Electrophysiologic Properties of the Mammalian Central Nervous System

Published on 26/03/2015 by admin

Filed under Neurosurgery

Last modified 26/03/2015

Print this page

rate 1 star rate 2 star rate 3 star rate 4 star rate 5 star
Your rating: none, Average: 0 (0 votes)

This article have been viewed 2956 times

CHAPTER 49 Electrophysiologic Properties of the Mammalian Central Nervous System

The study of excitable cells is, for a number of reasons, a fascinating one. Interest in electrophysiology and neuronal function spans many medical specialties because these excitable cells are those by which we move, think, and perform complex yet automatic tasks such as cardiovascular regulation. It is for these reasons that electrophysiologic studies have attracted the foremost physiologists in this century. Despite these outstanding contributions, several fundamental issues in neuroscience remain unresolved. Traditionally, clinical electrophysiology has used a more holistic approach than nonclinical neurophysiology has. Clinical insight into brain function (or dysfunction) is commonly achieved today by increasingly sophisticated imaging techniques that allow real-time observations. The booming advancement in molecular biology, as well as its fundamental contribution to medicine in general and neuroscience in particular, has unveiled an incredible level of ordered complexity in neuronal function. Basic scientists are producing a large quantity of molecular data, spanning from investigation of the role of a single protein in the electrical behavior of neurons to genetic markers of neurological disease. Despite the immense popularity of these approaches, it is important to remember that the electrical properties of individual neurons and the neuronal environment are the final effectors of brain activity and that diseases of the brain derive from the cellular level. It is thus foreseeable that recording of electrophysiologic signals will continue to provide a reliable method for in-depth investigation of central nervous system (CNS) function.

CNS function is dependent on homeostatic mechanisms that precisely regulate the extracellular level or concentration of neurotransmitters, ions, pH, and other variables. The neuronal cell membrane is a complex biochemical entity that interfaces between the cell and its environment. Its functions include directional transport of specific substances and maintenance of chemical gradients, particularly electrochemical gradients, across the plasma membrane. These ion gradients can be of high specificity (e.g., sodium versus potassium ions) and of great functional significance (e.g., in the production of action potentials). In addition, CNS function is supported by numerous non-neuronal mechanisms responsible for the control of extracellular and intracellular homeostasis (glial cells, cerebral vasculature). It has become increasingly evident that pathophysiologic changes in ion channel function play a major role in the etiology of certain disorders of the nervous system.

The following brief introductory chapter on CNS electrophysiology does not attempt to explain in detail the complex biophysical properties underlying communication between individual neurons or transduction of environmental and sensory signals into electrical activity in specific regions of the brain or spinal cord. Several excellent textbooks deal with specific aspects of CNS function and electrophysiology, and recent publications have described in a concise yet comprehensive manner the complex properties of the ion currents responsible for neuronal excitation. This chapter provides the reader with succinct background information on the electrical properties of neurons and additionally focuses on other aspects of brain function relevant to modern understanding of the pathophysiologic changes occurring in diseased brain. Such aspects include description of some of the mechanisms involved in brain homeostasis, the genesis of synchronous activity by electrotonic and ephaptic interactions, and the molecular changes in ion channels underlying neurological diseases. Because complete referencing of such a broad topic would entail a bibliography of thousands of references, relevant recent reviews, textbooks, and a nonexhaustive compilation of representative work are included.

Electrical Properties of Mammalian Cells

Matter is composed of atoms, which consist of positively charged nuclei and negatively charged electrons. Electrical phenomena occur whenever charges opposite in sign are separated or moved in a given direction; static electricity is the accumulation of electric charge. An electric current results when these charges flow across a permissive material, called a conductor. An ion current is a particular type of current carried by charges present on atoms or small molecules flowing in an aqueous solution. Separation of charges in an aqueous solution can be achieved by inserting an impermeable membrane in the solution itself. In mammalian cells, these membranes coincide with the plasma membrane, and its lipophilic composition ensures a remarkable level of electrical isolation for cells and tissues. Excitable as well as most nonexcitable cells are characterized by an asymmetric distribution of electric charges across the plasma membrane.

The biophysical bases for maintenance of this electric potential have been extensively investigated experimentally and modeled by mathematical simulations. Under normal resting conditions, mammalian cells allow transmembrane ion currents such that the internal portion of the cell is negatively charged; the presence of nonpermeant anions such as proteins also contributes to the maintenance of transmembrane potentials. This relatively stable state results in a net transmembrane potential of several millivolts and is commonly referred to as the resting membrane potential (RMP) (Fig. 49-1).

image

FIGURE 49-1 Relationship between resting potential and [K+]out. The line represents the “ideal” case predicted by the Nernst equation (see Equation 3). Experimental measurements show a significant departure from a linear dependency at physiologic [K+]out values. Resting potential values for excitable and nonexcitable central nervous system cells are shown. BBB, blood-brain barrier; RMP, resting membrane potential.

Most of what we know about the physiology of excitable cells was derived from electrical measurements (Fig. 49-2). Our knowledge of the complex properties of the CNS is based on the application to neuronal cells of simple physical rules governing the movement of charged particles. The physical principles of cell electrophysiology can be thus be compared with the biophysical rules governing flow of electric current through a so-called RC circuit (see Fig. 49-2), where electrons flow through a resistive component (in the case of living cells, represented by ion channels) set in parallel with a capacitive component (represented by the poorly conductive phospholipid bilayer). Ohm’s law describes the relationship between current flow (I) and the resulting voltage drop (E) across a resistor when no capacitive component is present. This is, of course, an abstract situation, in particular when dealing with ion fluxes across biologic membranes, but it allows one to understand the basic principles governing flow of electric current.

The following relationship constitutes Ohm’s law (or principle):

(1) image

where I represents the current flowing through a resistance (R) when voltage is applied. In a hypothetical cell, this relationship can be written as follows:

(2) image

where Em represents the difference in voltage (in millivolts) between the inside and the outside of the cell, Im represents the net current flowing at that particular time across the cell membrane, and Rm is the total membrane resistance. An important consequence of this relationship is that small changes in current will significantly affect cell RMP only when Rm is large. This is particularly important for inhibitory synaptic currents; for instance, in the case of γ-aminobutyric acid receptor A (GABAA) activation, inhibition is achieved not only by hyperpolarizing cell RMP further away from the firing threshold but also by greatly increasing postsynaptic conductance,* thus reducing the synaptic efficacy of concomitant excitatory signals.

A mathematical derivation of Ohm’s law useful for studies of biologic membranes was first formulated by Nernst,1 who described the relationship between intracellular and extracellular ion concentrations and changes in transmembrane potential attributable to permeation of an ion. In the case of potassium ions, the Nernst equation can be written as follows:

(3a) image

where R and F are constants, T is the temperature at which the observation is performed, and Z is the charge of the permeant ion. Note that if the charge (sign, or “Z” in Equation 3) of the permeant ion is changed (e.g., if we look at the Nernst equation for chloride), the direction of the gradient is changed as well. The logarithm of the ratio between intracellular and extracellular K+ dominates the right side of the equation because RT/ZF is constant under most biologic conditions when temperature can be maintained within a few degrees. This equation predicts, as expected from Ohm’s law, that net potassium flux will approach zero when intracellular and extracellular potassium levels are isosmolar, that is, when EK = 0 mV. Under physiologic conditions ([K+]out = 3.5 mM; [K+]in = 135 mM), the transmembrane potential at which potassium flux will be nil is around −90 mV.* This value constitutes the potassium equilibrium potential at these concentrations. Note that small changes in the extracellular potassium concentration will cause relatively large changes in the fraction of total membrane currents attributable to potassium ions. Because permeability to potassium is essential for maintenance of RMP, a net increase in extracellular potassium will cause a significant departure from RMP (see Fig. 49-1).

In practical form, the Nernst equation for potassium can be rewritten by converting to log10 and calculating RT/ZF at 20°C. This rearrangement leads to

(4) image

It can be seen that a 10-fold change in the concentration gradient for potassium can produce a 58-mV change in membrane potential. Because under normal conditions there is an almost 40-fold outward gradient for potassium ions and a 12-fold inward gradient for sodium ions, the resulting equilibrium potentials are −92 mV and +65 mV, respectively. Since the membrane at rest is much more permeable to potassium than to sodium (see later), RMP is closer to EK than to ENa.

Two additional considerations can help us understand the genesis of cell RMP: (1) if the membrane is exclusively permeant to K+ and no active, electrogenic transport of ions occurs, the cell potential will tend toward EK and only small movements of K+ will be sufficient to maintain RMP at −92 mV, and (2) if the membrane potential is clamped at EK, net potassium flux will be zero (as predicted by the fact that EK is the equilibrium potential for potassium). If the membrane potential is held positive with respect to EK, outward potassium currents will develop; conversely, if RMP is held at potentials negative with respect to EK, inward potassium fluxes will develop. Interestingly, the notion that cell RMP is controlled by potassium ions was first derived independently from direct observations: Julius Bernstein in 1902 correctly predicted that in most mammalian cells, potassium ions govern the transmembrane voltage difference.2 Direct experimental evidence was achieved only half a century later when a microelectrode could be placed into cells to directly measure RMP and the effects of changing [K+]out. Bernstein also proposed that selective potassium permeability was lost during the process of excitation, during which numerous “pores” opened to allow entry of other small ions (Cl and Na+).2 This theory explained several features of the regulation of RMP and generation of action potentials, including the depolarizing effects of [K+]out.

In fact, a prediction of the formalism just presented is that RMP changes linearly with [K+]out. As shown in Figure 49-1, experimental evidence contradicts this notion, at least in the case of mammalian neurons. Most neurons depolarize significantly after large changes in extracellular potassium, whereas changes around physiologic potassium concentrations do not significantly affect neuronal RMP. This can only be due to concomitant participation of other conductances. Although the exact nature of the ionic conductances contributing to the regulation of neuronal RMP varies between different neurons, a generic set of equations predict how these “parallel” conductances may affect Em. The most illustrious of these equations was provided by Goldman, who described the expected RMP in a cell endowed with more than one ion current mechanism as

(5) image

where GK, GNa, and GCl represent the conductances for potassium, sodium, and chloride.3 Note that these conductance values are multiplied by the relative chemical concentration gradient for each ion, thus combining the “passive” electro-osmotic tendency for ion permeation with the average conductance of the membrane for a particular ion. Note also that if GCl and GNa are close to zero, the transmembrane potential is governed almost exclusively by potassium ions and their conductance. This condition is common at resting potential in most neurons and glial cells, where IK (and to some extent Ipump) determines RMP (Fig. 49-3).

In 1949, Alan Hodgkin and Bernard Katz first applied the Goldman equation systematically to changes in membrane potentials evoked by altering external ion concentrations in the squid giant axon.4 They measured changes in RMP induced by changes in [K+]out, [Na+]out, and [Cl]out. They discovered that although changes in extracellular potassium dramatically changed RMP, comparable changes in [Na+]out had little effect. Changing [Cl]out had an intermediate effect. The following permeability ratios were obtained at rest:

(6a) image

When the measurements were performed at the peak of the action potential, however, these values changed dramatically to yield

(6b) image

Therefore, when the predominant membrane conductance is PK, the Goldman equation can be reduced to the Nernst equation for potassium:

(3b) image

whereas when PNa predominates, the following applies:

(3c) image

which is approximately the peak value of action potential overshoot. Direct measurement of changes in the relative permeability for Na+ and K+ contradicted one of the hypotheses formulated by Bernstein, who incorrectly predicted that neuronal excitation was due to loss of potassium permeability rather than activation of an inward sodium current (Table 49-1). Had this hypothesis been correct, the maximum depolarizing value reached during the action potential would have been around 0 mV and not at +30 mV as experimentally determined by Hodgkin and Katz.

TABLE 49-1 Glossary of Commonly Used Electrophysiologic Terms

Inward current Positive charges enter the cell (e.g., INa responsible for the action potential upstroke)
Outward current Positive charges leave the cell (e.g., IK during action potential repolarization)
Depolarization Change in RMP to less negative values (e.g., EPSP)
Hyperpolarization Change in RMP to more negative values (e.g., IPSP)
Inward-going rectification Tendency of some ionic currents to allow passage of inward-flowing but not outward-flowing ions (inward rectifier potassium currents)
Outward-going rectification Tendency of some currents to allow passage of outward-flowing but not inward-flowing ions (most other potassium currents activated by depolarization)
Voltage clamp Electrophysiologic technique allowing the study of ion currents and their modulation by voltage, or transmitters/second messengers
Current clamp Technique used during intracellular recordings to determine the resting properties and excitability of cells
Single-channel recording Modern variation (patch clamp) that allows study of the electrophysiologic properties of a single ion channel/protein
Multi–single-unit recording Extracellular recording from a neuron or a cluster of neighboring neurons

EPSP, excitatory postsynaptic potential; IPSP, inhibitory postsynaptic potential; RMP, resting membrane potential.

It has become apparent more recently that INa is not the only ionic current that can generate action potentials; calcium ions also play an important role in neuronal excitability. Similarly, IK is not the exclusive component of the electrical regulation of cell resting properties inasmuch as several other conductances are involved in the control of neuronal resting potential.

Ion Channels in Neurons and Glia

Ion channels are protein channels in cell membranes that allow ions to pass from the extracellular solution to the intracellular solution and vice versa. Similarly, transporters are specialized enzymes that carry specific ions or molecules across otherwise impermeant membranes or against electro-osmotic gradients. Not surprisingly, from a purely thermodynamic (or energetic) point of view, ion channels are less “expensive” to operate, whereas pumps or exchangers require considerable consumption of energy. Most ion channels and pumps are selective in that they allow only certain ions to pass, and an individual cell has ion channels with various ion selectivity. In the context of studies of biologic cell membranes, the term ion selectivity refers to the ability of all cell membranes to distinguish between various ions such as Na+, K+, Ca2+, and Cl. We will focus on Na+, K+, and Ca2+ channels. All of these voltage-gated channels are made up of one or more pore-forming α subunits and variable numbers of accessory subunits, denoted β, γ, and so on. The α subunits determine ion selectivity and mediate the voltage-sensing functions of the channel. This ion selectivity involves specific pores or channels in the cell membrane, with certain channels being specific for certain ions and the opening or closing (gating) of channels depending on conditions and various interactions with ligands binding to receptors. These receptors are in some cases part of the channel itself and in other cases neighboring entities that control channel dynamics. The selectivity of an ion channel can be “gated,” with the channel effectively opened or closed, and ion channels are said to be voltage gated or ligand gated, depending on how the change in selectivity is provoked. A summary of the most studied CNS ion channels is presented in Table 49-2.

Genesis of Fast Sodium Action Potentials and Properties of Sodium Channels

Neuronal cells use a single type of signaling based on all-or-nothing action potentials. Sodium action potentials such as those recorded in axons or cell bodies are relatively invariant in normal tissue, and thus the shape and duration of these electrical signals do not change significantly within neuronal subtypes in the nervous system. Calcium action potentials are similarly predictable, but the underlying ionic mechanism can be rather complex, depending on the cell type and its topographic location within the cell (see later). The terms sodium action potential and calcium action potential refer to the initial (depolarizing) phase of these rapid changes in membrane polarity. Although genetic or molecular alteration of INa and ICa can significantly affect neuronal firing and, ultimately, central and peripheral nervous system neurophysiology, it must be remembered that gross changes in neuronal excitability may also result by altering the repolarization phase of individual action potentials.

Action potentials have a characteristic shape once a certain threshold is reached. In normal tissue, stereotyped electrical events follow the initial depolarization (Fig. 49-4). The sequence can be described as follows: (1) RMP moves from an initial negative value (−65 to −80 mV for most neurons) toward the so-called threshold for activation of sodium channels (around −40 mV). This change can be slow and may occur spontaneously as a result of fluctuations in RMP; the threshold is reached rapidly when the initial depolarization is triggered by a synaptic potential (or a summation of synaptic potentials). (2) After reaching the threshold value, an extremely rapid (1 to 2 msec) depolarization occurs because of opening of sodium channels and massive influx of sodium ions into the cell. (3) Termination of the “upstroke” phase depends on a combination of factors: voltage– and time-dependent inactivation of INa and a concomitant decrease in the driving force for sodium occurring in parallel with voltage-dependent activation of IK. (4) Further increases in the permeability to potassium ions restore the pre–action potential RMP values and force the membrane toward EK. (5) Under most circumstances, an undershoot of cell resting voltage occurs (a few millivolts from RMP) as a result of residual activation of IK and the contribution of the electrogenic sodium-potassium adenosine triphosphatase (Na+,K+-ATPase) extruding 3 Na+ ions in exchange for 2 K+. The return to pre–action potential voltage favors the so-called removal of inactivation, a necessary step that allows a subsequent cycle of depolarization-induced action potential firing. This stereotyped and relatively simple sequence of events is typically recorded in axons; other more complex interactions of various INa and IK values may lead to slightly different voltage profiles.

From a functional standpoint, it is important to remember that the genesis of fast sodium action potentials is a hallmark of neuronal function, to the degree that during neurophysiologic recordings, the presence or absence of Na+ spikes is frequently used to determine the neuronal or glial cell type. Recently, this notion has been challenged, and glial “action potentials” have been reported with increasing frequency. These responses, however, appear to usually be associated with pathologic conditions (brain tumors, epilepsy), and the old perception that neuronal cells are the exclusive tenants of sufficient INa density to promote active responses is still generally accepted. Within the same neuronal cell, Na+ channels involved in the generation of action potentials can be located heterogeneously, and it is common for clusters of channels to be located at specific and crucial membrane segments. The most commonly encountered clustering of Na+ channels occurs at the node of Ranvier of myelinated axons, but clustering also occurs at synaptic contacts, dendrites, and cell bodies, in proximity to the initial segment of axons.

Early pharmacologic studies attempting to elucidate the ionic correlates of the action potential greatly benefited from the availability of naturally occurring toxins that specifically and powerfully block INa. The magic bullet for sodium channels was tetrodotoxin (TTX),5 a lethal poison produced by a selected species of puffer fish (Fugu). TTX is present in large quantities in the liver and female genital organs, but the real culinary treat is the testes of this hermaphroditic animal. Separation of the toxic portions is paramount to survival of the consumer. The sea anemone toxin (ATX) and α-scorpion toxin are also powerful blockers of INa. TTX and other natural toxins block the external portion of the ion channel. Interestingly, although most sodium channels are blocked by micromolar concentrations of TTX, genetic ablation of one amino acid (Tyr or Phe) in the sequence of the ion channel protein confers relative resistance to TTX because binding of the toxin is restricted to this region.6 The amazing specificity of TTX binding suggested that normally occurring variation in the amino acid sequence of ion channels may prove to be an extremely important determinant of not only the pharmacologic properties of the channels but also their inactivation/activation/voltage dependency. In fact, although the poisonous actions of TTX are primarily due to direct blockade of Na+ fluxes through the channel’s pore, ATX and α-scorpion toxin bind to a portion responsible for channel inactivation. Mutations in these regions cause faulty inactivation, a condition linked to neuropathogenesis.

The biophysical correlates of Na+ channel function are well understood today. The general scheme of

image

explains the properties of whole-cell INa recorded from neurons. The voltage dependency of each process justifies the initial depolarization required to promote the opening of channels; the consequent depolarization induced by sodium current promotes further opening of channels, the process being terminated by time- and voltage-dependent closure of the channels. The passage from closed to open (and visa versa) is referred to as activation (deactivation), whereas the passage from open to inactivated is called inactivation. Removal of inactivation occurs when the channel returns to its closed state.

From a structural point of view, Na+ channels are constituted by αβ1β2 heterotrimers, often with four repeated domains each with six-membrane–spanning subunits. The voltage sensor is located on the fourth transmembrane domain. Different subunits are represented differently in the central and peripheral nervous systems. According to the recent literature, the following tissue-specific localization and pharmacology can be derived: for α subunits, SCN1A is primarily expressed in brain tissue and is blocked by TTX and saxitoxin (SXT). The most abundant brain α subunit is encoded by SCN2A1; this subunit is also found in peripheral nerves, in the initial axonal segment, and in the nodes of Ranvier (TTX and SXT sensitive). SCN2A2 and SC2NA3 are predominantly expressed in the brain, whereas SCN4A encodes channels found in muscle. In addition to TTX, these channels are sensitive to ω-conotoxins (GIIIA, GIIIB, GIIIC). Mutations of these channels are responsible for hyperkalemic periodic paralysis, paramyotonia, and myotonia. The SCN5A subunit is expressed in the heart and in denervated skeletal muscle; in the heart, these channels are responsible for upstroke of the action potential. These subunits are resistant to blockade by TTX and SXT; mutations of these genes are involved in subtypes of long QT syndrome, a cardiac condition that is sometimes associated with epileptic seizures. SCN6A (uterus), SCN10A, SCN11A, and SCN9A (dorsal root ganglion) are all sensitive to TTX. The molecular nature of glial sodium channels and the contribution of a specific α subunit are not fully understood at present. β Subunits are bound covalently to α subunits and provide inactivation kinetics to Na+ channels. Mutations of these SCN1B and SCN1A subunits have been linked to generalized epilepsy with febrile seizures (see Catterall and colleagues7 and Gambardella and Marini8 for review of voltage-gated ion channel–related hereditary diseases).

The fact that even slight mutations cause profound deficits in sodium channel function and that these mutations result in neurological diseases leads to the hypothesis that replacement of defective channels by gene therapy may repristinate the loss of function caused by the initial genetic deficit. If a single gene mutation is responsible (as in cystic fibrosis and in a number of neurological disorders) and if the tissue to be transfected with the repair mechanism is accessible (such as skeletal muscle), it is possible that viral delivery of genetic products may alleviate the consequences of faulty ion channels. A positive outcome with this approach is somehow dependent on the pathogenesis of the disease itself. If the observed deficit is the consequence solely of the inherited mutation, replacement by a normal genotype is likely to be successful. If, however, the initial deficit has led to extensive rearrangement of neuronal connections or if the deficit causes extensive neuronal cell death (such as in epilepsy), it is unclear whether restoring normal function by simple gene replacement may be beneficial. It is also worth remembering that although a small fraction of neurological disorders are clearly imputable to a single gene mutation affecting a particular ion channel, the most common forms of disease result from a complex interaction of initial genotypic changes followed by adaptive responses, including apoptosis or necrosis. Finally, given that mutations of crucial ionic mechanisms such as INa affect cardiac, neuronal, and muscular function, it is possible that a large number of mutations are nonvital and that the vital mutations are possible because of redundancy of gene expression or compensatory overexpression of similar ion channel proteins.

Phenotypic changes caused by relatively minor alterations in ion channel gating sometimes become clinically relevant when concomitant deficits not necessarily associated with action potentials are present. This is the case, for example, in mutations of the SCN4A subunit leading to hyperkalemic periodic paralysis. For the paralytic symptoms to occur, patients must concomitantly experience variations in plasma potassium (by either K+ intake or exercise followed by rest). This leads to opening of Na+ channels that switch to a non-inactivating mode, thereby leading to the development of a persistent inward Na+ current. The ensuing depolarization of muscle membrane will further increase [K+]out via loss through voltage-dependent K+ channels and thus aggravate the initial trigger. Furthermore, the persistent depolarization causes inactivation of normal Na+ channels, which leads to rapid loss of tissue excitability and paralysis. This example demonstrates the complex interactions between normal and abnormal ion channels expressed in a certain cell type, the importance of the extracellular milieu in biophysical signaling via ion channels, and the difficulties associated with the diagnosis of altered ion channel phenotypes.

Calcium Action Potentials and Calcium Channels

The mechanism of calcium action potentials is somewhat different, but it follows the general principles of threshold for activation and rapid gating mechanisms. As is the case with sodium channels, Ca2+ channels are distributed heterogeneously in the CNS, and even within the same cells different subspecies of Ca2+ channels may be found. This inhomogeneous expression is functionally significant in that it allows Ca2+ influx to perform several different cellular tasks, including depolarization of dendrites and propagation of signals to the cell body, synaptic release of neurotransmitter, contraction, and second messenger function. As for sodium channels, membrane depolarization is the most common trigger for opening of channels; the kinetic properties of Ca2+ channels, however, are characterized by longer time constants. This kinetic behavior underlies the different durations of neuronal/axonal action potentials (INa, 1 to 3 msec) versus cardiac action potentials (several hundreds of milliseconds; large inward currents carried by ICa).* An additional difference between Na+ and Ca2+ action potentials (or the underlying ion channel mechanisms) is the fact that the threshold for ICa activation varies greatly between different Ca2+ channel families. Low-threshold Ca2+ channels (or low voltage activated) are also characterized by relatively rapid opening and closing and are also referred to as “T-type” (transient) currents. High-threshold Ca2+ channels (or high voltage activated) can be further subdivided into neuronal (N), L, and P types. The pharmacologic properties of the calcium channel families are equally complex (Table 49-3).9,10 Although sodium action potentials are typically triggered by depolarization and terminated by depolarization (inactivation of INa and voltage-dependent activation of IK), an additional mechanism is involved in repolarization: activation by intracellular Ca2+ channels that act as powerful terminators of cell depolarization. Ca2+ channels are modulated by important intracellular signals such as cyclic adenosine monophosphate (cAMP; cAMP-dependent protein kinases) and G proteins. These modulatory signals arise from receptor stimulation, thus coupling the activity of postsynaptic (or presynaptic in the case of presynaptic receptors) Ca2+ channels to the activity of neighboring cells.

Ca2+ channels contain four or five distinct subunits: α subunits display different tissue and peptide specificity. They are constituted by transmembrane-spanning proteins and act in both voltage sensor and selectivity filter capacities. The dihydropyridines verapamil and nifedipine bind to α1