ), or lower R_i (more conductive axoplasm) increases \lambda, allowing signals to spread farther. Typical values in large neuronal axons yield \lambda on the order of 1–5 mm, illustrating the scale of passive decay in neuronal . The \tau governs the speed of charging or discharging in response to injection, defined as \tau = R_m \cdot C_m. This parameter, typically 1–10 in neurons, reflects how quickly subthreshold potentials equilibrate temporally, with larger \tau indicating slower responses due to higher or . Electrotonic conduction refers to this passive, non-regenerative spread of subthreshold voltage changes, where the potential V(x) at distance x from the injection site decays exponentially as V(x) = V_0 e^{-x/\lambda}. Without active mechanisms, signals attenuate rapidly beyond a few length constants, limiting effective communication over long distances. In dendrites, underpins the integration of synaptic inputs, where subthreshold signals from multiple synapses summate spatially and temporally before reaching the . This passive filtering allows neurons to compute weighted averages of inputs, with dendritic modulating the strength and timing of potentials, as first rigorously analyzed in branched structures.

Synaptic Transmission and Termination

Chemical synapses

Upon arrival of an action potential at the presynaptic , depolarization opens voltage-gated calcium channels, allowing Ca²⁺ influx that triggers the fusion of synaptic vesicles with the presynaptic . This rapid Ca²⁺ entry, raising cytosolic concentrations from nanomolar to 50-100 μM near the channels, binds to synaptotagmin on vesicle membranes, promoting SNARE complex assembly and within 0.5-1 millisecond. The vesicles, typically 40-50 nm in diameter, release their contents into the synaptic cleft through transient fusion pores. The synaptic cleft, a narrow measuring 20-40 nm wide, enables rapid of released neurotransmitters to the postsynaptic . Neurotransmitters such as glutamate, the primary excitatory transmitter in the , and , the main inhibitory transmitter, traverse this gap in approximately 0.1 milliseconds due to their small molecular size and concentration gradients. This occurs without direct cytoplasmic continuity between cells, ensuring unidirectional signaling modulated by or enzymatic degradation. On the postsynaptic side, neurotransmitters bind to specialized receptors, eliciting either fast or slow responses. Ionotropic receptors, such as and NMDA subtypes for glutamate or receptors for , are ligand-gated ion channels that directly permit ion flow (e.g., Na⁺, K⁺, or Cl⁻), generating excitatory or inhibitory postsynaptic potentials within milliseconds. In contrast, metabotropic receptors, coupled to G-proteins (e.g., mGluR for glutamate or ), activate intracellular signaling cascades like second messenger production, leading to slower, modulatory effects lasting seconds to minutes. Synaptic transmission is quantal, with each vesicle releasing a packet (quantum) of , typically 1,000-10,000 molecules, producing a unitary postsynaptic response. Spontaneous release of single quanta generates miniature excitatory postsynaptic potentials (mEPSPs) or inhibitory potentials (mIPSPs), averaging 0.5-1 in amplitude, as first demonstrated at the . This quantal framework, established by Fatt and Katz, underlies the probabilistic nature of synaptic efficacy and plasticity.

Electrical synapses

Electrical synapses provide direct electrical coupling between neurons through s, enabling the rapid passage of ions and small molecules without the involvement of neurotransmitters. These junctions are formed by proteins, which assemble into hexameric structures known as connexons; each consists of two opposing connexons, one from each cell, creating a with a pore diameter of approximately 1.5 nm that allows bidirectional flow of current and metabolites. This bidirectional current flow facilitates the of action potentials across coupled cells, as seen in inhibitory interneurons of the , where electrical coupling promotes coordinated firing to regulate network activity.00373-7) Unlike chemical synapses, which rely on slower release and receptor activation, electrical synapses enable nearly instantaneous transmission with delays under 0.1 ms, supporting rapid synchronization in circuits requiring precise timing. Electrical synapses are prevalent in , where they form extensive networks for signal propagation, and in early development, aiding circuit maturation before chemical synapses dominate. In adult vertebrates, they persist in specific regions like the , connecting bipolar and amacrine cells to enhance visual processing through synchronized responses. Beyond electrical signaling, gap junctions support metabolic coupling by permitting the diffusion of small metabolites and second messengers, which can coordinate cellular in coupled populations. While advantageous for speed and synchrony, electrical synapses offer less cellular isolation than chemical synapses, as their direct connectivity can propagate unwanted activity, potentially contributing to epileptiform spread in pathological conditions like through enhanced hypersynchronization. This reduced modifiability limits fine-tuned control but ensures reliable, low-latency communication in specialized neural circuits.

Neuromuscular junctions

The (NMJ) represents a specialized between the of a and the of a fiber, facilitating the transmission of action potentials to initiate . This interface ensures high-fidelity signaling, distinct from synapses by its robust quantal release and structural adaptations for reliable neuromuscular activation. Structurally, the NMJ features a presynaptic motor terminal apposed to the motor end plate, a specialized region of the muscle fiber's characterized by extensive infoldings known as junctional folds. These folds increase the surface area and are densely packed with nicotinic receptors (nAChRs), reaching densities of approximately 10,000 per square micrometer, which amplifies the postsynaptic response to binding. The synaptic cleft between the presynaptic terminal and the folded measures about 50 nanometers, optimizing and receptor interaction. Upon arrival of an action potential at the presynaptic terminal, voltage-gated calcium channels open, triggering the synchronous of synaptic vesicles containing (ACh), the primary at the NMJ. This release occurs in large quanta, typically involving 50-300 vesicles per action potential in mammalian NMJs, each vesicle releasing around 10,000 ACh molecules into the synaptic cleft. The quantal nature of this release, first elucidated by and colleagues, ensures a probabilistic yet robust summation of miniature end-plate potentials into a full . The released binds to postsynaptic nAChRs, opening ligand-gated cation channels that permit sodium influx, generating an (EPP) that depolarizes the motor end plate. This , typically 40-50 mV in amplitude, exceeds the required to activate voltage-gated sodium channels in the adjacent , thereby triggering a propagating muscle action potential and . A key feature of NMJ transmission is the safety factor, defined as the ratio of EPP amplitude to the threshold for muscle action potential initiation (often around 3-5), which provides redundancy against fluctuations in transmitter release or receptor function, ensuring near-100% transmission fidelity under normal conditions. To terminate signaling and prevent desensitization, (AChE), anchored in the synaptic cleft and , rapidly hydrolyzes ACh into choline and acetate. This enzymatic action occurs with extraordinary efficiency, hydrolyzing unbound or dissociated ACh molecules in less than 1 , thereby limiting the duration of the EPP to about 5-10 milliseconds and allowing for high-frequency muscle activation. Disruptions at the NMJ underlie several disorders affecting transmission. is an autoimmune condition where antibodies target postsynaptic nAChRs, reducing receptor density and impairing EPP generation, leading to fatigable muscle weakness. , produced by , cleaves SNARE proteins in the presynaptic terminal, blocking ACh vesicle and causing by preventing EPP formation.

Variations Across Cell Types

Cardiac action potentials

Cardiac action potentials differ from those in neurons primarily due to their prolonged duration and characteristic plateau phase, which enable synchronized of heart muscle for effective pumping. These potentials occur in various cardiac cell types, including pacemaker cells in the sinoatrial () node and contractile myocytes in the atria and ventricles, with ion fluxes tightly regulated to maintain rhythmic activity. The cardiac action potential consists of five distinct phases. Phase 0 involves rapid depolarization driven by influx of sodium ions (Na⁺) through voltage-gated Na⁺ channels (primarily Nav1.5), shifting the membrane potential from approximately -90 mV to +20 mV. This upstroke is fast in ventricular myocytes but slower in SA node cells, where calcium (Ca²⁺) currents contribute more prominently due to fewer Na⁺ channels. Phase 1 features early repolarization from potassium (K⁺) efflux via transient outward K⁺ channels (I_to) and inactivation of Na⁺ channels. Phase 2, the plateau, is maintained by a balance of inward L-type Ca²⁺ currents through Cav1.2 channels and outward K⁺ currents, lasting 200-300 ms to allow sufficient time for excitation-contraction coupling. Phase 3 completes repolarization through delayed rectifier K⁺ currents (I_Kr and I_Ks, mediated by channels like KCNH2 and KCNQ1), restoring the membrane to its resting state. Phase 4 represents the resting potential, stabilized by inward rectifier K⁺ currents (I_K1); in pacemaker cells like those in the SA node, it includes spontaneous diastolic depolarization via the funny current (I_f through HCN channels), initiating the next action potential. Ion dynamics are tailored to cardiac function, with L-type Ca²⁺ channels playing a central role in the plateau phase to link electrical signaling to mechanical contraction, unlike the briefer neuronal action potentials dominated by Na⁺ and K⁺ fluxes. In SA node cells, the action potential upstroke is slower (conduction velocity ~0.1-0.2 m/s) and relies more on Ca²⁺ influx via Cav1.3 channels, reflecting their role in without a stable around -60 mV. Ventricular myocytes, in contrast, exhibit fast upstrokes (conduction velocity 1 m/s) and a more negative (-90 mV), ensuring rapid propagation for ventricular contraction. On the electrocardiogram (ECG), the corresponds to ventricular during phase 0 of the action potential in ventricular myocytes, typically lasting 60-100 ms in humans. Disruptions in function can lead to arrhythmias; for instance, often arises from mutations in K⁺ channel genes like KCNH2 or KCNQ1, which prolong phase 3 and increase the risk of . Similarly, mutations in the gene encoding the Na⁺ channel can enhance late Na⁺ currents, further extending the action potential duration.

Skeletal and smooth muscle action potentials

In fibers, action potentials are brief, typically lasting 2-5 milliseconds, and are initiated at the before propagating rapidly along the . These potentials travel deep into the fiber interior via transverse tubules (), which are specialized invaginations of the that ensure synchronous activation across the large fiber diameter. The in activates dihydropyridine receptors (DHPRs), L-type voltage-gated calcium channels embedded in the T-tubule membrane, which serve as voltage sensors rather than primary calcium conduits in . Through direct physical coupling, DHPRs induce conformational changes in adjacent ryanodine receptors (RyR1) on the (), triggering rapid calcium release from intracellular stores and linking the action potential to contraction. This excitation-contraction coupling results in a fast response, characterized by quick onset and relaxation due to efficient calcium by SR pumps. Smooth muscle action potentials, in contrast, exhibit greater variability in duration, often ranging from 10 to 100 milliseconds depending on the tissue type and physiological conditions, reflecting adaptations for sustained tone rather than rapid movement. Unlike , smooth muscle lacks , so calcium influx primarily occurs through voltage-sensitive calcium channels (VSCCs), predominantly L-type, in the plasma membrane during . This extracellular calcium entry activates , which phosphorylates to initiate cross-bridge cycling, while additional calcium can be mobilized from stores via (IP3) pathways in response to agonists. The resulting excitation-contraction coupling supports tonic contractions that can be maintained with low energy expenditure, enabling prolonged force generation in structures like blood vessels and airways. Propagation of action potentials differs markedly between the two muscle types, influencing their functional roles. In , the action potential spreads uniformly along the and penetrates the fiber via the T-tubule system, ensuring coordinated contraction of individual fibers under voluntary control. , often organized in syncytia, propagates potentials primarily along the and through junctions connecting adjacent cells, allowing electrical coupling and wave-like spread for graded, collective responses. Disruptions in action potential regulation can lead to severe disorders, as seen with tetanus toxin produced by . The toxin inhibits release in central inhibitory by cleaving synaptobrevin (), resulting in of motor neurons and sustained, high-frequency firing that prolongs muscle activation and causes spastic paralysis. This hyperactivity manifests as tetanic contractions, where repeated action potentials fuse into prolonged tension without relaxation.

Action potentials in non-neuronal cells

Action potential-like electrical signals occur in various non-neuronal cells, including those in plants, fungi, and protists, serving roles in rapid signaling and response to environmental stimuli. In plants, these signals are prominent in excitable species such as the Venus flytrap (Dionaea muscipula) and characean algae like Chara corallina. In the Venus flytrap, action potentials propagate along the trap lobes at speeds of 5–25 cm/s, triggering rapid closure upon mechanical stimulation of trigger hairs. These potentials in characean algae are generated in response to mechanical injury or electrical stimulation, influencing cytoplasmic streaming and photosynthetic activity. Unlike neuronal action potentials, plant versions are slower, with durations often exceeding 100 ms and propagation velocities of 0.04–0.6 m/s. The mechanisms of plant action potentials involve a sequence of ion fluxes distinct from those in neurons. Depolarization is primarily mediated by influx of Ca²⁺ through voltage-gated channels and efflux of Cl⁻ via anion channels, leading to membrane potential changes from a resting state of around -100 mV to peaks near 0 mV. Repolarization follows through K⁺ efflux and active H⁺ extrusion by plasma membrane H⁺-ATPases, restoring the electrochemical gradient. These signals propagate symplastically through plasmodesmata, connecting adjacent cells and coordinating responses across tissues. In wound responses, action potentials facilitate long-distance signaling, activating defense gene expression and hydraulic changes that propagate turgor waves. Similar electrical impulses appear in fungi and slime molds, though often graded rather than strictly all-or-none. In the fungus Neurospora crassa, intracellular recordings reveal action potential-like spikes with amplitudes of 10–100 mV and periods of 0.2–4 minutes, potentially involved in hyphal tip growth coordination. Oyster fungi (Pleurotus djamor) exhibit spontaneous high- and low-frequency spikes, with durations of approximately 3 minutes and 14 minutes, respectively, and responses to stimuli like ethanol or temperature, showing inter-colony signaling latencies of 26–51 seconds. Slime molds such as Physarum polycephalum generate propagating electrical potentials coupled with hydraulic flows, aiding in foraging and wound healing, where signals decrement in amplitude over distance, characteristic of graded propagation. These non-neuronal action potentials differ from the all-or-none neuronal type by being more variable in amplitude and often integrated with hydraulic or turgor signals, reflecting adaptation to non-conductive tissues. Their mechanisms, involving Ca²⁺ and Cl⁻ rather than Na⁺, underscore ancient origins predating metazoan neurons, tracing back to stem eukaryotes like and protists where they evolved for damage repair and motility.

Biophysical and Evolutionary Aspects

Taxonomic distribution and evolution

Action potentials are absent in prokaryotes, such as , which rely on alternative forms of chemical and mechanical signaling for cellular communication. In eukaryotes, they exhibit a broad taxonomic distribution, appearing in various protists, fungi, , and all metazoans. For instance, in protists like the Paramecium caudatum, calcium-based action potentials regulate ciliary beating to modulate swimming trajectories and escape responses. Similarly, action potential-like electrical spikes have been recorded in fungal mycelia, such as in the oyster fungus , where trains of impulses propagate along hyphae, potentially coordinating growth or resource distribution. In , action potentials occur independently of neural structures, as seen in , where they travel rapidly through conduits to trigger seismonastic leaf folding in response to mechanical stimuli. The evolutionary origins of action potentials trace back to stem eukaryotes around 1.5 billion years ago in the last eukaryotic , initially as calcium-mediated responses to membrane damage that coupled to and secretion for cellular protection. In metazoans, sodium-based action potentials emerged later, approximately 600 million years ago during the period, coinciding with the diversification of early animal lineages and the . Voltage-gated ion channel families critical for these potentials expanded convergently across major metazoan clades, including cnidarians, ctenophores, and bilaterians, rather than at a single common , enabling the independent development of neural complexity. action potentials, by contrast, evolved separately, utilizing distinct ion fluxes (primarily calcium and ) without reliance on the sodium channels predominant in animals. A key evolutionary advantage of action potentials is their capacity for rapid, all-or-none electrical propagation over long distances, which facilitated multicellular coordination and the division of labor between sensory, conductive, and effector cells in early metazoans. This mechanism supported adaptive responses like escape behaviors, enhancing survival in complex environments and contributing to the success of animal multicellularity. highlights differences between and vertebrates: invertebrate action potentials often depend on large, unmyelinated axons, as in the , which achieves high conduction velocities through sheer diameter for rapid signaling in neuromuscular systems. Vertebrate versions, however, incorporate myelin sheaths for , allowing efficient propagation in finer axons and optimizing energy use across diverse types. Recent insights underscore the expanded role of action potential-like events beyond neurons, including in glial cells where oligodendrocyte precursor cells can generate sodium-dependent action potentials to support migration and differentiation during development.

Quantitative biophysical models

The Hodgkin-Huxley model, developed in 1952, provides a foundational quantitative framework for simulating action potential generation in neuronal membranes by describing the dynamics of voltage-dependent sodium and ion channels. The model treats the membrane as an electrical circuit with and conductances that vary over time due to gating variables, capturing the rapid influx of sodium s during and efflux of ions during . The core equation governing V is: \frac{dV}{dt} = \frac{ I - g_{\mathrm{Na}} m^3 h (V - E_{\mathrm{Na}}) - g_{\mathrm{K}} n^4 (V - E_{\mathrm{K}}) - g_{\mathrm{L}} (V - E_{\mathrm{L}}) }{C_m} where I is the applied current, g_{\mathrm{Na}}, g_{\mathrm{K}}, and g_{\mathrm{L}} are maximum conductances for sodium, potassium, and leak channels, E_{\mathrm{Na}}, E_{\mathrm{K}}, and E_{\mathrm{L}} are potentials, and C_m is ; the gating variables m, h, and n ( for sodium, inactivation for sodium, and for potassium) follow first-order kinetics defined by voltage-dependent rate constants \alpha and \beta. This deterministic system of four coupled ordinary differential equations accurately reproduces the action potential's shape and threshold in the , with parameters fitted from voltage-clamp experiments. Extensions of the Hodgkin-Huxley model simplify its complexity while preserving key excitable dynamics, such as the FitzHugh-Nagumo model introduced in , which reduces the four-dimensional system to two variables: a fast activator () and a slow recovery variable (approximating activation and sodium inactivation). The FitzHugh-Nagumo equations are: \begin{align*} \frac{dv}{dt} &= v - \frac{v^3}{3} - w + I, \\ \frac{dw}{dt} &= \epsilon (v + a - b w), \end{align*} where v represents voltage, w the , and parameters \epsilon, a, and b control excitability and time scales; this simplification highlights bifurcations leading to oscillatory or spiking behavior without detailed channel kinetics. A related by Nagumo et al. in 1962 further adapted it for circuit simulations, emphasizing and properties in excitable media. Compartmental models extend the Hodgkin-Huxley framework to spatially distributed structures like axons or dendrites by dividing neurons into segments connected by axial resistances, allowing simulation of action potential along multi-segment geometries. The software, developed by Hines and Carnevale in 1997, implements this approach efficiently using implicit integration methods to handle variable time steps and morphologies, enabling realistic modeling of branched axons with nonuniform channel distributions. These models find applications in predicting action potential propagation failure, such as in discrete models where branch points or demyelination cause conduction block due to insufficient sodium current recruitment. They also simulate drug effects on excitability, for instance, by modifying conductance parameters to assess how alter threshold and firing rates in interaction studies. A key limitation of the standard Hodgkin-Huxley model is its deterministic nature, which assumes continuous, infinite populations of channels and neglects stochastic fluctuations from discrete channel openings, leading to inaccuracies in small compartments or low channel densities where can trigger spontaneous firing or alter reliability. extensions, such as representations of channel states, better capture this variability but increase computational demands.

Neurotoxins and disorders

Neurotoxins can profoundly disrupt action potential generation and propagation by targeting voltage-gated ion channels, particularly sodium channels, leading to paralysis or loss of excitability in affected tissues. Tetrodotoxin (TTX), a potent neurotoxin produced by certain bacteria in pufferfish and other marine organisms, selectively binds to and blocks voltage-gated sodium channels (NaV1.x), preventing sodium influx necessary for the depolarization phase of the action potential. This blockade inhibits action potential initiation and conduction in nerves and muscles, resulting in rapid paralysis and potentially fatal respiratory failure if ingested. Local anesthetics, such as lidocaine, exert their effects by binding preferentially to the inactivated state of voltage-gated sodium channels, stabilizing this conformation and reducing the availability of channels for reopening during repetitive firing. This use-dependent inhibition dampens high-frequency action potentials in sensory and motor neurons, providing localized analgesia by preventing pain signal transmission without completely abolishing low-frequency physiological activity. Disorders involving aberrant action potential dynamics often stem from genetic mutations in s, known as , which alter neuronal excitability. exemplifies hyperexcitability, where imbalances in excitatory and inhibitory signaling lower the threshold for action potential firing, leading to synchronized neuronal bursts and seizures; this can arise from enhanced excitatory synaptic transmission or reduced inhibition, facilitating abnormal propagation of action potentials across neural networks. A specific channelopathy, , results primarily from loss-of-function mutations in the SCN1A gene encoding the NaV1.1 subunit, which is crucial for action potential generation in inhibitory . These mutations impair interneuron firing, reducing and promoting network hyperexcitability that manifests as early-onset, therapy-resistant seizures. Botulinum toxin, produced by Clostridium botulinum bacteria, indirectly disrupts action potentials at the by cleaving SNARE proteins (e.g., SNAP-25), thereby inhibiting release from presynaptic motor neurons. This presynaptic blockade prevents of the muscle endplate, halting the initiation of action potentials in fibers and causing , which can be lethal if respiratory muscles are affected. Therapeutic interventions, such as anti-epileptic drugs (AEDs), counteract pathological action potential dysregulation by stabilizing neuronal membranes and raising the threshold for firing. Many AEDs, including and , target voltage-gated sodium channels by prolonging their inactivated state, thereby suppressing repetitive high-frequency action potentials while sparing single, physiological ones; this selective action reduces propensity in hyperexcitable circuits.

Historical and Experimental Context

Discovery and historical development

The concept of "animal electricity" emerged in the late through the pioneering experiments of , an physician and physicist. In 1786, while dissecting during a , Galvani observed that the muscles contracted when touched by a metal near a static electricity source, leading him to hypothesize an intrinsic electrical fluid within animals that drove muscular motion. He further demonstrated this by connecting frog nerves to different metals, such as iron and , which induced contractions without external electricity, attributing the effect to the frog's own "electric virtue." These findings, detailed in his 1791 treatise De Viribus Electricitatis in Motu Musculari Commentarius, sparked debates on bioelectricity and laid the groundwork for understanding nerve and muscle excitability, though Galvani's interpretations were later refined by contemporaries like . Advancing into the early , Julius Bernstein proposed the first membrane theory of bioelectric potentials in 1902. Drawing on measurements of resting potentials in frog muscle and , Bernstein suggested that cell membranes were selectively permeable to ions (K⁺) at rest, creating a potential that accounted for the negative intracellular voltage relative to the exterior. He posited that during , this selective permeability temporarily broke down, allowing other ions to flow and generate the action potential as a transient . This hypothesis, outlined in his seminal paper Untersuchungen zur Thermodynamik der bioelektrischen Ströme, provided a biophysical framework for excitability and influenced subsequent research, though it required ionic current mechanisms to fully explain propagation. The mechanisms of action potentials were decisively elucidated in the mid-20th century by and through their studies on the , conducted primarily from the 1930s to 1950s at the Laboratory of the Marine Biological Association in , . In 1939, they recorded intracellular action potentials, revealing rapid voltage changes, but faced challenges in isolating ionic contributions. By 1947, they developed the voltage-clamp technique, which held constant while measuring currents, allowing separation of sodium (Na⁺) influx during and potassium efflux during . Their comprehensive 1952 series of papers quantified these processes, demonstrating that action potentials result from dynamics, a model that revolutionized and earned them the 1963 in Physiology or Medicine (shared with John Eccles). Building on these foundations, advanced understanding of action potential-triggered synaptic transmission in the 1950s and 1960s, particularly the role of calcium (Ca²⁺) ions. Using the in frog preparations, Katz and Ricardo Miledi showed in 1965 that extracellular Ca²⁺ is essential for release, as reducing it diminished end-plate potentials while increasing magnesium blocked release competitively. Their "quantal" hypothesis posited that action potentials trigger vesicular release in discrete packets, modulated by presynaptic Ca²⁺ influx, explaining synaptic reliability and plasticity. This work, integral to Katz's 1970 in Physiology or Medicine (shared with and Ulf von Euler), bridged action potentials to chemical signaling. A key methodological milestone came in the late 1970s and 1980s with the patch-clamp technique, developed by Erwin Neher and Bert Sakmann. This innovation used glass micropipettes to form a high-resistance seal ("gigaseal") on small membrane patches, enabling recording of single currents at picopampere resolution. Their 1976 demonstration and refined 1981 methods allowed direct observation of channel gating during action potentials, confirming Hodgkin-Huxley's predictions at the molecular level and facilitating studies of channel diversity. For this breakthrough, Neher and Sakmann received the 1991 in or .

Experimental measurement techniques

The experimental measurement of action potentials has relied on intracellular recording techniques since the mid-20th century, with sharp microelectrodes enabling precise voltage measurements and control within single cells. These glass micropipettes, filled with conductive solutions like , are inserted into the cell interior to record changes directly, capturing the rapid and phases of action potentials with high fidelity. The technique was pivotal in the development of the voltage-clamp method, which holds the constant while measuring ionic currents, allowing dissection of sodium and contributions to action potential generation. A seminal application involved inserting two microelectrodes into the —one to measure voltage and another to inject —demonstrating that action potentials arise from voltage-dependent ionic conductances. Sharp electrodes remain valuable for intact tissues where minimal is needed, though their high resistance limits current injection compared to other methods. Extracellular recording methods provide a less invasive alternative, detecting action potentials through field potentials generated by population activity without penetrating cells. These techniques measure voltage changes in the using metal or electrodes placed near neuronal ensembles, capturing biphasic from synchronized firing. Field potentials, often recorded with a single , reflect summed synaptic and action potential currents, offering insights into dynamics in slices or cultures. Multi-electrode arrays (MEAs), consisting of dozens to thousands of closely spaced sites, enable simultaneous recording from multiple neurons, improving for mapping action potential propagation. Early MEAs, developed in the , used planar platinum-black electrodes to monitor extracellular from cultured neurons over extended periods, establishing the foundation for chronic studies. Modern high-density MEAs, with micrometer-scale spacing, facilitate spike sorting and localization of individual action potentials . The patch-clamp technique, introduced in the 1970s, revolutionized single-cell by allowing access to ionic currents underlying action potentials at the level of individual channels. In cell-attached mode, a glass pipette forms a high-resistance seal on the intact membrane, enabling recording of single-channel currents without disrupting the cell's interior, which is useful for studying native channel properties during spontaneous action potentials. Whole-cell configuration ruptures the patch beneath the pipette, dialyzing the cell interior for voltage-clamp control and direct measurement of total currents driving action potentials, though it can alter intracellular milieu over time. This method confirmed the discrete, probabilistic opening of voltage-gated sodium and potassium channels during the action potential upstroke and . Patch clamping has been adapted for action potential studies in diverse preparations, from isolated neurons to slices, providing sub-millisecond . Optical techniques offer non-contact methods to record and manipulate action potentials, leveraging light-sensitive probes for high-throughput imaging across populations. Voltage-sensitive dyes, amphipathic molecules that insert into membranes and fluoresce or absorb light proportional to voltage changes, enable visualization of action potential wavefronts in excitable tissues. These dyes, first demonstrated to detect millisecond-scale signals in axons, allow simultaneous monitoring of hundreds of neurons without electrical artifacts. Optogenetics extends this by using light to trigger action potentials via genetically encoded channels like channelrhodopsin-2, a light-gated cation channel expressed in target cells to evoke precise, millisecond-resolution depolarizations. This tool, initially applied to control neuronal firing in culture and slices, facilitates causal studies of action potential roles in circuits. In vivo measurements have advanced with techniques like two-photon microscopy, which uses infrared lasers to excite voltage indicators deep in scattering brain tissue, minimizing photodamage while resolving action potentials in behaving animals. Two-photon excitation, enabling volumetric imaging at depths up to 1 mm, has captured somatic and dendritic action potentials in cortical layers with subcellular precision. Recent 2020s developments include high-density silicon probes, such as Neuropixels arrays with thousands of recording sites along a , allowing stable, chronic isolation of up to hundreds of single units in freely moving . These probes, with site densities exceeding 100 per mm, improve yield and stability for long-term action potential tracking during behavior.

Modern computational modeling

Modern computational modeling of action potentials has advanced significantly beyond classical deterministic frameworks, incorporating elements, multiscale integration, and data-driven approaches to capture the complexity of neuronal signaling. Software environments like and enable detailed simulations that account for channel noise, reflecting the probabilistic opening and closing of ion channels in real neurons. , a widely used tool for modeling individual neurons and networks, supports simulations by implementing Markov kinetic schemes for voltage-gated channels, allowing researchers to quantify how channel noise affects action potential reliability, such as propagation failures in thin axons. Similarly, facilitates simulations across scales, from molecular reactions to network activity, using Gillespie algorithms for discrete events like channel gating, which reveal noise-induced variability in action potential timing and amplitude. These tools have been instrumental in studying how intrinsic noise influences neuronal excitability without relying on deterministic approximations. Multiscale models extend these capabilities by linking molecular-level details, such as Markov state models of kinetics, to higher-level network dynamics, providing a unified for action potential across cellular compartments and circuits. In such models, Markov state models describe the conformational changes of individual channels with , capturing transitions that feed into compartmental simulations of dendritic and axonal action potentials, ultimately informing network-level behaviors like synchronized firing. For instance, multiscale approaches have integrated presynaptic with synaptic release probabilities to simulate action potential-triggered dynamics, bridging timescales from microseconds to milliseconds. This integration allows for the exploration of emergent properties, such as how molecular noise to alter circuit-level information encoding. In the 2020s, and have emerged as powerful tools for predicting action potential characteristics directly from genetic data, accelerating the personalization of models for disease-related variants. frameworks, including convolutional neural networks and models, analyze genetic variants in genes to forecast their impact on channel function, thereby predicting alterations in action potential , , and . For example, models trained on electrophysiological data from variants in voltage-gated sodium and calcium channels can estimate gain- or loss-of-function effects, enabling simulations of how genetic mutations distort action potentials in conditions like . These AI-driven predictions complement traditional biophysical modeling by handling high-dimensional genomic inputs, with neural networks achieving accuracies exceeding 80% in classifying variant pathogenicity based on simulated action potential perturbations. At the brain-scale, projects like the employ large-scale simulations to model action potentials within reconstructed cortical columns, integrating thousands of detailed neurons to replicate dynamics. These simulations use multicompartmental models with stochastic channel kinetics to generate realistic spiking patterns in rat somatosensory cortex, validating against experimental data on connectivity and . By scaling up to millions of synapses, such efforts reveal how action potential synchrony emerges in microcircuits, informing hypotheses on cortical computation. Despite these advances, challenges persist in parameter fitting and model validation, particularly for ensuring biological in complex simulations. Parameter estimation often involves optimization techniques like to match simulated action potentials to experimental traces, but ill-posed problems arise due to trade-offs between channel densities and kinetics, leading to non-unique solutions. Validation increasingly relies on , where light-sensitive channels are incorporated into models to predict neuronal responses to precise perturbations, as demonstrated by empirically derived models of channelrhodopsin-2 that accurately replicate voltage and light dependencies in action potential modulation. These methods highlight the need for hybrid experimental-computational pipelines to constrain parameters and verify predictions against high-resolution optogenetic data.

References

  1. [1]
    Physiology, Action Potential - StatPearls - NCBI Bookshelf - NIH
    An action potential is a rapid sequence of changes in the voltage across a membrane. The membrane voltage, or potential, is determined at any time by the ...
  2. [2]
    Chapter 1. Resting Potentials & Action Potentials
    Action potentials are the basic events the nerve cells use to transmit information from one place to another.
  3. [3]
    A quantitative description of membrane current and its application to ...
    This paper by Hodgkin and Huxley provides a quantitative description of membrane current and its application to nerve conduction and excitation.
  4. [4]
    [PDF] History of Electrophysiology
    Ludimar Hermann introduced the term of 'action current' for this electrical wave. potential as early as 1868. This finding was confirmed by Hermann in 1881, ...Missing: coined | Show results with:coined
  5. [5]
    Action potentials and synapses - Queensland Brain Institute
    Action potentials are the fundamental units of communication between neurons and occur when the sum total of all of the excitatory and inhibitory inputs makes ...
  6. [6]
    Physiology, Resting Potential - StatPearls - NCBI Bookshelf - NIH
    The membrane is permeable to K+ at rest because many channels are open. In a normal cell, Na+ permeability is about 5% of the K+ permeability or even less, ...
  7. [7]
    Physiology, Sodium Potassium Pump - StatPearls - NCBI Bookshelf
    Mar 13, 2023 · The Na + K + pump is an electrogenic transmembrane ATPase first discovered in 1957 and situated in the outer plasma membrane of the cells on the cytosolic side.
  8. [8]
  9. [9]
    Resting Membrane Potential - Nernst - TeachMe Physiology
    Jul 13, 2023 · The Na/K ATP-ase pump plays an essential role in maintaining the sodium and potassium concentrations by actively transporting these ions against ...
  10. [10]
    Crystal Structure of the Human Two–Pore Domain Potassium ...
    Two–pore domain potassium (K2P) channels conduct K+ ions across the plasma membrane of eukaryotic cells. They help to maintain the cellular resting ...
  11. [11]
    Threshold Potential - an overview | ScienceDirect Topics
    Threshold potential is the transmembrane voltage at which a neuron's depolarization opens Na+ channels, initiating an action potential.
  12. [12]
    Summation of Synaptic Potentials - Neuroscience - NCBI Bookshelf
    The summation of EPSPs and IPSPs by a postsynaptic neuron permits a neuron to integrate the electrical information provided by all the inhibitory and ...Missing: initiation | Show results with:initiation
  13. [13]
    Excitatory and Inhibitory Postsynaptic Potentials - NCBI - NIH
    PSPs are called excitatory (or EPSPs) if they increase the likelihood of a postsynaptic action potential occurring, and inhibitory (or IPSPs) if they decrease ...
  14. [14]
    The Refractory Period - Neuroscience - NCBI Bookshelf - NIH
    The refractoriness of the membrane in the wake of the action potential explains why action potentials do not propagate back toward the point of their initiation ...
  15. [15]
    Refractory Period - an overview | ScienceDirect Topics
    During the relative refractory period, initiation of a second action potential is inhibited but not impossible; a larger-than-normal stimulus can result in ...Missing: influence | Show results with:influence
  16. [16]
    A quantitative description of membrane current and its application to ...
    A quantitative description of membrane current and its application to conduction and excitation in nerve. A. L. Hodgkin,. A. L. Hodgkin.
  17. [17]
    The Hodgkin-Huxley Heritage: From Channels to Circuits - PMC
    They described sodium selectivity, voltage-dependent activation, and fast inactivation, and they developed a quantitative model for action potential generation ...
  18. [18]
    The 'all or none' contraction of the amphibian skeletal muscle fibre
    The Journal of Physiology · Volume 38, Issue 2-3 pp. 113-133 The Journal ... The 'all or none' contraction of the amphibian skeletal muscle fibre. Keith Lucas,.
  19. [19]
    Neuroanatomy, Neuron Action Potential - StatPearls - NCBI Bookshelf
    Neurons are electrically excitable, reacting to input via the production of electrical impulses, propagated as action potentials throughout the cell and its ...
  20. [20]
    The Myelin Sheath - Basic Neurochemistry - NCBI Bookshelf
    In unmyelinated fibers, impulse conduction is propagated by local circuits of ion current that flow into the active region of the axonal membrane, through the ...Missing: electrotonic | Show results with:electrotonic
  21. [21]
    Increased Conduction Velocity as a Result of Myelination - NCBI - NIH
    For example, whereas unmyelinated axon conduction velocities range from about 0.5 to 10 m/s, myelinated axons can conduct at velocities up to 150 m/s. The major ...<|control11|><|separator|>
  22. [22]
    Neuroanatomy, Unmyelinated Nerve Fibers - StatPearls - NCBI - NIH
    Continuous transmission of this action potential thus causes nerve conduction in unmyelinated axons. Categories of Unmyelinated Nerve Fibers. Most nociceptive ...
  23. [23]
    Histology, Schwann Cells - StatPearls - NCBI Bookshelf - NIH
    May 1, 2023 · Schwann cells serve as the myelinating cell of the PNS and support cells of peripheral neurons. A Schwann cell forms a myelin sheath by wrapping ...Structure · Function · Histochemistry and... · Microscopy, Electron
  24. [24]
    Oligodendrocytes: Myelination and Axonal Support - PMC
    This review summarizes our current understanding of how myelin is generated and also the role of oligodendrocytes in supporting the long-term integrity of ...
  25. [25]
  26. [26]
  27. [27]
    Saltatory Conduction along Myelinated Axons Involves a Periaxonal ...
    Dec 26, 2019 · The propagation of electrical impulses along axons is highly accelerated by the myelin sheath and produces saltating or “jumping” action ...
  28. [28]
    Demyelination in multiple sclerosis - PMC - PubMed Central - NIH
    The second part of the review describes the axonal consequences of demyelination. Segmental demyelination results in conduction block or slowing of conduction ...
  29. [29]
    Conduction in Segmentally Demyelinated Mammalian Central Axons
    Oct 1, 1997 · The prominent symptoms associated with central demyelinating diseases such as multiple sclerosis (MS) are primarily caused by conduction ...
  30. [30]
    The electrical constants of a crustacean nerve fibre - Journals
    Theoretical equations are derived for the response of a nerve fibre to the sudden application of a weak current. The equations describe the behaviour of the ...
  31. [31]
    Synaptic Transmission - Basic Neurochemistry - NCBI Bookshelf - NIH
    Chemical transmission between nerve cells involves multiple steps · 1. Synthesis of the neurotransmitter in the presynaptic nerve terminal. · 2. Storage of the ...Chemical transmission... · Quantal analysis defines the... · Ca2+ is necessary for...
  32. [32]
    Organization of Cell Types (Section 1, Chapter 8) Neuroscience ...
    There is a distinct synaptic cleft or intercellular space of approximately 20-40 nm. An endoplasmic reticulum is present that regulates Ca2+ level ...
  33. [33]
    Physiology, Synapse - StatPearls - NCBI Bookshelf
    Mar 27, 2023 · The chemical neurotransmitter released must diffuse across this cleft to elicit a response in the receptive neuron. Because of this, the ...Missing: dimensions | Show results with:dimensions
  34. [34]
    Spontaneous subthreshold activity at motor nerve endings - PMC
    FATT P., KATZ B. An analysis of the end-plate potential recorded with an intracellular electrode. J Physiol. 1951 Nov 28;115(3):320–370. doi: 10.1113 ...
  35. [35]
    Electrical Synapses - Neuroscience - NCBI Bookshelf - NIH
    (A) Gap junctions consist of hexameric complexes formed by the coming together of subunits called connexons, which are present in both the pre- and postsynaptic ...
  36. [36]
    The Permeability of Gap Junction Channels to Probes of Different ...
    Estimates based on size cutoffs for permeants suggest that the effective pore diameter varies over a range of 8–16 Å, depending on the connexin isoform (Beblo ...
  37. [37]
    Electrical synapses: structure, locations, diagram | Kenhub
    Nov 26, 2024 · Each gap junction is composed of two perpendicular connexons; one in the presynaptic and one in the postsynaptic membrane. Connexins are ...
  38. [38]
    On the Diverse Functions of Electrical Synapses - Frontiers
    Electrical synapses are the neurophysiological product of gap junctional pores between neurons that allow bidirectional flow of current between neurons.
  39. [39]
    Electrical synapses and their functional interactions with chemical ...
    Rather than conceiving synaptic transmission as either chemical or electrical, this article emphasizes the notion that synaptic transmission is both chemical ...
  40. [40]
    Electrical synapses convey orientation selectivity in the mouse retina
    Dec 11, 2017 · Gap junctions are particularly prevalent in the retina, where all five major cell classes are connected by electrical synapses. In vertebrates, ...
  41. [41]
    Approaches to Study Gap Junctional Coupling - PMC
    Mar 10, 2021 · Gap junction channels connect the cytosol of neighboring cells and allow the exchange of ions and small molecules, such as metabolites (Giaume ...
  42. [42]
    Modulation of metabolic communication through gap junction ...
    Gap junction (GJ) channels assembled from connexin (Cx) proteins provide a structural basis for direct electrical and metabolic cell–cell communication.
  43. [43]
    Gap Junctions and Epileptic Seizures – Two Sides of the Same Coin?
    Electrical synapses (gap junctions) play a pivotal role in the synchronization of neuronal ensembles which also makes them likely agonists of pathological ...
  44. [44]
    Epileptogenic Agents Enhance Transmission at an Identified Weak ...
    If convulsants have a similar action in mammalian cortex, enhanced transmission at weak electrical synapses may provide abnormal pathways for the flow of ...
  45. [45]
    Physiology, Neuromuscular Junction - StatPearls - NCBI Bookshelf
    Feb 17, 2025 · These folds are densely packed with nicotinic acetylcholine receptors (nAChRs), which function as ligand-gated ion channels.
  46. [46]
    The Neuromuscular Junction in Health and Disease - PubMed Central
    Dec 3, 2020 · The neuromuscular junction (NMJ) is a highly specialized synapse between a motor neuron nerve terminal and its muscle fiber that are responsible for converting ...
  47. [47]
    Post-synaptic specialization of the neuromuscular junction
    Jun 19, 2022 · The sarcolemma within NMJ folded to form some invagination portions called junctional folds (JFs), and they have important roles in maintaining ...
  48. [48]
    Bernard Katz, quantal transmitter release and the foundations ... - NIH
    Fatt & Katz (1952) reached the fundamental conclusion that minis result from the spontaneous release of ACh from the presynaptic motor neuron. Their experiments ...
  49. [49]
    Acetylcholine Release - an overview | ScienceDirect Topics
    Typically, 50–300 vesicles are exocytosed in response to an action potential, with ∼10000 acetylcholine molecules released per vesicle. Once released from ...
  50. [50]
    Chapter 4: Synaptic Transmission and the Skeletal Neuromuscular ...
    The synapse at the neuromuscular junction has three characteristic features of chemical synapses in the nervous system. First, there is a distinct separation ...
  51. [51]
    Physiology, Acetylcholinesterase - StatPearls - NCBI Bookshelf - NIH
    Jan 19, 2023 · Acetylcholinesterase (AChE) is a cholinergic enzyme primarily found at postsynaptic neuromuscular junctions, especially in muscles and nerves.
  52. [52]
    Acetylcholinesterase - an overview | ScienceDirect Topics
    ACh molecules that do not bind immediately with a receptor or those released after reacting with a receptor are hydrolyzed almost instantly (in less than 1 ms) ...
  53. [53]
    Myasthenia Gravis - StatPearls - NCBI Bookshelf - NIH
    Myasthenia gravis is an autoimmune disorder affecting the neuromuscular junction. It manifests as a generalized muscle weakness which can involve the ...
  54. [54]
    BOTULINUM TOXIN - PMC - NIH
    Intramuscular administration of botulinum toxin acts at the neuromuscular junction to cause muscle paralysis by inhibiting the release of acetylcholine from ...
  55. [55]
    A comparative review on heart ion channels, action potentials ... - NIH
    Sep 8, 2021 · This review compares types of human action potentials, the dominant ion currents during action potential phases, alteration in ion channels activities in ...
  56. [56]
    Cardiac Ion Channels | Circulation: Arrhythmia and Electrophysiology
    Apr 1, 2009 · This review describes the generation of the normal cardiac action potential. The properties of the major ionic currents are the examined in detail.Missing: paper | Show results with:paper
  57. [57]
    [PDF] Comparison between neuronal action potential and action potentials ...
    In skeletal muscle cells, the action potential duration is approximately 2-5 ms. In contrast, the duration of cardiac action potentials ranges from 200 to 400 ...
  58. [58]
    Physiology, Skeletal Muscle Contraction - StatPearls - NCBI Bookshelf
    May 1, 2023 · Propagation of action potentials to the motoneuron and subsequent depolarization results in the opening of voltage-gated calcium (Ca2+) channels ...
  59. [59]
  60. [60]
    Physiology, Ryanodine Receptor - StatPearls - NCBI Bookshelf
    Sep 26, 2022 · Ryanodine receptors are essential for excitation-contraction coupling, linking action potentials and contraction of the striated muscle by releasing calcium ...Missing: characteristics | Show results with:characteristics
  61. [61]
    Smooth muscle cell calcium activation mechanisms - Berridge - 2008
    Oct 30, 2008 · The Ca2+ transients also activate Ca2+-sensitive chloride channels (CLCAs) that result in membrane depolarization to activate voltage-operated ...
  62. [62]
    Physiology, Smooth Muscle - StatPearls - NCBI Bookshelf
    Aug 14, 2023 · Smooth muscle contraction depends on calcium influx. Calcium increases within the smooth muscle cell through two different processes. First, ...
  63. [63]
    Calcium Signaling in Smooth Muscle - PMC - PubMed Central - NIH
    Membrane potential plays an important role in all excitable cells, including smooth muscle, where it regulates [Ca2+]i and thereby smooth muscle contraction.
  64. [64]
    Muscle Physiology - UW Pressbooks
    Smooth muscle cells lack t-tubules and have less SR. Excitation-contraction coupling in smooth muscle differs substantially from skeletal or cardiac muscle.
  65. [65]
    10.3 Muscle Fiber Excitation, Contraction, and Relaxation
    Propagation of an action potential along the sarcolemma is the excitation portion of excitation-contraction coupling and must be coupled to the release of ...
  66. [66]
    Tetanus: Pathophysiology, Treatment, and the Possibility of Using ...
    Jan 8, 2013 · The result is a partial, functional denervation of the lower motor neurons, which leads to their hyperactivity and to increased muscle activity ...
  67. [67]
    Local Tetanus Begins with a Neuromuscular Junction Paralysis ...
    Jun 14, 2024 · The most frequent form of tetanus is a generalized spastic paralysis that affects all skeletal muscles with painful contractures beginning from ...
  68. [68]
    The potential of plant action potentials | Synthese
    Nov 21, 2023 · This paper helps correct the imbalance in representations of action potentials by offering an overview of the mechanism for plant action potentials.3 Action Potentials · 3.2 Neuronal Action... · 4.2 Plant Cells Vs Neurons...
  69. [69]
    Action Potentials in a Giant Algal Cell: A Comparative Approach to ...
    The giant alga Chara corallina generates action potentials (APs) in response to mechanical stimulation, injury, or direct electrical stimulation.
  70. [70]
    Debunking a myth: plant consciousness - PMC - PubMed Central
    Nov 16, 2020 · Plant action potentials travel more slowly than those of animals, 0.04–0.6 m s−1 versus 0.5–100 m s−1, respectively, and with long refractory ...<|separator|>
  71. [71]
    Plant Ion Channels: Gene Families, Physiology, and Functional ...
    Rather, the depolarizing phase of action potentials is mediated by Cl− efflux channels and by Ca2+ channels (3–6).Plant Ion Channels: Gene... · Plant Chloride/anion... · Potassium Channels
  72. [72]
    [PDF] what do plants need action potentials for? - Esalq
    In plants the depolarization phase of AP consists of Cl-- and Ca2+-fluxes. The following phase—a repolarization—relies in turn on K+ fluxes and active H+ flows ...
  73. [73]
    Membrane voltage as a dynamic platform for spatiotemporal ...
    Unlike neuronal and other animal tissues, action potentials in plants are mediated by a temporal sequence of ion flux through Ca2+ and Cl- channels with voltage ...
  74. [74]
    Excitability and propagation of the electrical impulse in Venus flytrap
    In plants three types of electrical signals have been reported, action potentials, variation potentials and system potentials. Action potentials are related to ...
  75. [75]
    Fungal communication via electrical signals has inspired the ... - NIH
    Apr 5, 2023 · Spikes resembling neuronal action potentials were first reported as early as 1976 by making intracellular recordings from mycelium of the red ...
  76. [76]
    On spiking behaviour of oyster fungi Pleurotus djamor - Nature
    May 18, 2018 · We demonstrated that the fungi generate action potential like impulses of electrical potential. Trains of the spikes are observed.
  77. [77]
    [PDF] On electrical spiking of Ganoderma resinaceum - bioRxiv
    Jun 18, 2021 · It was demonstrated that patterns of electrical activity of the slime mould change in response to stimulation with volatile chemicals, ...
  78. [78]
    From damage response to action potentials: early evolution of ...
    Jan 5, 2016 · Elaborating on calcium-triggered membrane depolarization, we reason that the first action potentials evolved alongside the membrane of sensory- ...<|control11|><|separator|>
  79. [79]
    Excitable Membranes and Action Potentials in Paramecia - NIH
    Oct 15, 2015 · The ciliate Paramecium caudatum possesses an excitable cell membrane whose action potentials (APs) modulate the trajectory of the cell swimming through its ...
  80. [80]
    Early metazoan life: divergence, environment and ecology - Journals
    Dec 19, 2015 · The origin of basal metazoan clades occurred during the Cryogenian and early Ediacaran, to about 600 Ma. There have been many reviews of the ...
  81. [81]
    Convergence of ion channel genome content in early animal evolution
    We find that animals with nervous systems have broadly similar complements of ion-channel types but that these complements likely evolved independently.<|control11|><|separator|>
  82. [82]
    Action Potential Energy Efficiency Varies Among Neuron Types in ...
    Jul 1, 2010 · AP energy consumption varies greatly among vertebrate and invertebrate neurons, with several mammalian neuron models using close to the capacitive minimum of ...
  83. [83]
    Oligodendrocytes Changing the Rules: Action Potentials in Glia and ...
    A category of oligodendrocyte precursor cells (OPCs) has been identified that can fire action potentials, and their excitation is driven by synapses from axons.
  84. [84]
    The neuropathic pain triad: neurons, immune cells and glia - PubMed
    ... action potentials to the spinal cord, and onward transmission of the warning signal to the brain ... The neuropathic pain triad: neurons, immune cells and glia.
  85. [85]
    Impulses and Physiological States in Theoretical Models of Nerve ...
    July 1961, Pages 445-466. Journal home page for Biophysical Journal. Articles. Impulses and Physiological States in Theoretical Models of Nerve Membrane.
  86. [86]
    The NEURON Simulation Environment | Neural Computation
    Aug 15, 1997 · The NEURON simulation program provides a powerful and flexible environment for implementing such models of individual neurons and small networks of neurons.Missing: original | Show results with:original<|control11|><|separator|>
  87. [87]
    Asymptotic construction of pulses in the discrete Hodgkin-Huxley ...
    ... Hodgkin-Huxley model for myelinated nerves is given. Predictions of the shape and speed of the waves and the thresholds for propagation failure are obtained.
  88. [88]
    Parameterization for In-Silico Modeling of Ion Channel Interactions ...
    We describe an approach for developing a model for ion channel drug interactions. The method describes the process of extracting rate constants from ...
  89. [89]
    Limitations of the Hodgkin-Huxley Formalism: Effects of Single ...
    Abstract. A standard membrane model, based on the continuous deterministic Hodgkin-Huxley equations, is compared to an alternative membrane model, based.
  90. [90]
    TETRODOTOXIN BLOCKAGE OF SODIUM CONDUCTANCE ...
    It is concluded that tetrodotoxin, at very low concentrations, blocks the action potential production through its selective inhibition of the sodium-carrying ...Missing: channels | Show results with:channels
  91. [91]
    Recent advances in the study of mechanism of action of marine ...
    Tetrodotoxin and saxitoxin block the sodium channel selectively without any effect on other types of voltage-activated and transmitter-activated ion channels.
  92. [92]
    Setting up for the block: the mechanism underlying lidocaine's use ...
    Voltage-gated sodium channels play a critical role in the generation and propagation of action potentials in neurons and muscle cells. Lidocaine inhibition of ...
  93. [93]
    The Sodium Channel as a Target for Local Anesthetic Drugs - Frontiers
    Oct 31, 2011 · Na channels are the source of excitatory currents for the nervous system and muscle. They are the target for a class of drugs called local anesthetics (LA).
  94. [94]
    Basic Mechanisms Underlying Seizures and Epilepsy - NCBI
    Given that the basic mechanism of neuronal excitability is the action potential, a hyperexcitable state can result from increased excitatory synaptic ...
  95. [95]
    SCN1A channelopathies: Navigating from genotype to neural circuit ...
    Apr 17, 2023 · Over 80% of cases of Dravet Syndrome (DS) are caused by SCN1A variants, and 80–90% of these arise through de novo mutations. DS presents within ...
  96. [96]
    Dravet syndrome: a sodium channel interneuronopathy
    It is primarily caused by heterozygous loss-of-function mutations in the SCN1A gene that encodes brain voltage-gated sodium channel type-1, termed NaV1.1. Here ...
  97. [97]
    Antiepileptic Drugs: Overview, Mechanism of Action, Sodium ...
    Jun 7, 2024 · AEDs that target the sodium channels prevent the return of these channels to the active state by stabilizing them in the inactive state. In ...
  98. [98]
    Mechanisms of Action of Antiseizure Drugs and the Ketogenic Diet
    INaP facilitates epileptic burst firing by reducing the threshold for action potential generation, sustained repetitive firing, and augmentation of ...
  99. [99]
    Luigi Galvani's path to animal electricity - ScienceDirect.com
    Galvani carried out the experiments with artificial electricity from November 1780 up to February 1783, and he passed to the investigation of the effects of ...Missing: primary | Show results with:primary
  100. [100]
    [PDF] Alan Lloyd Hodgkin - Nobel Lecture
    1. Action potential (dotted curve) and increase in conductance (white band) in squid axon at about 6°C. From Cole and Curtis13.Missing: coined | Show results with:coined
  101. [101]
    The effect of calcium on acetylcholine release from motor nerve ...
    It is concluded that the action of calcium is concerned directly with the release of the transmitter, and not indirectly—as is sometimes suggested—by ...
  102. [102]
    A Direct Comparison of Whole Cell Patch and Sharp Electrodes by ...
    High-impedance, sharp intracellular electrodes were compared with whole cell patch electrodes by recording from single spinal neurons in immobilized frog ...Missing: seminal papers