Canal Calcio T

46
Molecular Physiology of Low-Voltage-Activated T-type Calcium Channels EDWARD PEREZ-REYES Department of Pharmacology, University of Virginia, Charlottesville, Virginia I. Introduction 118 II. Electrophysiology of Native T-type Currents 120 A. Low-theshold calcium spikes 120 B. Neuronal 122 C. Heart 125 D. Kidney 127 E. Smooth muscle 128 F. Skeletal muscle 129 G. Sperm 129 H. Endocrine tissues 130 I. Cell lines 131 J. Single-channel recordings 131 III. Regulation 134 A. Hormonal inhibition 134 B. Hormonal stimulation 135 C. Guanine nucleotides 136 D. Protein kinases 136 E. Voltage 137 IV. Molecular Cloning of T-type Channels 138 A. Cloning 138 B. Structure 140 C. Distribution 141 D. Electrophysiology of recombinant channels 141 E. Auxiliary subunits? 146 V. Pharmacology 147 A. Antihypertensives 147 B. Antiepileptics 148 C. Anesthetics 149 D. Antipsychotics 149 VI. Conclusions 150 A. Physiological roles 150 B. Future directions 150 Perez-Reyes, Edward. Molecular Physiology of Low-Voltage-Activated T-type Calcium Channels. Physiol Rev 83: 117–161, 2003; 10.1152/physrev.00018.2002.—T-type Ca 2 channels were originally called low-voltage-activated (LVA) channels because they can be activated by small depolarizations of the plasma membrane. In many neurons Ca 2 influx through LVA channels triggers low-threshold spikes, which in turn triggers a burst of action potentials mediated by Na channels. Burst firing is thought to play an important role in the synchronized activity of the thalamus observed in absence epilepsy, but may also underlie a wider range of thalamocortical dysrhythmias. In addition to a pacemaker role, Ca 2 entry via T-type channels can directly regulate intracellular Ca 2 concentrations, which is an important second messenger for a variety of cellular processes. Molecular cloning revealed the existence of three T-type channel genes. The deduced amino acid sequence shows a similar four-repeat structure to that found in high-voltage-activated (HVA) Ca 2 channels, and Na channels, indicating that they are evolutionarily related. Hence, the 1 -subunits of T-type channels are now designated Ca v 3. Although mRNAs for all three Ca v 3 subtypes are expressed in brain, they vary in terms of their peripheral expression, with Ca v 3.2 showing the widest expression. The electrophysiological activities of recombinant Ca v 3 channels are very similar to native T-type currents and can be Physiol Rev 83: 117–161, 2003; 10.1152/physrev.00018.2002. www.prv.org 117 0031-9333/03 $15.00 Copyright © 2003 the American Physiological Society on October 7, 2011 physrev.physiology.org Downloaded from

Transcript of Canal Calcio T

Page 1: Canal Calcio T

Molecular Physiology of Low-Voltage-ActivatedT-type Calcium Channels

EDWARD PEREZ-REYES

Department of Pharmacology, University of Virginia, Charlottesville, Virginia

I. Introduction 118II. Electrophysiology of Native T-type Currents 120

A. Low-theshold calcium spikes 120B. Neuronal 122C. Heart 125D. Kidney 127E. Smooth muscle 128F. Skeletal muscle 129G. Sperm 129H. Endocrine tissues 130I. Cell lines 131J. Single-channel recordings 131

III. Regulation 134A. Hormonal inhibition 134B. Hormonal stimulation 135C. Guanine nucleotides 136D. Protein kinases 136E. Voltage 137

IV. Molecular Cloning of T-type Channels 138A. Cloning 138B. Structure 140C. Distribution 141D. Electrophysiology of recombinant channels 141E. Auxiliary subunits? 146

V. Pharmacology 147A. Antihypertensives 147B. Antiepileptics 148C. Anesthetics 149D. Antipsychotics 149

VI. Conclusions 150A. Physiological roles 150B. Future directions 150

Perez-Reyes, Edward. Molecular Physiology of Low-Voltage-Activated T-type Calcium Channels. Physiol Rev 83:117–161, 2003; 10.1152/physrev.00018.2002.—T-type Ca2� channels were originally called low-voltage-activated(LVA) channels because they can be activated by small depolarizations of the plasma membrane. In many neuronsCa2� influx through LVA channels triggers low-threshold spikes, which in turn triggers a burst of action potentialsmediated by Na� channels. Burst firing is thought to play an important role in the synchronized activity of thethalamus observed in absence epilepsy, but may also underlie a wider range of thalamocortical dysrhythmias. Inaddition to a pacemaker role, Ca2� entry via T-type channels can directly regulate intracellular Ca2� concentrations,which is an important second messenger for a variety of cellular processes. Molecular cloning revealed the existenceof three T-type channel genes. The deduced amino acid sequence shows a similar four-repeat structure to that foundin high-voltage-activated (HVA) Ca2� channels, and Na� channels, indicating that they are evolutionarily related.Hence, the �1-subunits of T-type channels are now designated Cav3. Although mRNAs for all three Cav3 subtypes areexpressed in brain, they vary in terms of their peripheral expression, with Cav3.2 showing the widest expression. Theelectrophysiological activities of recombinant Cav3 channels are very similar to native T-type currents and can be

Physiol Rev

83: 117–161, 2003; 10.1152/physrev.00018.2002.

www.prv.org 1170031-9333/03 $15.00 Copyright © 2003 the American Physiological Society

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 2: Canal Calcio T

differentiated from HVA channels by their activation at lower voltages, faster inactivation, slower deactivation, andsmaller conductance of Ba2�. The Cav3 subtypes can be differentiated by their kinetics and sensitivity to block byNi2�. The goal of this review is to provide a comprehensive description of T-type currents, their distribution,regulation, pharmacology, and cloning.

I. INTRODUCTION

Rises in intracellular calcium trigger a variety of pro-cesses including muscle contraction, chemotaxis, geneexpression, synaptic plasticity, and secretion of hor-mones and neurotransmitters. Although there are manychannels and pumps involved in controlling intracellularCa2� levels, voltage-gated Ca2� channels play a key rolein this process (50). Calcium is not only an importantsecond messenger, but its entry can also depolarize theplasma membrane, and thereby activate other voltage-gated ion channels. This property is especially importantfor neuronal T-type channels, which can generate low-threshold spikes that lead to burst firing and oscillatorybehavior (175, 377). This activity is especially prominentin the thalamus, where it plays an important role in sen-sory gating, sleep, and arousal (277). The term thalamo-

cortical dysrhythmias has been coined to describe patho-logical changes in these oscillations, and they have beenimplicated in a wide range of neurological disorders in-cluding absence epilepsy, the tremor associated with Par-kinson’s disease, tinnitus, neuropsychiatric disorders, andneurogenic pain (186, 242).

The ability of neurons to fire low-threshold Ca2�

spikes suggested the existence of low-voltage-activated(LVA) Ca2� channels. Since 1975 it has been recognizedthat there were at least two distinct types of Ca2� channel(reviewed in Ref. 153). Hagiwara, Ozawa, and Sand, usingtwo-microelectrode voltage-clamp recordings from star-fish eggs, found channels that were activated after smalldepolarizations of the membrane, LVA, and other chan-nels that required larger depolarizations of the membrane,high voltage activated (HVA) (154). LVA currents inacti-vated faster, more completely, and at more negative mem-brane potentials than HVA currents. Similarly, LVA andHVA currents were described in the marine pilewormNeanthes (123). Of historical note, the first recordings ofmammalian LVA channels were made by Moolenar andSpector in 1978 (290), who used the two-microelectrodevoltage-clamp technique on the mouse neuroblastomacell line N1E-115. However, these authors only observedone type of Ca2� channel, which in 10 mM Ca2� activatedat �55 mV and peaked at �20 mV. Subsequent studieshave shown that N1E-115 cells provide a convenient sys-tem to study the properties of native T-type currents (seesect. III). Using quinidine to block K� currents, Fishmanand Spector reported in 1981 (120) on the existence oftwo types of mammalian Ca2� channel. One type acti-vated at �50 mV, peaked at �20 mV, and inactivated

rapidly (� �15–30 ms). The second type peaked at 0 mVand inactivated slowly (� �2,000 ms). The existence ofLVA currents was firmly established by the work of manygroups, including those headed by Kostyuk (115, 420),Carbone and Lux (62), Armstrong (12), Bean (32), Feltz(54), and Tsien (303, 306). Whole cell patch-clamp record-ings showed that depolarizations in the �60 to �20 mVrange elicited rapidly inactivating currents, while higherdepolarizations elicited noninactivating currents. Plots ofthe current versus test potential (I-V) showed two com-ponents, with the LVA component appearing as a humpon the back of a larger HVA component. LVA currentsalso turned off, or deactivated, slower than HVA currents,producing slow tail currents (62). This property was stud-ied in great detail in GH3 cells, where slowly deactivating(SD) tail currents were found to activate at lower thresh-olds than fast deactivating (FD) currents (12). Tsien andcolleagues (306) extended the classification of dorsal rootganglion (DRG) channels and proposed that these chan-nels be called T type for transient, L type for long lasting,and N type for neither T nor L type. LVA currents werealso found to be resistant to rundown, unlike HVA cur-rents that required cAMP, ATP, and Mg2� in the pipettefor stability (115). T- and L-type channels were also de-scribed in cardiac myocytes from guinea pig ventricle anddog atrium and could be distinguished by their voltagedependence, kinetics, single-channel currents, pharma-cology, and conductance of Ca2� and Ba2� (32, 303). Theobservation that T-type currents inactivate at lower mem-brane potentials than L-type currents led to the develop-ment of an assay to separate these two components (32).The activity of both channels was recorded from a well-hyperpolarized potential such as �90 mV, then the activ-ity of L-type channels was recorded at a potential whereT-type channels are inactivated and L type are not (typi-cally �50 mV), then the L-type currents are subtractedfrom the T- plus L-type currents to isolate the T-typecurrent. Additional methods, such as tail current analysis,are required to isolate T-type currents in most other tis-sues due to the presence of HVA channels that also inac-tivate at �50 mV. Another possible contamination is volt-age-gated Na� currents that appear to conduct Ca2�

(ICa,TTX); however, these can be differentiated by theirsensitivity to tetrodotoxin (TTX) (162). Hallmark featuresof T-type channel electrophysiology are as follows: 1)they begin to open after small depolarizations of theplasma membrane (LVA); 2) their currents during a sus-tained pulse are transient; 3) they close slowly uponrepolarization of the membrane, generating a SD tail cur-

118 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 3: Canal Calcio T

rent; 4) they have a tiny, and equivalent, single-channelconductance of Ba2� and Ca2�; 5) they are relativelyinsensitive to dihydropyridines; and 6) their steady-stateinactivation occurs over a similar voltage range as activa-tion. In fact, T-type channels display a window current,i.e., there is a small range of voltages where T-type chan-nels can open, but do not inactivate completely. Thisproperty may be particularly important in controlling in-tracellular Ca2� levels (45). In conclusion, studies onnative channels have provided a toolkit for separatingCa2� channel subtypes, and these tools have proven use-ful in the characterization of both native and recombinantchannels.

T-type channels are expressed throughout the body,including nervous tissue, heart, kidney, smooth muscle,sperm, and many endocrine organs. These channels havebeen implicated in variety of physiological processes in-cluding neuronal firing, hormone secretion, smooth mus-cle contraction, myoblast fusion, and fertilization. In gen-eral, the electrophysiological properties of T-typecurrents recorded from various cell types are similar, butdifferences have been noted in how they inactivate and intheir pharmacology. This heterogeneity can be explainedin part by the existence of three T-type channels that areencoded on separate genes (90, 228, 324).

Voltage-gated Ca2� channels have been classifiedby their electrophysiological and pharmacologicalproperties, and more recently by their amino acid se-quence identity. There are at least 10 genes encoding�1-subunits of voltage-gated Ca2� channels (Fig. 1).Alignment of their deduced amino acid sequences sug-

gests that gene duplication and divergence of an ances-tral Ca2� channel gene gave rise to LVA and HVAsubfamilies. Further duplication of the HVA gene gaverise to Cav1 and Cav2 subfamilies. This duplicationmust have occurred well over 500 million years ago,since the nematode Caenorhabditis elegans containsone member of each subclass. Eventually the Cav1subfamily evolved into four genes, while both the Cav2and Cav3 subfamilies evolved into three. Cloning andexpression studies have established that the Cav3 fam-ily encodes LVA T-type channels (sect. IV), while theCav1 subfamily encodes L-type channels (although ex-pression of Cav1.4 has not been reported yet). Thesechannels play a critical role in excitation-contractioncoupling in cardiac, skeletal, and smooth muscle. Infact, the Cav1.1 gene has evolved beyond a Ca2� chan-nel, and its physiological role in skeletal muscle is as avoltage sensor, coupling membrane depolarization di-rectly to calcium release channels of the sarcoplasmicreticulum (SR) (340). In contrast, cardiac and smoothmuscle contraction requires influx of extracellular Ca2�

through Cav1.2 channels. In heart, this influx activates alarger release of Ca2� from intracellular pools via amechanism called calcium-induced calcium release(CICR) (37). Cav1.2 channels are an important thera-peutic target in the control of blood pressure and car-diac rhythm. The physiological roles of Cav1.3 channelshave been difficult to discern since pharmacologicaltools have not been described that can separate itsactivity from Cav1.2, and because these channels ap-pear to be coexpressed in a number of tissues. Incontrast, the activity of Cav2 family members can beselectively blocked by peptide toxins. Splice variationof the Cav2.1 gene results in channels that differ in theirsensitivity to the spider toxin �-Aga-IVA and thereforeencode both P- and Q-type channels (55). The Cav2.2gene encodes N-type channels, which are characterizedby their irreversible and potent block by the snail toxin�-conotoxin-GVIA. Studies with these toxins indicatethat Cav2.1 and Cav2.2 channels mediate the Ca2� influxinto presynaptic terminals that trigger neurotransmitterrelease. In contrast to these channels, it has been dif-ficult to determine which native Ca2� current is carriedby Cav2.3 channels. Recombinant Cav2.3 channels re-semble T-type channels in their sensitivity to nickel andtheir voltage dependence of inactivation, leading to theearly suggestion that they might encode T-type chan-nels (376). However, Cav2.3 channels require strongerdepolarizations for channel opening, i.e., are HVA chan-nels, and they deactivate 10-fold faster than T-typechannels (228). Studies with SNX-482, a tarantula toxinthat selectively blocks recombinant Cav2.3 channels,suggests that these channels mediate R-type currents insome tissues (299).

FIG. 1. Evolutionary tree of voltage-gated Ca2� channels. Tree isbased on an alignment of the putative membrane-spanning regions andpore loops of the human channels. Alignment of the full-length se-quences yields a similar pattern, although all the branch points areshifted to the left due to inclusion of nonconserved sequences. Low-voltage-activated (LVA) channels appear to have diverged from an an-cestral Ca2� channel before the bifurcation of the high-voltage-activated(HVA) channel family into Cav1 and Cav2 subfamilies. (Adapted fromPerez-Reyes E, Cribbs LL, Daud A, Yang J, Lacerda AE, Barclay J,Williamson MP, Fox M, Rees M, and Lee J-H. Molecular characterizationof T-type calcium channels. In: Low Voltage-activated T-type Calcium

Channels, edited by Tsien RW, Clozel J-P, and Nargeot J. Chester, UK:Adis International, 1998, p. 290–305.)

T-TYPE CALCIUM CHANNELS 119

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 4: Canal Calcio T

Electrophysiological studies of recombinant chan-nels show that Cav3.1 (formerly �1G) and Cav3.2 (�1H)have similar activation and inactivation kinetics, but canbe differentiated by their recovery from inactivation andsensitivity to block by nickel (209, 230, 351). The kineticproperties of these �1-subunits closely resemble nativeT-type channels. In contrast, recombinant Cav1 and Cav2channels require auxiliary subunits for normal gating be-havior. This suggests that native T-type channels may beformed by a single �1-subunit. Cav3.3 (�1I) channels alsogenerate LVA currents; however, these currents activateand inactivate much more slowly than T-type channels.Native currents with these properties have only beendescribed in a limited number of studies (19, 99, 177, 316,398). Although there is no drug that is highly selective(�10-fold) for T-type over other ion channels, block ofT-type channels appears to contribute to the therapeuticusefulness of antihypertensives, antiepileptics, anesthet-ics, and possibly antipsychotics. Therefore, T-type chan-nels provide a novel target for drug development.

II. ELECTROPHYSIOLOGY OF NATIVE

T-TYPE CURRENTS

A. Low-Threshold Calcium Spikes

Low-threshold calcium spikes (LTS) have been de-scribed in slices and in isolated neurons from a variety ofbrain nuclei such as inferior olive (243, 244), thalamicrelay (96, 183), medial pontine reticular formation (146),lateral habenula (437), septum (10), deep cerebellar nu-clei (241), CA1-CA3 of the hippocampus (163, 307), asso-ciation cortex (122), paraventricular (399) and preopticnuclei of the hypothalamus (385), dorsal raphe (60), glo-bus pallidus (296), and the subthalamic nucleus (40).Typically the spike is crowned by a burst of action poten-tials (Fig. 2). Addition of TTX to block voltage-gated Na�

channels abolishes the fast action potentials, effectivelyisolating the slowly activating and inactivating LTS. Nu-merous studies have shown that the LTS is mediated by a

FIG. 2. Low-threshold Ca2� spikes generate burst firing. A: many neurons can generate two distinct patterns of actionpotential firing in response to a depolarizing stimulus. Regular, or tonic, firing is elicited when the neuron is depolarizedfrom a resting membrane potential near �55 mV. In contrast, when the membrane potential is below �70 mV, the samedepolarizing stimulus triggers a high-frequency burst of action potentials. [From Huguenard JR. Low-voltage-activated(T-type) calcium-channel genes identified. Trends Neurosci 21: 451–452, 1998, with permission from Elsevier Science.]B: a representative example of currents that generate burst firing. Depolarization of the plasma membrane by hyper-polarization-activated current (Ih) leads to activation of T-type currents (IT), and a second phase of depolarization calledthe low-threshold Ca2� spike. Riding on top of the low-threshold Ca2� spike are a burst of Na� spikes mediated by fastvoltage-gated Na� channels. High-threshold Ca2� and K� currents can also be activated by the low-threshold calciumspikes. Ca2� entry during the burst leads to activation of Ca2�-activated K� currents, which in combination withvoltage-gated K� channels repolarize the membrane. (From Bal T and McCormick DA. Synchronized oscillations in theinferior olive are controlled by the hyperpolarization-activated cation current Ih. J Neurophysiol 77: 3145–3156, 1997.)C: thalamic neurons of transgenic mice lacking expression of Cav3.1 do not fire bursts. Current-clamp recordings arefrom neurons held at �60, �70, or �80 mV, then depolarized or hyperpolarized by current injections of varyingmagnitudes as indicated below the set of traces. When the resting membrane potential is �60 mV, a depolarizing stimulustriggers tonic firing in both wild-type (�/�) and transgenic (�/�) animals. When the membrane potential (Vm) is loweredto �70 mV, a hyperpolarizing stimulus triggers burst firing in neurons from wild-type, but not transgenic animals. Whenthe resting membrane potential is �80 mV, a depolarizing stimulus only triggers burst firing in neurons from wild-typeanimals. (From Kim D, Song I, Keum S, Lee T, Jeong MJ, Kim SS, McEnery MW, and Shin HS. Lack of the burst firingof thalamocortical relay neurons and resistance to absence seizures in mice lacking �1G T-type Ca2� channels. Neuron

31: 35–45, 2001, with permission from Elsevier Science.)

120 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 5: Canal Calcio T

Ca2� conductance; it is not observed in Ca2�-free externalsolutions, and it can be blocked by divalent cations suchas Co2� and Ni2�. Definitive proof that thalamic LTS aremediated by T-type channels was provided by studies ontransgenic mice, where knockout of the Cav3.1 gene abol-ished these spikes and burst firing (Fig. 2C) (206).

With a few notable exceptions (7, 191), LTS cannotbe triggered by depolarization of the neuron from theresting membrane potential, which is typically between�60 and �65 mV. However, LTS can be observed after ahyperpolarizing pulse is delivered. This process has beencalled “deinactivation” and is due to recovery of channelsfrom inactivation. Table 1 lists a number of studies wherethe voltage and time dependence for recovery of thesespikes has been examined in detail. LTS began to appearafter the neuronal membrane was hyperpolarized below�69 mV, and full-amplitude spikes were observed whenthe membrane was below �73 mV. The rate of recoverywas measured by varying the duration of a hyperpolariz-ing pulse, then testing for the presence of an LTS uponreturn to the resting membrane potential. These studiesfound that the LTS deinactivated relatively quickly, withhalf of it recovering in �100 ms and full recovery after 200ms. Therefore, one physiological role of the LTS is as apacemaker. Although increases in intracellular concentra-tions of Ca2� act as a second messenger for a variety ofprocesses, in this case the important property is thatinflux of the divalent cation leads to direct depolarizationof the membrane. The amplitude of this depolarizationwas typically 25 mV (Table 1), which raised the mem-brane potential to approximately �40 mV. Since thethreshold of many voltage-gated Na� channels is around�55 mV, the LTS can trigger a burst of action potentials.

These Na�-dependent action potentials are brief (1–2 ms)and of high amplitude (�80 mV), and in some thalamicneurons of very high frequency (�400 Hz). High-thresholdCa2� channels will also begin to open at potentials morepositive than �40 mV (385). This will lead to more influxof Ca2� and, if present, activation of Ca2�-dependent K�

channels, which can lead to an afterhyperpolarization(AHP) (40, 244). Low-threshold channels can recoverfrom inactivation during the AHP and could begin to fireagain as the membrane potential returns to its restinglevel. Hyperpolarizations can also activate Ih channels,which allow the influx of both Na� and K�, thereby depo-larizing the cell and directly triggering another LTS (Fig. 2).In this scenario Ih are the primary pacemaker current (254).Clearly there are a variety of ionic conductances that medi-ate pacemaker activity, and it is an oversimplification toimplicate a single channel as the mediator of a pacemakercurrent. This is especially true in the heart (61).

Another way that low-threshold channels are in-volved in generating oscillations includes reciprocal con-nections with an inhibitory interneuron. Due to its impor-tant role in controlling brain activity, the best studied ofthese circuits involves the GABAergic neurons of thethalamic reticular nucleus that regulate the activity ofthalamic relay neurons (377). Interestingly, the firing pat-terns of thalamic relay neurons vary with the state ofconsciousness (277, 378). In the awake state and duringrapid-eye-movement sleep, they fire in tonic mode: theresting membrane potential is relatively depolarized, theLTS is inactivated, and excitatory inputs faithfully triggeraction potentials. In deep sleep, or during thalamocorticaldysrhythmias, these neurons fire in an oscillatory modecalled burst firing; the resting membrane potential is hy-

TABLE 1. Properties of neuronal low-threshold Ca2� spikes

Brain Region

RestingMembrane

Potential, mVThreshold,

mVAmplitude,

mV

Deinact VThreshold,

mV

Deinact VMidpt,

mV

Deinact TimeMidpt,

ms

Deinact TimeMax,ms

ReferenceNos.

Subthalamic nucleus �50 20 �78 80 120 40Septal nucleus �60 14 �63 �77 70 175 10Hippocampus, CA1–CA3 �60 �63 11 307Thalamic nuclei �64 �54 23 �55 �65 60 170 183Lateral thalamic nuclei �60 �65 20 �69 50 96Lateral geniculate nuclei 30 �75 �66 457Lateral geniculate nuclei �60 �68 35 �64 �72 456Lateral habenular nucleus �62 �65 30 �65 �68 437Paraventricular nucleus �66 �60 40 200 300 399Medial preoptic nucleus �60 34 �73 �77 120 400 385Dorsal raphe nucleus �67 �60 �70 400 800 60Cerebellar nuclei �57 24 �72 �78 241Cerebellar nuclei �58 �65 20 �72 �77 200 400 2Inferior olivary nucleus �64 �65 35 �70 �75 40 75 243, 244Hypoglossal motoneurons ��60 �65 10 �83 200 400 421

The properties of low-threshold Ca2� spikes (LTS) were measured using current-clamp recordings. Threshold was defined by the voltagewhere a depolarizing pulse triggers a LTS. The amplitude of the LTS is shown and does not include the fast Na�-dependent action potential. Thevoltage and time dependence for deinactivation (recovery) of the LTS spike is shown. Deinact., deinactivation; midpt, midpoint; CA1–CA3,hippocampal pyramidal neurons.

T-TYPE CALCIUM CHANNELS 121

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 6: Canal Calcio T

perpolarized, allowing full-amplitude low-threshold spik-ing (Fig. 2A). Such firing is thought to underlie spike-wavedischarges observed during absence seizures (176). Sim-ilar patterns of brain activity have been detected in avariety of neurological disorders by magnetoencephalog-raphy and have been termed thalamocortical dysrhyth-mias (242).

B. Neuronal

1. Isolated neurons

LVA T-type currents have been characterized usingwhole cell voltage-clamp recording of neurons isolatedfrom a variety of brain regions. Very large T-type currents(�1 nA) have been recorded from cholinergic neurons ofthe basal forebrain (6), from relay neurons of the ventro-basal thalamus (222), and from sensory neurons of dorsalroot ganglia (436). Prominent T-type currents have beenfound in neurons from many brain regions (Table 2). I-Vrelationships are typically measured by holding the cell at�90 mV, and then currents are elicited by depolarizingpulses that increase in 10-mV increments. T-type currentsbegin to activate when the membrane is depolarizedabove �60 mV. At threshold potentials, the currents ac-tivate relatively slowly, requiring many milliseconds toreach peak, and also inactivate relatively slowly. At higherpotentials both activation and inactivation are faster. Thisproduces a stereotypical pattern in the family of currenttraces where successive records cross each other, remi-niscent of voltage-gated Na� channels. In general, HVAchannels do not produce this criss-crossing pattern be-cause inactivation rates are much slower than activationrates (334). I-V curves commonly show two components,with the LVA currents peaking at �30 mV while the HVAcurrents peak around 0 mV. In many neurons, T-typecurrents can be measured in isolation with test pulses to�30 mV and below. However, at higher potentials bothLVA and HVA currents are activated, complicating esti-mates of the LVA activation curve. Both I-V curves showa peak then decrease at more positive potentials, which isdue to a reduction of the electrochemical gradient forcalcium ions as the test potential approaches the theoret-ical reversal potential (�100 mV; see Fig. 7 for example).Therefore, the fraction of channels activated by a testpulse continues to increase beyond the peak of the I-Vcurve. This explains why estimates of the midpoint ofchannel activation (V0.5) derived by normalizing the cur-rent at each test potential to the maximum current ob-served (I/Imax) produces values that are more negative(�50 mV) than estimates that take into account changesin driving force (�41 mV). To circumvent this problem,activation curves have also been estimated by measuringthe amplitude of slowly deactivating tail currents at aconstant voltage. Most of these studies used a test pulse

of constant duration (isochronal), which assumes that therates of channel activation and inactivation are voltageindependent. However, this is not the case, so this methodcan underestimate channel activation at negative poten-tials where channels open slowly. It can also underesti-mate activation if channels inactivate during the testpulse. This problem can be overcome by varying theduration of the activating pulse, so that the tail current iselicited at the peak of the current (288). So far thismethod has only been applied to recombinant channels.Another experimental variable that affects the position ofthe activation curve is the concentration and choice ofcharge carrier. Millimolar concentrations of positivelycharged divalent cations can bind to surface charges onboth the channel and the membrane, thereby reducing theeffective transmembrane voltage gradient (164). An effectcalled surface charge screening. Increasing external[Ca2�] from 2 to 10 mM shifts the apparent gating ofT-type channels by �10 mV (128).

A hallmark of T-type currents is that they are tran-sient; currents reach a peak then decay. In solutionswhere K� channels are blocked, this decay is due tochannel inactivation. The inactivation rate is measured byfitting the current trace with an exponential function andis typically reported with a � in milliseconds. Plots ofinactivation rate versus test potential reveal two phases:inactivation is slow at �60, gets faster between �50 and�30, then is constant above �30 mV. Activation rateshave a similar voltage dependence, leading to the notionthat channels must activate before inactivating, and thatthe inactivation process is essentially voltage-indepen-dent. Although most T-type currents inactivate relativelyrapidly (�30 ms), there are also reports of slowly inacti-vating currents (Table 2). This variability was interpretedas evidence for multiple T-type channel genes (179). Thishypothesis was supported by the cloning of T-type chan-nels that differed in this property; Cav3.1 and 3.2 displayfast inactivation, while Cav3.3 displays slowly inactivatingcurrents. Notably, Cav3.3 mRNA is expressed in the samebrain regions as the slow currents (177, 228, 393). Exclud-ing these cases, the average inactivation � from 22 studiesis 20 ms.

The voltage dependence of inactivation is also mea-sured at “steady-state” by holding the membrane at vary-ing potentials (prepulse), then assaying for available T-type channels with a test pulse to �30 mV (h�). Toapproximate steady-state, many studies have used a pre-pulse duration of 1 s, which is considerably longer thanthe time required to inactivate open channels. However,increasing the prepulse duration to 10 s results in a�10-mV shift in the apparent h� curve (53, 389). Despitethe use of different prepulses and concentration of diva-lent cations, there is good agreement between the 34studies shown in Table 2, and the average midpoint of theh� curve is �77 mV (Table 2). Somewhat surprisingly this

122 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 7: Canal Calcio T

TABLE 2. Properties of T-type currents in isolated neurons

TissueDivalent,

mMThreshold,

mVI–V Peak,

mVImax,pA

V0.5,mV

h�,mV

�inact,ms �recov, ms Nickel, �M

ReferenceNos.

Basal forebrainStriatum 5 Ca �60 �40 �90 �53M �88 288 170Accumbens 5 Ca �64 �33 �51 �43C �80 397 80Septum 2.5 Ca �54 �1,320 �40M �49 16 5 6Septum 2 Ba �70 �40 �500 �59M �84 147

Amygdala 10 Ca �60 �20 �20 �38C �70 300 188Cerebral cortex

Temporal 2 Ca �60 �30 �47 �86 400 50 (28%) 354Piriform 5 Ca �60 �30 �250 �44T �86 22 257

HippocampusDentate granule 2 Ca �50 �30 �224 �65 25 458L–Minterneurons

2 Ca �60 �30 �100 �48M �84 12 100 400 126

CA1 pyramidal 10 Ca �60 �30 �150 �46C �79 15 430 (63%) 3897,000

CA1, 3 pyramidal 2 Ba �57 �100 �76 25 820 400Thalamus

Ventrobasal 3 Ca �60 �25 �350 �52T �84 20 251 85Ventrobasal 3 Ca �60 �30 �280 �59T �81 30 350 200 (66%) 179Ventrobasal 5 Ca �70 �1,400 �96 15 353 222LGN 10 Ca �48 �15 �186 �67 36 299 500 (83%) 387LGN 5 Ca �60 �23 �200 �45M �64 20 615 50 (100%) 159LGN-interneuron 2 Ca �70 �110 �55M �81 20 218 318LGN-relay 2 Ca �70 �289 �64M �86 20 206 318Reticular 3 Ca �50 �10 �131 �49T �78 80 570 200 (54%) 179Reticular 2 Ca �70 �50 �128 �64T �85 25 275 50 (38%) 409Lateral habenula 1 Ca �65 �35 �150 �62T �81 58W 507 177

HypothalamusHypothalamus 10 Ca �65 �30 �93 40 940 600 4Hypothalamus 10 Ca �60 �30 �150 �46C �79 15 215 389Hypothalamus 5 Ca �50 �20 �400 20 50 (50%) 293Supraoptic 2 Ca �64 �60 �250 �65 200 110Supraoptic 2 Ca �60 �30 �400 42 50 (84%) 119Paraventricular 10 Ca �59 �33 �402 �67 12 100 (67%) 253

Midbrain and ponsArea postrema 5 Ca �65 �10 �36 �72 10 (100%) 157Cochlear nucleus 2 Ca �75 �40 �400 �59M �89 500 (50%) 101

CerebellumPurkinje 10 Ca �60 �20 �400 �40C �79 23 3,300 110 189

Dorsal rootganglionDRG 2.5 Ca �65 �35 �2,000 �82 100 (95%) 436DRG 10 Ba �60 �25 �41C �60 25 330 295

Nodose ganglion 5 Ca �60 �20 �400 �48 30 400 54Nodose 10 Ca �55 �15 �400 �29C �44 30 226

Pelvic ganglion 10 Ca �50 �25 �500 �38T �57 10 231Olfactory 2.6 Ca �57 �36 �50 �82 429

Olfactory 3 Ca �70 �40 �60 �45C �66 100 (71%) 200, 201Retina

Rod bipolar 10 Ca �60 �30 �35 �36T �59 30 1,000 (80%) 316Cone bipolar 10 Ca �60 �20 �66 �35T �60 55 1,000 (85%) 316

Oligodendrocytes 20 Ba �50 �25 �150 �63 10 47

The divalent cation concentration used to measure the currents is shown. Properties of the current-voltage relationship are described in termsof the voltage threshold where currents become detectable, the voltage where they peak (I–V peak), and the maximum amplitude observed (Imax).Methods used to calculate midpoints of the activation curves (V0.5) are denoted with the following superscripts: C, chord conductance; T, tailcurrents; M, I/Imax. The midpoint of the steady-state inactivation curve is reported (h�). Kinetics of inactivation (�inact) can vary with test potentials;therefore, the values shown represent the voltage-independent rate, which was typically measured at �10 mV. WCurrents decayed in a biexponentialmanner; to simplify comparisons, a weighted tau was calculated according to the following equation �W � A1�1 � A2�2, where A represents thefraction of current decaying with the respective time constant. The values shown for recovery from inactivation represent the voltage-independentrate, which was typically measured at �90 mV. Tail current kinetics continuously vary with repolarization potential, so when possible the valuesshown represent those obtained at �90 mV. In some cases recovery from inactivation was found to be biphasic, so the values represent the tau ofthe fast component, the fraction of channels that recover fast (in parentheses), and the tau of the slow component. Nickel responses represent eitherthe IC50 value obtained or the concentration tested and the percent block observed (in parentheses). LGN, lateral geniculate nucleus; MDN,magnocellular dorsal nucleus; L–M, lacunosum-moleculare of hippocampus; DRG, dorsal root ganglion; inact., inactivation; recov, recovery.

T-TYPE CALCIUM CHANNELS 123

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 8: Canal Calcio T

value is considerably more negative than the apparentthreshold of activation, indicating that channels can inac-tivate at potentials where they do not appear to open.Similar h� values have been obtained with the recombi-nant Cav3 channels (see Table 8), where this closed-stateinactivation has been studied in greater detail (127). Al-though most HVA channels inactivate at more positivepotentials, R-type and recombinant Cav2.3 channels inacti-vate over a similar voltage range as T-type channels (334).

The time course of recovery from inactivation is animportant property of T-type channels. LTS are oftentriggered after an inhibitory postsynaptic potential(IPSP), and this is due to fast recovery of T-type channelsduring the IPSP (deinactivation), followed by their open-ing as the membrane returns to its resting potential. Re-covery is often measured by varying the time between(interpulse) an inactivating pulse and a test pulse. In moststudies the interpulse voltage was �90 mV, and similarresults are obtained at more negative voltages. In con-trast, recovery is slower at potentials where channels canalso inactivate. There is considerable variability in thereported values for recovery, ranging from 100 to 3,300 ms(Table 2). Although most studies found that recoveryfollowed a monoexponential time course, others havefound evidence for a biexponential process. This discrep-ancy might be due to differences in the duration of theinactivating pulse, where long pulses allow channels toaccumulate in a second inactivated state from which theyrecover slowly. As observed in sensory neurons (53),recombinant channels recover with a monoexponentialtime course from short pulses and with a slower, biexpo-nential time course from long pulses, although this prop-erty is less pronounced for Cav3.3. Differences in recoverykinetics of T-type currents between different neurons canalso be due to which Cav3 isoform they express (209). Ofthe three isoforms, Cav3.1 channels recover the fastest(�120 ms from short pulses), and this time course issimilar to that observed for recovery of both native T-typecurrents (�300 ms; Table 2) and LTS (�150 ms; Table 1).

Early studies found that LVA channels were moresensitive to block by 50–100 �M nickel than HVA chan-nels (124, 152, 298). Subsequent studies found that manyneuronal T-type channels required �10-fold higher con-centrations for block, with many studies reporting IC50

values of �300 �M (Table 2). Studies on recombinantchannels provided an explanation for this diversity; onlyCav3.2 channels are highly nickel sensitive (IC50 �10 �M),while Cav3.1 and Cav3.2 are 20-fold less sensitive (230).Recombinant HVA channels also differ in their Ni2� sen-sitivity, with Cav2.3 being relatively sensitive (455). Al-though the mechanisms of block of recombinant HVA andLVA channels are complex and subject to multiple inter-pretations, a consistent finding is that block is greatest atpotentials near threshold. This voltage dependence wouldexaggerate block of LVA channels, making nickel appear

more selective than it is. In summary, block by 10–50 �MNi2� can be used to implicate Cav3.2 channels, but addi-tional evidence, such as a slowly deactivating tail current,is required to rule out Cav2.3 channels. Both these crite-ria, plus PCR, were recently applied to show the expres-sion of Cav3.2 in sympathetic ganglion neurons (231).

The temperature sensitivity of T-type currents hasbeen measured in thalamic neurons (85), DRG neurons(305), N1E-115 neuroblastoma cells (298), and GH3 cells(343). Heating from 22 to 37°C caused over twofold in-creases in current amplitudes and accelerated channelactivation and inactivation. In contrast, heating did notaffect the position of the I-V curve (305). The effects oftemperature have been found to be nonlinear, making itdifficult to calculate the effect of a 10°C increment (Q10)in temperature (298, 343). In the linear range, Q10 valueswith an average of 2.5 have been found for effects onamplitude, activation kinetics, and inactivation kinetics.

The pH sensitivity of T-type currents has been stud-ied in CA1 hippocampal neurons (405), ventrobasal tha-lamic neurons (365), and cardiac myocytes (83, 414).Acidification of the external solution decreased currentamplitudes, whereas alkalinization had the opposite ef-fect. In contrast, T-type currents are relatively insensitiveto changes in intracellular pH (405). The effect of externalpH is in part mediated by changes in the single-channelconductance, which when measured with 110 mM Ca2�

increased from 3.5 pS at pH 6 to 10.8 pS at pH 9 (414).Small changes in pH have also been shown to shift thevoltage dependence of activation and inactivation; chang-ing pH from 7.3 to 6.9 shifted both of these parameters by�3 mV, while increasing the pH to 7.7 had the oppositeeffect (365). Larger shifts in the voltage dependence ofactivation (�15 mV) have been observed using largershifts in pH (from 7.5 to 9.8) (83). The observed pKa

values were �7, indicating that T-type currents will beaffected by modest changes in extracellular pH (365, 414).

2. Slices

LVA currents have been characterized in slices pre-pared from numerous brain regions (Table 3). The moststriking difference between these data and that obtainedwith isolated neurons is that currents are much larger inslices. This difference has been attributed to loss of chan-nels that were localized on dendrites (97). In fact, T-typechannels appear to be preferentially localized to dendrites(77, 135, 195, 199, 328).

T-type currents recorded from slices also appear toactivate and inactivate at more negative potentials thanobserved in isolated neurons. The average threshold inslices was �70 mV, and peak currents were observed at�45 mV (Table 3), while in isolated neurons thresholdwas �60 and the peak occurred at �30 mV (Table 2). Thisdifference has been attributed to poor voltage clamp of

124 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 9: Canal Calcio T

neurons in slices (97). Consistent with this suggestion, thevoltage dependence of single T-type channels recordedfrom dendrites is similar to that observed in isolatedneurons (256).

In summary, T-type currents are found in neuronsthroughout the brain, with particularly large currentsfound in thalamic, septal, and sensory neurons. Theiractivation near the resting membrane potential and theirfast recovery from inactivation allow them to generateLTS. Their preferential localization in dendrites suggestsT-type channels play an important role in synaptic inte-gration.

C. Heart

Cardiac T-type currents have been the focus of manystudies due to their possible role as a pacemaker current.These studies have established that T-type current densi-ties are highest in conduction and pacemaker cells, andmuch reduced or nonexistent in ventricular myocytes.T-type currents are highest near birth and gradually de-cline but may reappear in pathological conditions. Cur-rents are larger in lower species and have a wider distri-bution. For example T-type currents are prominentthroughout the guinea pig and hamster heart but virtuallyabsent in adult ventricular myocytes of rat, cat, and dog.Although Cav3.2 was cloned from a human heart cDNAlibrary, T-type currents have not been detected in humanmyocytes (39, 235, 313).

Cardiac T-type currents were initially described indog atrium (32) and guinea pig ventricle (283, 303). Thesestudies showed that cardiac T-type currents resembledthose recorded from neurons, but differed from L-typechannels in terms of their kinetics (transient), pharmacol-ogy (dihydropyridine insensitive), conductance of diva-lent cations (Ca2� � Ba2�), lack of regulation by �-adren-ergic agonists, and smaller single-channel conductance ofBa2�. Subsequent work has established their presence in

myocytes isolated from dog Purkinje (367), rabbit sino-atrial node (152), cat atrium and latent pacemaker cells(462), hamster ventricle (362), and rat atrium (446). Ingeneral, the currents were small, displaying a peak cur-rent density below �3 pA/pF. In contrast, prominent T-type currents (��10 pA/pF) have been recorded frommyocytes isolated from shark (271), chick (202), finch(48), and mollusks (452).

Calcium channels were originally classified by theirinactivation kinetics: rapidly inactivating or transientchannels were called T type, while slowly inactivating, orlong-lasting channels were called L type (303, 306). Itshould be noted that this only holds for currents carriedby Ba2�, as cardiac L-type currents carried by Ca2� caninactivate at a similar rate as T-type currents. This is dueto calcium-induced inactivation of the L-type channel,which appears to be mediated by calmodulin tethered tothe carboxy terminus of the channel (111). This inhibitionis triggered with every heartbeat as Ca2� entry triggers alarger release of Ca2� from internal stores (37). In con-trast, Ca2� currents through T-type channels inactivate abit slower than Ba2� currents (116, 410), which excludesa similar regulation by calmodulin. Another major differ-ence between these channels is that L-type channels con-duct Ba2� threefold better than Ca2�, whereas T-typechannels conduct both equally well (discussed further insect. IIJ).

A proposed physiological role for cardiac T-typechannels is as a pacemaker current during diastolic de-polarization. These studies have been hampered by thelack of a selective blocker, relying heavily on the ability of40 �M nickel to block T- but not L-type currents (152,298). Current-clamp studies of spontaneous action poten-tials showed that nickel slowed the late phase of depolar-ization, and hence slowed the firing of rabbit sinoatrialnodal cells (SAN). Similar results have been obtained by anumber of groups using rabbit SAN (98, 353) or cat latentpacemaker cells (462). Tetramethrin (0.1 �M) was re-

TABLE 3. Properties of T-type currents in brain slices

Brain RegionDivalent,

mMThreshold,

mV

I–V

Peak,mV

Imax,pA

V0.5,mV

h�,mV

�inact,ms

�recov,ms Nickel, �M

ReferenceNos.

Cerebellum 0.5 �60 �33 �690 �51M �86 14 100 (19%) 292Thalamus

LGN 1 �65 �53 �1,500 �61M �77 26 100 200 (53%) 93LGN 2 �70 �45 �1,812 �61M �84 17 151Laterodorsal, IT,f 2 Ca �80 �55 �670 �86 32 398Laterodorsal, IT,s 2 Ca �80 �65 �630 �98 55, 69 25 (50%) 398

Hypothalamus MDN 2 �83 �50 �1,051 �70M �82 12 100 500 (100%) 302Brain stem respiratory 1.5 �65 �40 �550 �59M �81 200 (50%) 107Dorsal horn 2 Ba �54 �20 �1,000 �45M �86 14 200 (91%) 347

The voltage dependence of activation was estimated by normalizing the current observed at each test potential to the maximum ( M ) (I/Imax),fit with a Boltzmann function, then reported as the midpoint voltage (V0.5). A single concentration of nickel was tested, and the percent block isshown in parentheses. LGN, lateral geniculate nucleus; MDN, magnocellular dorsal nucleus.

T-TYPE CALCIUM CHANNELS 125

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 10: Canal Calcio T

ported to produce selective block as well, and to producesimilar effects on pacemaker cycle (152); however, sub-sequent studies failed to confirm this selectivity (165).

Similarly, a number of studies have found lower sen-sitivity and selectivity for nickel (Table 4). Two possibleexplanations for these disparate results are 1) that heartexpresses two isoforms of the T-type channel, Cav3.1 andCav3.2, and these isoforms differ in their sensitivity tonickel, and 2) that isoform expression is age and speciesdependent. Cloned Cav3.2 channels are blocked at 20-foldlower concentrations than Cav3.1 (or Cav3.3), displayingan IC50 of �10 �M (230). There is a strong correlationbetween nickel sensitivity and Cav3.2 expression, suggest-ing that native Cav3.2 channels are as sensitive as thecloned channel (323). This suggests that nickel (�100�M) can be used to implicate Cav3.2 channel expression.Expression in atrium appears to be species specific:Cav3.1 appears to predominate in rat and mice (91, 236),while in guinea pigs it is Cav3.2 (323). In fact, carefulmeasurement of the nickel dose response in guinea pigatrium revealed a biphasic response, with 80% of thechannels being blocked with an apparent IC50 of 23 �M,while 20% of the channels were inhibited with an IC50 of1,350 �M. Rabbit and cat SAN cells appear to predomi-

nantly express Cav3.2 channels. In contrast, in situ hybrid-ization of mouse heart suggests that Cav3.1 channels pre-dominate, although both isoforms were readily detectedand both were enriched in SAN relative to atrium (49). Itshould be noted that distinct in situ probes might differ intheir ability to detect their cognate sequence, makingsuch comparisons difficult. PCR amplification of mouseheart detected the expression of a splice variant of Cav3.1that differs in the III-IV loop (91).

Expression of T-type currents in developing hearthas also been studied. T-type currents are readily detect-able in neonatal hearts, increase slightly to a peak be-tween postnatal days 4 and 8, then decline slowly to asteady-state (236, 246). In other studies T-type currentswere only detected in neonatal rats, disappearing by post-natal day 21 (236). In contrast to L type, the biophysicalproperties of T-type currents are unchanged during neo-natal development. Changes in nickel sensitivity have notbeen investigated but may be informative. No differencein T-type current density was detected in SAN myocytesfrom newborn (3–10 days old) and adult (41–48 days old)rabbits (329). Expression of Cav3.2 mRNA is higher infetal human heart than adult, whereas Cav1.2 shows theopposite pattern (331).

TABLE 4. Properties of T-type currents in peripheral tissues

TissueDivalent,

mMThreshold,

mVI–V Peak,

mV pA/pFImax,pA

V0.5,mV

h�,mV

�inact,ms

�deact,ms �recov, ms Nickel, �M

ReferenceNos.

HeartAtrium 5 Ca �50 �30 �0.3 �24 20 32Atrium 5.4 Ca �50 �30 �1.0 �75 �40M �57 15 197 50 (66%) 444Atrium 2 Ca �50 �30 �2.6 �42C �68 13 1.3 160 236Atrium 1.8 Ca �50 �30 �350 �24T �59 23 323Atrium 2.5 Ca �60 �40 �115 �38C �59 10 116SAN 2.5 Ca �47 �10 �2.1 �88 �23C �75 3.2 144 40 (100%) 152Latent pacemaker 2.7 Ca �50 �20 �3.3 �90 �31C �57 40 (54%) 462Purkinje 2 Ca �60 �30 �1.7 �391 �48C �68 10 100 100 (90%) 165Purkinje 5 Ca �40 �25 �2.9 �700 �47M �37 4 83 (67%) 50 (47%) 410

444Ventricle 5.4 Ca �50 �10 �100 11 100 (100%) 129Ventricle 20 Ba �50 �30 �5.8 �858 �43C �77 13 362

Skeletal muscle 1.8 Ca �55 �42C �63 13 4510 Ca �45 �15 �3.6 �400 �24C �41 16 224 (65%) 5.4 38

5,160Spermatocytes 10 Ca �60 �20 �6.5 �218 �47C �64 7 1.6 118 200 (75%) 350Adrenal

Glomerulosa 20 Ba �45 �30 �7T �56 7 82Fasciculata 10 Ca �50 �15 �8.2 �272 �17G �50 18 1.7 400 (50%) 20 287

4,800Pituitary

Pars intermedia 10 Ca �50 �15 �150 5T 1.8 84Lactotropes 5 Ca �48 �12 �90 �28M �66 15 100 (80%) 240Melanotropes 11 Ba �50 �20 �30 �61 19 204Gonadotropes 20 Ca �40 �20 �75 �27C �61 15 3 262

Pancreas�-Cell 10 Ca �40 �16T 21 2.5 320�-Cell 15 Ba �60 �30 �35 �14M 100 (100%) 25

Definitions are as in Table 2. Methods used to calculate activation curves are denoted with the following superscripts: C, chord conductance;T, tail currents; G, constant-field equation of Goldman-Hodgkin-Katz; M, I/Imax. In some cases recovery from inactivation was found to be biphasic,so the values represent the tau of the fast component, the fraction of channels that recover fast (in parenthesis), and the tau of the slow component.SAN, sinoatrial node.

126 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 11: Canal Calcio T

Cardiac hypertrophy appears to trigger a return tothe neonatal pattern of gene expression (388), leading toreexpression of T-type channels in rat and cat ventricularmyocytes (173, 268, 308). RNase protection assays indi-cate that Cav3.1 transcripts are upregulated in a rat modelwhere hypertrophy is induced by infarction (173). Al-though levels of Cav3.2 transcripts were not examined,the sensitivity of reexpressed currents to nickel blocksuggests a contribution of Cav3.2 channels as well (173,268). T-type currents are also reexpressed in dedifferen-tiated rat ventricular myocytes (114). T-type currentshave been found to be increased twofold in cardiomyo-pathic Syrian hamsters relative to other hamster strains(46, 362). Studies using 8-mo-old animals found that thevoltage dependence of activation and inactivation wasshifted to more negative potentials in ventricular myo-cytes from cardiomyopathic animals (362). In contrast,studies using 1-day-old hamsters found that only inactiva-tion was shifted (46). The relevance of these observationsto human pathologies is unclear as T-type currents havenot been detected in myocytes isolated from normalatrium (235) or diseased ventricle (39, 337).

Calcium influx through voltage-gated channels acti-vates intracellular Ca2� release channels, leading to largerelevations in intracellular Ca2� and hence muscle con-traction. The role of L-type channels in this process is wellestablished. T-type channels are also capable of triggeringthis CICR, albeit much more weakly in both cardiac (370,461) and skeletal muscle (133). T-type current-inducedcontraction developed more slowly, suggesting that thesechannels are not localized near SR stores as are L-typechannels. This observation coupled with the lower ex-pression of T-type channels in working myocytes indi-cates that they do not play a major role in excitation-contraction coupling. Consistent with this notion,mibefradil has little effect on cardiac inotropy at doses

that lower blood pressure (356). In contrast, T-type chan-nels may be localized near SR in atrial pacemaker cells ofthe cat (180). These authors suggested a novel pacemak-ing mechanism where Ca2� influx through the T-typechannel triggers subsarcolemmal Ca2� sparks, which inturn activate Na�/Ca2� exchange currents that depolarizethe membrane to threshold.

A role for T-type channels in triggering atrial natri-uretic peptide (ANP) secretion has been inferred from theeffects of mibefradil (236, 434). Evidence that mibefradilacted on T-type channels was provided by the observationthat L-type blockers were much less potent blockers, orstimulated ANP secretion.

D. Kidney

Relatively little is known about the physiologicalroles of T-type channels in kidney. T-type currents haveonly been recorded from smooth muscle cells (SMC)isolated from rat interlobular and arcuate arteries (143).Heterogeneity in Ca2� currents was observed, with ap-proximately one-third of the freshly isolated myocytesdisplaying only T-type currents (“T-rich”), one-third only Ltype, and the rest a mixture. Currents in T-rich cells weresome of the largest ever recorded for cardiovascularsmooth muscle (�156 pA in 2.5 mM Ca2�; Table 5), whichallowed their recording in physiological Ca2� solutions.Consistent with their assignment as T-type channels, cur-rents activated and inactivated at voltages 20 mV morenegative than observed for L-type currents. In contrast tomost T-type currents, these SMC currents activated (I-Vpeak �10 mV) and inactivated (h� � �50 mV) at voltagesthat were slightly more positive, which might be charac-terized as mid-voltage activated. They also inactivatedabout twofold slower during a pulse (�inact � 46 ms at �10

TABLE 5. Properties of LVA currents in smooth muscle

TissueDivalent,

mMThreshold,

mVI–V Peak,

mV Imax, pAh�,mV

�inact,ms Nickel, �M

ReferenceNos.

ArteriesCoronary 10 Ba �60 �10 �125 �65 14 130, 131Coronary 20 Ba �40 �30 �250 �55 332Aortic 20 Ca �60 �30 �140 �80 20 600 3Aortic 1.5 Ca �60 �30 ��100 �65 339Aortic 20 Ba �40 �12 24 339Mesenteric 10 Ba �40 �30 �63 372Rabbit ear 110 Ba �30 �10 �20 �55 34Rat tail 20 Ba �50 �38 46 428

VeinsPortal 5 Ca �45 �10 �263 �50 13 246

OrgansLung 10 Ca �50 �10 �200 448Uterus 1.8 Ca �60 �30 �300 �70 453Bladder 1.8 Ca �56 �10 �200 �60 100 (100%) 382Colon 2 Ca �60 �10 �385 �58 20 30 (50%) 445

Definitions are as in Table 2.

T-TYPE CALCIUM CHANNELS 127

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 12: Canal Calcio T

mV). Atypical currents with similar properties have beenreported for SMC isolated from rat portal vein (315) andfrom the terminal branches of guinea pig mesenteric ar-teries (291).

Quite surprisingly, the human tissue that expressesthe most mRNA for Cav3.2 is the kidney (90, 438). Incontrast, both Cav3.1 and Cav3.3 are predominantly ex-pressed in brain (228, 288, 289, 324). Skøtt and co-workers(11, 156) have studied the distribution Cav3.1 and 3.2 inkidney and concluded that Cav3.1 was in the tubules,while Cav3.2 was in renal smooth muscle (11, 156). RT-PCR indicated that mRNAs for Cav3.2, and to a lesserextent Cav3.1, were expressed in rat (but not rabbit)glomerular afferent and efferent vessels, and vasa recta.The same study reported that K�-induced contraction ofperfused rabbit afferent arterioles could be blocked bylow concentrations of mibefradil (IC50 �10 nM) thatshould be selective for T-type channels. Nickel could alsoblock this response; however, the concentrations re-quired (1 mM) would be expected to block almost all Ca2�

channels. Similar results were obtained with rat isolatedperfused hydronephrotic kidneys: either 1 �M mibefradilor 100 �M nickel (a relatively selective dose) dilatedafferent and efferent arterioles constricted with ANG II(314). Selectivity of mibefradil for T-type channels wasdemonstrated by coadministration with the L-typeblocker nifedipine. In these experiments 1 �M nifedipinecaused a modest dilation of the efferent arteriole but didnot prevent the more pronounced dilation induced bymibefradil.

Cav3.1 was detected in dot blots of human fetal, butnot adult, kidney (288). RNase protection assays of ratkidney regions indicated that Cav3.1 was predominantlyexpressed in the inner medulla (11). RT-PCR of microdis-sected nephron segments indicated that Cav3.1 was ex-pressed in distal convoluted tubules, connecting tubules,and collecting ducts. This study also examined the distri-bution of Cav3.1 protein using a polyclonal antibodyraised against the first 22 amino acids of the deducedCav3.1 sequence (11). Immunoreactivity was detected inthe apical domains of the distal convoluted tubules and inprincipal cells of connecting tubules and inner medullarycollecting ducts. Preincubation of the antibody with pep-tide blocked the signal; however, Western blots were notpresented to demonstrate the selectivity of the antibody.These results suggested that T-type channels, along withendothelial calcium channels (ECaC), might be involvedin Ca2� reabsorption.

Calcium channel blockers are widely used in thetreatment of hypertension. It is generally accepted thatthey work by blocking L-type channels in vascular smoothmuscle, leading to vasodilation. Recent studies suggestthat part of their effect may be mediated by changes inrenal hemodynamics (172). In intact dogs, both nifedipineand mibefradil increased renal blood flow, whereas only

nifedipine affected glomerular filtration rate. Mibefradilhas also been reported to decrease renin secretion in ratswith renal artery clips, but had no effect on renin secre-tion from isolated juxtaglomerular cells (423). In conclu-sion, these studies suggest that T-type channels on affer-ent and efferent glomerular arterioles may be animportant therapeutic target (89). A possible advantage toa T-selective antihypertensive drug is that it may alsoprevent glomerular damage (192).

E. Smooth Muscle

In addition to L-type currents, SMCs isolated fromarteries, veins, and organs have also been reported tohave LVA currents. With the exception of the kidneyresults noted above, cardiovascular SMCs have tiny LVAcurrents. Therefore, most studies have had to use veryhigh concentrations of charge carrier, which shifts gatingto positive potentials, and currents have only been char-acterized in terms of their I-V relations and sensitivity toholding potential. These characteristics are not sufficientto establish an LVA current as T type, since HVA currentsactually represent a spectrum of mid- to high-voltage-activated currents. HVA currents also inactivate over awide range of potentials, with some R-type currents inac-tivating over the same voltage as T-type channels. Theseconcerns take on additional weight with the finding thatsome vascular SMCs express non-L-type HVA currents(291) such as P-type currents (155). Native P-type cur-rents can be mid-voltage activated (33). In addition, anickel- and mibefradil-sensitive LVA current has beendescribed in colonic myocytes that gates similar to T-typechannels but has different permeability properties (214).LVA channels can be classified as T type by their slow tailcurrents, which surprisingly have not been reported, andby their single-channel properties, which have been re-ported (34, 130, 450). Because T-type channels inactivateat the resting membrane potentials of most SMCs (�75 to�60 mV, reviewed in Ref. 168), it has been hard to discerntheir physiological role. Perhaps they contribute to basalintracellular Ca2� concentrations via their window cur-rent, or after transient hyperpolarizations induced byspontaneous transient outward currents (STOCs).

LVA currents have been found in the following arte-rial SMCs: coronary (130, 281, 332), aortic (3), mesenteric(150, 311, 372), rabbit ear (34), rat tail (327, 428), andmiddle cerebral arterioles (167). RT-PCR detected theexpression of both Cav3.1 and Cav3.2 in rat mesentericarterioles (150). Single-channel studies have establishedthe presence of 7.5-pS T-type channels in freshly isolatedguinea pig coronary SMCs (130). This study also reportedthat whole cell LVA currents could only be observed inhalf of the cells. In contrast, LVA currents were neverdetected in freshly isolated coronary SMCs from rabbits

128 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 13: Canal Calcio T

(269) or humans (332). LVA currents were found if thecells were cultured, and these currents appear to be Ttype based on their voltage dependence and 1:1 conduc-tance of Ca2� and Ba2� (332). Similar results have beenobtained with aortic SMCs; T-type currents are expressedin proliferating cultures, but neither in confluent cultures(3, 339) nor in freshly isolated cells (220). In fact, T-typecurrents were only detected in cells that were in G0 or G1

phases (220), leading to the hypothesis that they mightplay a role in proliferating SMCs (338). Consistent withthis notion is the finding that mibefradil could reduceneointima formation following balloon injury to rat ca-rotid arteries (355). However, it should be noted thatT-type currents were not observed in SMCs prepared frominjured rat aorta, although a transient downregulation ofL-type channels was documented (333).

LVA currents have also been detected in SMCs iso-lated from veins, such as saphenous (450), portal (246),and azygous (282). T-type single-channel currents wererecorded in saphenous vein SMCs (450). Isradipine (PN200–110) was found to block T-type currents of rat portalvein with an IC50 of 0.5 �M, while L-type currents wereblocked at 1 nM (246). The sensitivity of these T-typecurrents was greater than observed in other preparations(see sect. VA), suggesting that Cav3 isoforms may differ intheir pharmacology. Current-clamp recordings revealedthat portal vein SMCs could fire action potentials thatresembled those observed with cardiac myocytes (246).Isradipine (10 nM) inhibited the plateau phase with noeffect on the rising phase, suggesting that T-type channelsmight play a pacemaker role in these cells.

LVA currents have been described in SMCs frombronchi (185, 448), ileum (373), colon (445), bladder(382), and uterus (453). The results with pregnant humanuterine SMCs are notable because the T-type currents (�3pA/pF in 1.8 mM Ca2�) were larger than the L-type cur-rents (453). These currents activated and inactivated(h� � �70 mV) at potentials typical of most T-type cur-rents (Table 5). Current-clamp recordings indicated that adepolarizing pulse could trigger an action potential; how-ever, the threshold (�20 mV) was higher than observedwith neuronal LTS. Sizable (�200 pA) T-type currentswere also described for porcine bronchial smooth muscle,being detected in 30% of the cells, but absent from tra-cheal SMCs (448). LVA currents (�1.4 pA/pF) were foundto disappear when bladder SMCs were cultured (382).Concomitant with this loss was a shift in the threshold foraction potential generation from �25 to �15 mV, sugges-tive of a role for T-type channels.

F. Skeletal Muscle

Newborn rodent skeletal muscle was also used inearly studies to establish the existence of T-type channels

as separate from L-type channels (30, 81). Developmentalstudies of mice and rats have shown that the T-typecurrent is only found in embryonic and newborn muscleand disappears by 3 wk of age (29, 38, 142). Concomi-tantly, the slow L-type current increases. Studies on themuscular dysgenesis (mdg) mouse clearly establishedthat the T- and L-type channels were encoded on separategenes (30, 67, 212), and it is the L-type channel (Cav1.1)that plays a critical role in excitation-contraction coupling(395). Recent studies have shown that the T-type currentis important for myoblast fusion (45). In particular, it wasshown that T-type channels could generate sufficient win-dow currents to alter intracellular Ca2� concentrations,and trigger fusion. This hypothesis was further supportedby pharmacological experiments using low concentra-tions of Ni2� and amiloride, and by antisense oligonucle-otides directed against Cav3.2 (45). The selective expres-sion of Cav3.2 in skeletal muscle fibers has also beendemonstrated using single-cell PCR (38). The electrophys-iological properties of these currents (Table 4) and Ni2�

sensitivity closely resemble recombinant Cav3.2 currents,although differences have been noted in kinetics of acti-vation, inactivation, and recovery from inactivation (38).Early studies also noted kinetic differences as a functionof development (142), suggesting that these differencesare not due to rapid events such as phosphorylation, butrather slower events such as gene transcription of auxil-iary subunits (see sect. IVE).

G. Sperm

Calcium channels appear to play a role in mediatingthe egg-induced acrosome reaction that precedes fusion(for review, see Ref. 94). T-type channels have been im-plicated in this process, although other Ca2� conduc-tances may be involved (134, 435). One reason for thisuncertainty is that it is virtually impossible to patch clampmature sperm (335). Therefore, electrophysiological stud-ies have used spermatogenic cells, which are primaryspermatocytes in the pachytene stage. The only voltage-dependent Ca2� channels in these cells are T type, andtheir biophysical (350) and pharmacological properties(15) have been well characterized. PCR results indicatethat Cav3.2 is the predominant isoform expressed in hu-man testis (182, 375). Consistent with this result is thehigh sensitivity (IC50 � 34 �M) with which nickel blocksthe spermatocytic T-type current (15), although it shouldbe noted that Sertoli cells also express nickel-sensitiveT-type currents (225). Similar concentrations of nickelalso block the acrosome reaction (121, 375). Similarly, thefollowing compounds have been found to block sper-matocyte T-type channels and the acrosome reaction atsimilar concentrations: isradipine (PN 200–110), nifedi-pine, pimozide, amiloride, mibefradil, W-7, and trifluoper-

T-TYPE CALCIUM CHANNELS 129

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 14: Canal Calcio T

azine (13, 15, 121, 247, 350, 375). Spermatocytic T-typechannels appear to be regulated by tyrosine and calmod-ulin-dependent protein kinases (14, 247).

H. Endocrine Tissues

1. Adrenal

T-type channels have been implicated in the secre-tion of aldosterone (76) and cortisol (109). Calcium cur-rents, and their regulation by secretagogues, have beenextensively studied in cells isolated from bovine adrenalzona glomerulosa and fasciculata. Glomerulosa cells ex-press a mixture of T- and L-type channels (82), whereasfasciculata cells predominantly express T-type channels(22, 287). Although these cells can be considered nonex-citable due to their lack of Na� channels, they are stillcapable of generating Ca2�-dependent action potentials(22). In contrast, adrenal chromaffin cells express HVACa2� and Na� currents but no LVA currents (16).

Serum K� is regulated by secretion of aldosteronefrom adrenal glomerulosa cells. Glomerulosa cells areexcellent K� sensors and depolarize after small increasesin serum K�, which in turn leads to the activation ofT-type channels, Ca2� influx, and stimulation of aldoste-rone synthesis and release. For example, increasing K�

from 2 to 5 mM causes the resting membrane potential ofisolated cells to depolarize from �97 to �78 mV (76).Aldosterone secretion is also stimulated by ANG II, aneffect mediated by stimulation of T-type channel activity(discussed in sect. IIIB).

Cortisol secretion from adrenal zona fasciculata cellsis regulated by ACTH, which depolarizes these cells byinhibiting a K� conductance (286). Depolarization wouldactivate T-type channels, leading to a rise in intracellularCa2�, and ultimately cortisol synthesis and secretion. Cor-tisol secretion can be inhibited by a variety of blockersincluding mibefradil, and this block occurs over the sameconcentration range as their block of the T-type channels(109, 141, 345). ACTH may also regulate the expression ofT-type channels (22).

Adrenal T-type channels are encoded by Cav3.2. Insitu hybridization of rat and bovine glands detected highlevels of Cav3.2 expression in the cortex, with only traceamounts of Cav3.1 and Cav3.3 (358). The biophysical prop-erties and sensitivity to Ni2� block of both glomerulosaand fasciculata cells are also consistent with their beingencoded by Cav3.2 (287, 358).

2. Pituitary

LVA Ca2� currents have been recorded from cells of theanterior and intermediate lobes of the pituitary gland includ-ing (84, 392, 439) lactotropes (240), corticotropes (261),gonadotropes (379), and thyrotropes (204). These cells gen-

erate spontaneous action potentials and Ca2� spikes, lead-ing to secretion of hormone (see Ref. 406 and referencestherein). Hormonal regulation of these LVA currents sup-ports the idea that these currents are involved in secretion(see sect. III). The voltage- and time-dependent properties ofthese currents are similar, but not identical to other nativeT-type currents (Tables 2 and 4). Notably the peak of the I-Vcurve is slightly depolarized. In addition, one study reportedthat the LVA current was fivefold larger in Ba2� than Ca2�

(379), a property typically ascribed to HVA currents. Assuggested previously, not all LVA currents are carried byT-type channels, and other criteria must be applied (19, 33,376). One such discriminating property is their tail deactiva-tion kinetics, and slowly deactivating T-type currents areclearly present in pituitary cells (84). Similarly, these chan-nels can be distinguished at the single-channel level by theirconductance for Ba2�, and again, pituitary T-type currentshave been clearly identified (203). In situ hybridization stud-ies suggest all three Cav3 isoforms are expressed, althoughCav3.2 was the predominant isoform detected (393). Consis-tent with this result, T-type currents were reported to benickel sensitive, with 80% of the current blocked by 100 �MNiCl2 (240). Currents recorded from GH3 cells are not nickelsensitive (IC50 � 777 �M), suggesting that this tumor-de-rived cell line most likely expresses Cav3.1 channels (160). Inconclusion, T-type channels appear to play a pacemaker rolein anterior pituitary, although they may also be involved instimulus-secretion coupling (258).

3. Pancreas

T-type channels appear to play a similar pacemaker rolein insulin secretion from pancreatic �-cells of the islets ofLangerhans. Increases in plasma glucose and its metabolismin �-cells leads to increases in cellular ATP and inhibition ofKATP channels (300). Closure of KATP channels initiates thepacemaker cycle by depolarizing the �-cell membrane toabout �55 mV. Riding on top of this slow plateau depolar-ization are rapid calcium-dependent spikes (reviewed in Ref.352). Insulin secretion can be blocked by a wide variety ofagents, indicating that these spikes activate L-, P-, and R-typecurrents (238, 258, 418).

T-type currents in pancreatic �-cells have been char-acterized at the whole cell (Table 4; Refs. 25, 320) andsingle-channel level (17, 348). Expression is species de-pendent, with little or no expression in rodents, butreadily detectable in humans (25, 426). Therefore, it isnotable that LVA currents were found in �-cells fromdiabetes-prone NOD mice (426). The expression of LVAcurrents in mouse cells has been reported to be stimu-lated by a 6-h treatment with cytokines (25 U/ml of inter-leukin-1� plus 300 U/ml of interferon-�; Ref. 427). How-ever, the I-V curve for the cytokine-induced currentpeaked between �10 and �20, while under similar re-cording conditions this group found that bona fide, slowly

130 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 15: Canal Calcio T

deactivating, T-type currents peaked between �20 and�10 mV (426). Human �-cells express both LVA and HVAcurrents, and in 20% of the cells tested, the LVA currentwas dominant (25). Voltage-clamp recordings indicatedthat 100 �M Ni2� could block the LVA current, whilecurrent-clamp recordings indicated that it slowed actionpotential frequency. Similar studies using the rat insuli-noma cell line INS-1 found that low concentrations ofnickel could completely block T-type currents (IC50 � 30�M) and partially block insulin secretion (42). Nickel-sensitive T-type currents (IC50 � 3 �M) have also beenrecorded from the mouse insulinoma cell line NIT-1 (426).Li and co-workers (463) also cloned a splice variant ofCav3.1 from an INS-1 cDNA library. The high sensitivity tonickel block suggests that Cav3.2 channels are also ex-pressed, and this conclusion is supported by Northernanalysis (438).

I. Cell Lines

T-type currents have been reported in a number ofcell lines such as 3T3 fibroblasts (73) and many lines oftumor origin, such as neoplastic B lymphocytes (128), therelated mouse neuroblastoma-derived lines N1E-115 (298,368), NG108–15 (196, 334), 140–3 (144), N18 (397), andND7–23 (213); human neuroblastoma lines IMR-32 (65)and SK-N-MC (18); rat pituitary lines GH3 (160, 270);human TT cells (also called h-MTC) and rat 6–23 (clone 6)cells from medullary thyroid tumors (43, 44); human Y79retinoblastoma cells (24); AT-1 from mouse atrium (351);rat smooth muscle lines A7r5 and A10 (272); and thepancreatic insulinoma lines INS-1 (rat; Ref. 42), HIT-T15(hamster; Ref. 259), and NIT-1 (mouse; Ref. 426). Many ofthese cell lines were reported to express almost exclu-

sively T-type currents, allowing characterization of thesechannels in the absence of contaminating HVA currents(Table 6). The study of T-type currents in cell lines hasadvanced our understanding of gating, permeation, andpharmacology (73, 160, 368). For example, Chen and Hess(73) developed models of gating using recordings from3T3 fibroblasts. They also showed that the T-type currentsof NG108–15 cells had different kinetics, leading them topredict the existence of multiple channel isoforms. Asimilar conclusion was reached from studies on TT cells,which also recovered from inactivation much moreslowly than observed previously (44). Differences in thepharmacology of T-type channels also suggested the ex-istence of distinct channel isoforms (160). The cloning ofmultiple Cav3 isoforms that differ in their kinetics, phar-macology, and recovery from inactivation supports thishypothesis (209, 228, 404).

J. Single-Channel Recordings

Single-channel studies provided conclusive evidencethat T-type channels were distinct from HVA Ca2� chan-nels (63, 303, 306). T-type channels were distinguished bytheir smaller currents, insensitivity to dihydropyridines,and voltage dependence of activation and inactivation. Inaddition to providing criteria for resolving T-type cur-rents, single-channel studies have provided insights intoboth hormonal regulation of activity and gating.

With isotonic Ba2� as the charge carrier (110 mM),the single-channel current at a test potential of 0 mV is�0.3 pA for T-type channels (Table 7) and �1.5 pA forL-type channels. Single-channel conductance is defined asthe slope of the line relating single-channel amplitudesversus test potential and is commonly reported in pico-

TABLE 6. Properties of T-type currents in cell lines

Cell LineDivalent,

mMThreshold,

mVI–V Peak,

mVImax,pA

V0.5,mV

h�,mV

�inact,ms

�deact,ms �recov, ms Nickel, �M

ReferenceNos.

HumanTT 10 Ca �40 �15 �211 �27T �51 16 2 1,900 5 (57%) 44IMR-32 10 Ca �50 �20 �80 15 200 (52%) 65Y79 20 Ba �50 �20 �87 �32C �40 20 24

RatC-cell 6-23 10 Ca �50 �15 �500 �31M �57 20 2 1,260 5 (22%) 43GH3 10 Ca �50 �20 �600 �33G �71 20 5 200 (80%) 777 160

1,000Mouse

MAb-7B 2.5 Ca �65 �35 �80 �82 16 128N1E-115 50 Ba �50 �20 �300 47 298NG108-15 10 Ba �60 �15 �140 �27C �53 20 2 334ND7-23 10 Ba �50 �30 �119 �67 18, 32 845 100 (60%) 2133T3 20 Ba �50 �20 �500 �65 10 2 100 73AT-1 2.5 �60 �35 �500 �48C �64 14 2.5 150 (73%) 351

1,200

Definitions are as in Table 2. Methods used to calculate activation curves are denoted with the following superscripts: C, chord conductance;T, tail currents; G, constant-field equation of Goldman-Hodgkin-Katz; M, I/Imax.

T-TYPE CALCIUM CHANNELS 131

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 16: Canal Calcio T

seimens (pS). When recorded in Ba2�, the conductance ofT-type channels is 7–8 pS (Table 7), which is smaller thanthat observed for either N-type (13–15 pS) or brain L-typechannels (25–28 pS) (118, 125, 417). Somewhat surpris-ingly, all three families have a similar Ca2� conductance.For example, the L-type channel conductance dropsthreefold in Ca2� (149, 367). Recombinant Cav2.1 andCav2.2 channels also preferentially conduct Ba2� (�2-fold), whereas Cav2.3 conducts both ions equally well(56). In contrast, native T-type channels conduct Ca2� andBa2� (and Sr2�) equally well (64, 104, 367, 368). Theconductance plots differ in their relative position, withT-type channels gating at more negative ranges, and ex-trapolating to a reversal potential of approximately �40mV, while HVA currents appear to reverse at �60 mV (34,52, 118, 124, 194, 203, 213, 367, 417).

Estimates of permeability based on reversal poten-tials yield the opposite rank order, with Ca2� being morepermeable than Ba2� (128). Similar results have beenobtained using recombinant T-type (363) and native L-type channels (161) and suggest that Ca2� binds with highaffinity in the pore and permeates when displaced by asecond Ca2�. A very interesting property of all voltage-gated Ca2� channels is that they become highly perme-able to Na� in the absence of Ca2�. Apparently Na� arenot capable of displacing the bound Ca2�, and hence donot permeate. This block is unidirectional, since monova-lent cations can displace Ca2� bound in the pore when

they approach from the intracellular side of the channel(255, 363). The selectivity sequence for monovalentcations through T-type channels is Li�� Na� � K� �Rb� � Cs� (128, 255). In fact, T-type channels begin topass outward K� currents at physiological concentrationsof Ca2� and voltage (�28 mV; Ref. 128). HVA channelscan also pass outward currents, but this requires depolar-ization of the membrane beyond �70 mV (161). Our un-derstanding of Ca2� channel permeation is far from com-plete, and subject to multiple interpretations (275, 442).

Single-channel currents have also been studied atCa2� concentrations closer to the physiological levels,displaying a single-channel conductance of �2 pS in 5 mM(64) and 4.6 pS in 10 mM Ca2� (21). Whole cell studiesusing varying concentrations of Ca2� indicate that cur-rents increase sharply over the 1–10 mM range, thensaturate at higher concentrations. These studies have al-lowed estimates of the affinity of the open channel forCa2�. The average KD value from seven studies is 4.9 mM,with considerable variability (range 0.3�10 mM) (4, 22,54, 152, 160, 389, 429).

The transitions between open and closed states havebeen studied in detail in DRG neurons (64), cardiac myo-cytes (104), fibroblasts (73), and N1E-115 neuroblastomacells (368). T-type channels respond to depolarizingpulses by opening in bursts, that is, channels open andclose many times, become silent for a while, and at somevoltages may burst again. The fast opening and closings

TABLE 7. Single-channel properties

Source Divalent, mMCurrent at 0

mV, pAConductance,

pS SelectivityReference

Nos.

Rat DRG 40 Ca �0.2 5.2 Ca � Ba 64Chick DRG 110 Ba �0.4 8 124Mouse DRG 60 Ba �0.25 7.2 Ba � Sr 217Hippocampal CA3 pyramidal 95 Ba �0.3 7.8 118Septum and diagonal band 100 Ba �0.2 7.8 147Cerebellar Purkinje 110 Ba �0.3 9.0 52Pituitary menalotropes 110 Ba �0.4 8.1 203Hypoglossal motoneuron 110 Ba �0.3 7 417Retinal ganglion 96 Ba �0.36 8 194Cardiac ventricle 110 Ca �0.33 6.8 Ba � Ca 104Cardiac Purkinje 110 Ba �0.4 9.0 Ba � Ca 367Ear artery SMC 110 Ba �0.3 6.9 34Coronary artery SMC 110 Ba �0.38 7.5 130Saphenous vein 90 Ba �0.35 8 450Adrenal glomerulosa 110 Ba �0.25 8.7 Ba � Ca 27Pancreatic �-cells 100 Ba �0.3 6.4 348Neuroblastoma N1E-115 60 Ba �0.25 7.2 Ca � Ba � Sr 368Neuroblastoma ND7-23 110 Ba �0.25 7.9 213Neuroblastoma 140-3 110 Ba �0.3 6.84 144Low divalent

DRG 5 Ca 2 64Cardiac ventricle 10 Ca �0.16 4.7 21Coronary artery SMC 10 Ca �0.16 4.7 21Coronary artery SMC 10 Ba �0.2 4.6 130

Single-channel properties were measured from cell-attached patches using the indicated concentration of divalent cation as charge carrier. Theslope conductance of high-voltage-activated channels is similar to T-type channels when recorded in Ca2� solutions but can be distinguished by theircurrent amplitudes at a test potential of 0 mV. DRG, dorsal root ganglion; SMC, smooth muscle cell.

132 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 17: Canal Calcio T

can be modeled as transitions between a closed state C(C4 in Fig. 3) and an open state O. The transition rates outof the open state determine the mean open time, which istypically 0.5–2 ms (Table 7). Multiple bursts are seenduring depolarizing pulses just above the threshold forchannel opening. At more depolarized potentials a singleburst is observed, indicating that channels entered aninactivated state. Analysis of the time a channel spends inclosed states reveals a bimodal distribution; closingswithin a burst are short (�1 ms), whereas the time be-tween bursts is longer (�10 ms). Channels can also openpartially, leading to a subconductance state that has asmaller current (27, 64, 104). In contrast to other ionchannels (66), the gating behavior of this subconductancestate was found to be similar to the main conductancestate (64).

A distinguishing kinetic feature of T-type channels isthat whole cell current traces recorded during an I-Vprotocol produce a criss-crossing pattern, where succes-sive traces cross each other. This is largely due to slowactivation kinetics at threshold potentials (�60 to �40mV). At the single-channel level, T-type channels show along latency to the first opening, especially at thresholdpotentials, where 40 ms may pass before the first opening.This delay presumably reflects slow transitions betweenclosed states and can lead to a sigmoidal rise in the wholecell current. This rate is voltage dependent, such thatchannels open faster at more depolarized potentials (104).

In most sweeps (�70%) the channel does not open atall, resulting in a blank sweep (73, 104). This suggests thatchannels might inactivate without opening. Such closedstate inactivation has been observed in other channelsand is particularly prominent in Cav2 channels (321).Closed state inactivation also provides an explanation for

the whole cell observation that steady-state inactivationcan be observed at potentials more negative than channelopening. This inactivation is slow and voltage dependent(364). Inactivation from open (or the proximal closedstate C4) has similar kinetics and limits the channel to oneburst per sweep. Whole cell recordings suggest that thisrate is relatively voltage independent at potentials above�20 mV. Several models of T-type channel gating havebeen proposed that can simulate the observed activity.One reason these models differ is that they are based ondisparate results. Unresolved questions include the fol-lowing: Does the voltage dependence of the mean opentime explain the slowly deactivating tail currents ob-served at the whole cell level (64)? One study found thatopen times were shorter at more negative test potentials(64), another found no difference (104), while two foundthat they increased (217, 368). To complicate mattersmore, two studies have found evidence for a second typeof opening of longer duration (27, 147). Discrepancy inmean open time measurements has been attributed to thelow signal-to-noise ratio of channel openings (73). A re-lated question is: Does the burst duration vary with testpotential? Most studies report the burst duration at asingle voltage, although one study found that it increasedfrom 5 ms at �40 mV to 14 ms at �10 mV (104). Thesemodels might also differ because of T-type channel het-erogeneity, as evidenced by the distinct gating behavior ofthe three recombinant Cav3 channels (228). Recently, amodel has been proposed that can simulate the gatingbehavior of both recombinant Cav3.1 and Cav3.3 channels,although different rate constants for the transitions arerequired (127, 364). The model assumes that depolariza-tion leads to sequential movement of each of the four S4voltage sensors (see sect. IVB), culminating in a finalclosed state that precedes the open state. The open statecan either deactivate (return to C4) or inactivate (IO).Closed state inactivation is allosterically coupled to S4movement and becomes prominent after three S4 sensorshave moved (Fig. 3). Movement of the fourth voltagesensor does not affect the inactivation rate, so open chan-nels inactivate at the same rate as closed channels in C3and C4 states. The speed of voltage sensor movement andsubsequent transitions among closed states appears to beslower in Cav3.3 than Cav3.1, leading to slower activationkinetics. The channels also differ in their equilibriumbetween C4 and O states, with Cav3.3 favoring C4 whileCav3.1 favors O. Therefore, Cav3.1 channels inactivatemore from open states, whereas Cav3.3 channels inacti-vate more from closed states. The model accounts formost, but not all, of the observed properties of T-typechannels; in particular, it does not attempt to explainmultiple open states or ultraslow (�1 s) inactivation(127). In summary, voltage-gated Ca2� channels can bedistinguished by their single-channel properties, with T-

FIG. 3. Simplified gating scheme. The model represents transitionsof T-type channels between rested (R), closed (C), inactive (I), and openstates (O). Changes in the membrane potential are thought to inducemovement of the S4 voltage sensors, which leads to rearrangements inthe channel structure. Movement of three of the four S4 regions issufficient to induce closed states that inactivate relatively quickly. Theproximal closed state (C4) can also show closed state inactivation orlead to channel opening. Single T-type channels show a propensity toopen in bursts during a prolonged depolarization, and this representstransitions between C4 and O states. The model has been simplified fromthe version presented by Frazier et al. (127) by omission of C1, C2, I1, andI2. The effects of voltage suggest that S4 sensors must move back towardtheir original position before channels can recover from inactivation.

T-TYPE CALCIUM CHANNELS 133

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 18: Canal Calcio T

type channels showing a smaller conductance in Ba2� andequal conductance of Ca2� and Ba2�.

III. REGULATION

Among the first criteria used to distinguish LVA fromHVA channels was their resistance to rundown in solu-tions that lack cAMP, ATP, and Mg2�, suggesting that LVAchannels were not phosphorylated by cAMP-dependentprotein kinase (115). Consistent with this hypothesis, hor-mones that stimulate adenylyl cyclase, such as the �-ad-renergic agonist isoproterenol, had no effect on T-typecurrents recorded from either canine atrial myocytes (32),rabbit sinoatrial nodal cells (152), canine Purkinje cells(166, 410), guinea pig ventricular myocytes (415), rabbitear artery (35), or guinea pig hippocampal CA3 neurons(117). Stimulation of L-type currents via cAMP-dependentpathways provided a positive control for most of thesestudies. Although some studies have reported a smallstimulation of T-type currents by isoproterenol (9, 283),this may be due to either incomplete separation of T- fromL-type currents (as evidenced by BAY K 8644 stimulation)or due to secondary regulation induced by changes ininternal Ca2� concentrations (411). In contrast to theapparent lack of protein kinase A (PKA) regulation, hor-monal regulation of T-type currents has been reported inmany systems. Both inhibition and stimulation have beenreported, and in some cases the same agonist did both.For example, the GABAB agonist baclofen has been re-ported to cause stimulation at 2 �M but inhibition at 100�M (361).

A. Hormonal Inhibition

Many hormones and neurotransmitters have beenreported to inhibit T-type currents, including dopamine,serotonin, somatostatin, opioids, ANP, and ANG II. Al-though most of these studies were performed on isolatedneurons, regulation has also been noted in secretory cellsof the pituitary and adrenal.

Dopamine inhibition of T-type currents has been de-scribed in DRG neurons (260), adrenal glomerulosa cells(312), pituitary melanotrophs (204), lactotrophs (240),and cells from the pars intermedia (309). Chick DRGcurrents were shown to be inhibited 60% by either 10 �Mdopamine or norepinephrine (260). Inhibition occurredwith no apparent effect on kinetics. Recovery upon wash-out was not observed in whole cell recordings. Similarinhibition was observed at the single-channel level, al-though in these experiments washout was observed. Sin-gle-channel current amplitude was not affected, suggest-ing that inhibition was due to a reduction in theprobability of opening (Po).

Dopamine (10 nM) inhibited T-type currents of rat

anterior pituitary lactotrophs by 25–40% (239, 240). Do-pamine shifted the h� curve from �63 to �77 mV (5 Ca)and accelerated current decay during the pulse. It had noeffect on voltage dependence of activation. Dopamine’seffects were reversible, mimicked by the dopamine ago-nist (D2 class) bromocryptine (10 nM), and prevented bythe D2 antagonist sulpiride (100 nM). Inhibitory regulationcould be abolished by omitting GTP from the intracellularsolution or by pretreatment with pertussis toxin. Inhibi-tory modulation could also be disrupted by inclusion ofantibodies that specifically recognize the �-subunit of Go,while antibodies against Gi subunits had no effect (239).Parallel studies showed that D2 receptors coupled to po-tassium channels via a different G protein, Gi-3�.

Dopamine inhibited rat adrenal glomerulosa T-typecurrent by 33% (103). Inhibition was mediated by the D1

receptor class as evidenced by 1) inhibition with the D1

agonist SKF 82958; 2) block by the D1 antagonist SCH23390; 3) but not by the D2 antagonist spiperone. Dopa-mine-mediated inhibition was blocked by agents that dis-rupt signaling by 1) G proteins [2 mM guanosine 5-O-(2-thiodiphosphate) (GDP�S)], 2) adenylyl cyclase (100 �M2,3-dideoxyadenosine), and 3) protein kinase activity(10 �M H-89). Consistent with an action through D1 re-ceptors, dopamine caused a stimulation of cAMP forma-tion. Somewhat surprisingly, bath application of 8-bromo-cAMP had no effect on basal T-type currents but couldblock the effects of dopamine. Addition of 100 nM G��subunits to the pipette solution did not affect T-typecurrents. However, G�� inhibition was observed afteraddition of 1 mM 8-bromo-cAMP to the bath. These re-sults suggested that G�� subunits might bind directly tothe T-type channel.

Somatostatin (10 nM) was found to inhibit T-typecurrents by 50% in rat somatotrophs (72). Somatostatincaused the currents to decay faster during a sustainedpulse and shifted the midpoint of the steady-state inacti-vation curve (h�) curve from �52 to �63 mV (chargecarrier was 5 mM Ca2�). This regulation was abolished bypertussis toxin. Neurotensin (200 nM) and substance P(200 nM) inhibited by 25% the T-type currents from ratcholinergic neurons (263). The �-opioid agonist Tyr-D-Ala-Gly-Met-Phe-Gly-ol (DAGO) was found to inhibit DRGcurrents by 50% (360). Enkephalin was found to inhibitT-type currents 20–40% in the neuroblastoma cell lineNG108–15 (196). In these cells there was no effect ofeither the phorbol ester 4�-phorbol 12-myristate 13-ace-tate (PMA), arachidonic acid (10 �M), or forskolin. Bra-dykinin (0.1 �M) inhibited 57% of the T-type current fromND7–23 cells, a cell line derived from DRG (213). Thisstudy also showed that (�)-baclofen (1 �M) could inhibitcurrents by 56%, whereas guanosine 5-O-(3-thiotriphos-phate) (GTP�S) inhibited by 20%. Pertussis toxin treat-ment did not block baclofen inhibition. Baclofen (20 �M)

134 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 19: Canal Calcio T

has also been reported to inhibit the T-type currents inhippocampal neurons by 53% (126).

Muscarinic inhibition of T-type currents from hengranulosa cells has also been reported (425). In contrastto most studies that report �40% inhibition, this studyreported 90% inhibition. Part of this difference may be dueto their use of the perforated patch technique, which isexpected to preserve the integrity of the intracellularmilieu. The effect of carbachol was blocked by atropine.In contrast, muscarinic agonists have been reported tohave no effect on LVA channels in cholinergic neurons (5).

ANP (3 nM) inhibited T-type currents of bovine ad-renal glomerulosa cells by 28% (274). Inhibition was de-pendent on the holding potential, with little block whenmembrane potential (VM) was �90 mV, but near completeblock when VM was �45 mV. This voltage dependence ofblock also caused a �10 mV shift of the h� curve to morenegative potentials. In contrast, ANP did not affect thevoltage dependence of activation as measured using tailcurrents (26). The effects of ANP were also studied at thesingle-channel level. Single T-type channels were identi-fied by their 8.3-pS conductance. Inhibition was accom-panied by a decrease in the number of active sweeps, withno change in the amplitude of single openings. The abilityof bath-applied ANP to modulate channels recorded in thecell-attached mode indicated that ANP induced the pro-duction of a soluble second messenger.

ANG II and the AT2 selective ligand CGP-42112 inhib-ited by 25% the T-type current in NG108–15 cells (58, 59).Inhibition was associated with a �5-mV shift in the I-V butno effect on the h� curve. Inclusion of 3 mM GDP�S in thepipette blocked the ANG II effect. In contrast, 3 mMGTP�S did not block the ANG II inhibition. This result issurprising since this concentration of GTP�S should havemaximally stimulated G protein signaling. Washout wasnot reported. A role for tyrosine phosphorylation in thesignaling pathway was suggested by the ability of 50 �Msodium orthovanadate to disrupt ANG II modulation ofthe T-type current.

B. Hormonal Stimulation

T-type currents can also be stimulated acutely by avariety of hormones. One of the first studies showed that10 �M serotonin could stimulate rat spinal motoneuronT-type currents by 66% in a slice preparation (36). Also ina slice preparation, dorsal horn neurons were found to bestimulated by substance P (1 �M), although some neuronswere inhibited (347). Cholinergic agonists also stimulateT-type currents, and this was demonstrated at the single-channel level. Carbachol and muscarine doubled the num-ber of openings of single T-type channels in adult guineapig hippocampal CA3 neurons (117). This stimulation wasblocked by 0.1 �M atropine. The modulation was likely

due to an increase in the probability of opening of eachchannel, as there was no change in the single-channelconductance, which was 7.1 pS in a solution containing 95mM Ba2�. Under similar conditions, carbachol inhibitedsingle L-type currents. Carbachol (50 �M) and serotonin(30 �M) were also shown to stimulate hippocampal T-type currents 54 and 61%, respectively (126). These effectswere blocked by the appropriate receptor antagonists andwere quickly reversible. Serotonin appeared to shift theI-V to the left (peak shifted from �25 to �30 mV), but thiseffect was not quantitated (126). Erythropoietin increasedT-type currents 20% in the human neuroblastoma cell lineSK-N-MC (18). It also stimulated increases in intracellularCa2� that was dependent on external Ca2� but not inter-nal stores. Norepinephrine acting via �-adrenoceptorswas reported to increase T-type currents in canine Pur-kinje cells by 69% (410).

Endothelin has been reported to stimulate T-typecurrents from both cardiac and smooth muscle myocytes(129, 181). In neonatal rat ventricular myocytes, endothe-lin-1 displayed an EC50 of 1.3 nM and a maximal stimula-tion of 54% (129). Heat inactivation of the endothelin, oromission of GTP from the pipette, occluded the effects.The stimulation induced by 10 nM endothelin wasblocked by 0.2 �M staurosporine or 20 �M H-7. Proteinkinase C (PKC) agonists stimulated the currents directly,and this could be blocked by H-7 as well. In guinea pigsmooth muscle myocytes, endothelin stimulated the ac-tivity of a 12-pS channel, that may or may not be carriedthrough T-type channels (181). Endothelin has been re-ported to have no effect on T-type currents recorded fromsmooth muscle cells from rat renal resistance arteries(143).

Acetylcholine stimulated T-type currents in NIH 3T3cells that had been transfected with recombinant m3 andm5 muscarinic receptors (322). Stimulation was modest(25%) and was accompanied by a leftward shift in the I-Vrelation. Parallel assays demonstrated that cAMP levelswere increased. T-type currents were also stimulated bytreatments designed to stimulate cAMP-mediated signal-ing, such as 500 �M 8-bromo-cAMP or 10 �M forskolin.Likewise, the PKA inhibitor 10 �M Rp-adenosine 3,5-cyclic monophosphothionate (RpcAMPS) inhibited basalactivity and occluded the response to acetylcholine.These results suggested that T-type channels might bedirectly phosphorylated by PKA. An apparent contradic-tion to this hypothesis was that T-type currents were notregulated in cells transfected with the m1 muscarinicreceptor, although these cells showed robust increases incAMP. The authors speculated that m1 receptors mightsimultaneously activate inhibitory PKC pathways that ob-scured the PKA-mediated stimulation. First they showedthat the PKC activator phorbol 12,13-dibutyrate (PDBu;0.5 �M) directly inhibited T-type currents as observed inother cells types. Next they showed that preincubation

T-TYPE CALCIUM CHANNELS 135

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 20: Canal Calcio T

with the PKC inhibitor calphostin C (0.5 �M) allowedm1-mediated stimulation of the T-type current. Basal T-type currents were not regulated by acetylcholine in cellstransfected with either m2 or m4. However, inhibition ofT-type currents, as well as inhibition of cAMP formation,could be observed in these cells after stimulation of ad-enylyl cyclase by 10 �M forskolin. These results indicatethat hormonal regulation of T-type channels may dependon both basal phosphorylation levels as well as cross-talkbetween PKA and PKC pathways.

Barrett and colleagues (82) have extensively studiedthe stimulation of adrenal glomerulosa T-type Ca2� chan-nels by ANG II. Initial studies established the presence ofT-type channels in these cells through the use of tailcurrent analysis (82). Slowly deactivating currents gatedwith the voltage dependence expected of T-type channels,while fast-deactivating currents behaved like HVA, L-typechannels. ANG II increased the T-type tail currents by 50%(82). The effects of ANG II were also studied at thesingle-channel level (273). ANG II increased both thenumber of active sweeps and increased the number ofopenings per sweep, with no appreciable change in theamplitude of single openings. The ability of bath-appliedANG II to modulate channels recorded in the cell-attachedmode indicated that ANG II induced the production of asoluble second messenger. The effects of ANG II wereblocked by the ANG II receptor antagonist saralasin (1�M). ANG II also shifted the voltage dependence of acti-vation by 10 mV to more negative potentials, with nochange in the h� curve. Pretreatment of the cells withpertussis toxin abolished these effects, indicating thatANG II receptors couple through G proteins, possibly Gi

(251). The final step in this signal transduction pathwayappears to be phosphorylation of the T-type channel bycalmodulin-dependent protein kinase II (CaMKII; see be-low) (27, 252). However, not all studies have found anANG II stimulation of adrenal glomerulosa currents, withreports of either no effect (250) or frank inhibition (103,344).

C. Guanine Nucleotides

As noted above, T-type currents are regulated by Gprotein-coupled receptors. Therefore, this regulationshould require the presence of GTP and should be modi-fied by activators (GTP�S) and inhibitors (GDP�S) of Gproteins. Accordingly, no hormonal regulation was ob-served when GTP was omitted from the patch pipette(129, 239) or when it was replaced with GDP�S (59, 103).

The effects of GTP�S are difficult to interpret be-cause it can activate all G proteins. With the use ofphotoactivation of caged GTP�S in DRG neurons, lowconcentrations were observed to stimulate T-type cur-rents by 54%, while higher concentrations inhibited them

up to 87% (361). Inhibition was accompanied by an accel-eration of inactivation, such that the � decreased from 22to 18 ms. Neither cholera toxin pretreatment nor photore-lease of caged GDP�S had any effect. Pertussis toxinpretreatment blocked GTP�S inhibition, but not stimula-tion. GTP�S has also been shown to inhibit T-type cur-rents (40%) of rat nodose ganglion neurons (148). Thiseffect was blocked by pretreatment with pertussis toxin.

Somewhat surprisingly, G proteins do not appear tomediate the regulation of T-type currents by serotoninand nociceptin. Serotonin inhibition (25%) of T-type cur-rents has been observed in Xenopus sensory neurons(383, 384). Inhibition was not observed when the T-typechannels were recorded in the cell-attached patch config-uration, indicating that its effects were membrane delim-ited. Inhibitory regulation was not blocked by pertussistoxin pretreatment or inclusion in the pipette of GDP�S,GTP�S, or 5-guanylyl-imidodiphosphate (GMP-PNP). Theability of these manipulations to block the regulation ofHVA currents in the same neurons provided a convenientpositive control. It was concluded that G proteins are notinvolved in the serotonin inhibition of T-type channels. Asimilar conclusion was reached in a study on the nocicep-tin inhibition of DRG T-type currents (1). Nociceptin in-hibited currents with an IC50 of 0.1 �M, and in 22 of 77neurons it blocked 100% of the current at 1 �M. Inclusionin the pipette of GTP�S, GDP�S, or AlF3 did not alterresponse, leading the authors to conclude that a G proteinis not involved. It should be noted that nociceptin had noeffect on T-type currents from small nociceptive trigemi-nal ganglion neurons (51). The endogenous cannabinoidanandamide also blocks T-type currents independently ofG proteins and receptors (71).

The ability of pertussis toxin treatment to block reg-ulation indicates that Gi/Go proteins are involved. Supportfor this hypothesis comes from studies that used specificantibodies in the patch pipette to block regulation (239,251). Although the ��-subunits of G proteins have beensuggested to interact directly with T-type channels (103),more studies are required to establish this point.

D. Protein Kinases

With the notable exception provided by studies onfibroblasts (322), T-type currents appear not to be regu-lated by PKA. In contrast, T-type currents appear to beinhibited by members of the PKC family and stimulated bytyrosine kinases and CaMKII.

The effects of PKC have been evaluated by activatingthe kinase directly with compounds such as diacylglyc-erol 1-oleolyl-2-acetylglycerol (OAG), or various phorbolesters: PMA, 12-O-tetradecanoylphorbol-13-acetate (TPA),or PDBu. Although many studies have found no effect ofthese compounds (310), both stimulation and inhibition of

136 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 21: Canal Calcio T

activity have been reported. Studies on rat DRG neuronsfound that 10 nM PMA inhibited the T-type current by27%. Likewise, the inactive analog 4�-phorbol had noeffect. Of note, the PMA effect was not observed at roomtemperature, requiring preincubation at 29°C or higher(359). Similarly, 100 nM TPA was found to inhibit theT-type current by 26% in ventricular myocytes and Pur-kinje cells. A phorbol ester analog that does not activatePKC had no effect, while the protein kinase inhibitor H-8prevented TPA inhibition (411). In contrast, 50 �M H-7was not able to block the inhibition of hippocampal T-type currents by 10 �M TPA or 5 �M OAG (407). Inaddition, the TPA and OAG effects occurred faster thanwould be expected (200 ms), leading Toselli and Lux(407) to conclude that the compounds were blocking thechannel directly. PKC may in part mediate the inhibitoryeffects of arachidonic acid (459).

PKC-mediated stimulation of T-type currents has alsobeen reported (129). Currents in rat neonatal ventricularmyocytes were stimulated 30% by either 0.2 �M PDBu orPMA. The inactive analog 4�-PDBu had no effect. Stimu-lation was blocked by inclusion of 20 �M H-7 in thepipette.

The ability of CaMKII to stimulate adrenal T-typecurrents has been studied at both the whole cell andsingle-channel level, and recently with recombinant chan-nels. Inclusion of Ca-CaM in the pipette shifted the volt-age dependence of activation to more negative potentials(252). The CaMKII inhibitor KN-62 or the synthetic pep-tide inhibitor that corresponds to residues 290–309blocked this effect. Stimulation was also demonstrated atthe single-channel level in both cell-attached and excisedpatches (27). In cell-attached patches, manipulation ofintracellular Ca2� led to an increase in channel activitythat could be blocked by KN-62, but not the inactiveanalog KN-04. Channel activity could also be stimulatedwhen excised inside-out patches were exposed to cal-modulin and appropriate concentrations of Ca2�. Thisstimulation could be blocked with the autocamtide 2-re-lated inhibitory peptide (AIP). Stimulation could also bemimicked by the direct addition of a constitutively activemutant of CaMKII. Heat-inactivated kinase had no effect.As observed previously with ANG II (273), this regulationappeared to be due to an increase in active channels.These studies clearly establish that native bovine glo-merulosa T-type channels are regulated by CaMKII. Adre-nal cortical T-type channels are encoded by �1H, or Cav3.2(358). Recent studies have shown that calcium/calmodu-lin can stimulate the activity of cloned human Cav3.2channels when coexpressed with CaMKII�c in HEK-293cells (443). Stimulation could be blocked by CaMKII in-hibitors added to either the bath (3 �M KN-62) or to theintracellular pipette used for whole cell recording (AIP, 2�M). Replacement of pipette ATP with the nonhydrolyz-able analog 5-adenylyl-imidodiphosphate (AMP-PNP)

also blocked regulation. As observed with native chan-nels, stimulation was accompanied by an 11-mV hyperpo-larizing shift in the voltage dependence of activation, withno shift in the h� curve. In contrast, the activity of humanCav3.1 was not regulated under similar conditions. Thesestudies strongly suggest that the modulatory phosphory-lation site(s) reside directly on the �1-subunit of T-typechannels and that these sites are subtype specific, in thiscase only found on Cav3.2. Although CaMKII regulation ofT-type currents has not been reported in other systems, itmight have mediated the 30% stimulation observed incardiac myocytes (411).

E. Voltage

An intriguing property of many HVA Ca2� currents isthat their activity can be increased, or “facilitated,” bystrong depolarizing pulses. These prepulses can inducechannel conformations that are 1) more susceptible tophosphorylation, 2) have a lower affinity for divalentcations in the pore, or 3) have a lower affinity for Gprotein ��-subunits. The prepulse can also directly inducehigh activity gating modes (mode 2) of cardiac L-typechannels. Both native (below) and recombinant (140)T-type currents have also been reported to be facilitatedby prepulses.

The T-type currents of guinea pig coronary smoothmuscle myocytes were stimulated twofold by 200-ms pre-pulses to voltages above �30 mV (131). Studies where theprepulse potential was varied indicated that saturationoccurs at �10 mV. Use of longer prepulses (10 s) shiftedthis voltage dependency to the left, such that saturationoccurred at �60 mV. Changing the duration of the pre-pulse showed that the peak effect occurred between 160and 320 ms, which then decayed and was gone by 5 s. Thispotentiation was greater when the interpulse voltage was�100 compared with �80 mV, and at 36 versus 24°C.Potentiated currents inactivated faster (�inact 5.6 vs.14.5 ms).

Facilitation has also been observed in bone marrowcells, where T-type currents were stimulated twofold by a750-ms prepulse to �150 mV (330). Compared with theresults obtained in smooth muscle, stronger depolariza-tions were required; stimulation was not observed withprepulses lower than �30 mV, and saturated at �100 mV.Changing the duration of prepulse (�10-mV pulse)showed that half-maximal stimulation occurred at �250ms, and saturated by 1,000 ms. Changing the intervalbetween the prepulse and the test pulse showed thatpotentiation peaked after 1 s and decayed in a biexponen-tial manner: 9.4 and 43.5 s. Potentiation was not affectedby omission of ATP or GTP or bath application of non-specific protein kinase inhibitors (e.g., 100 �M H-7). Po-tentiated currents decayed slightly faster (�inact decreased

T-TYPE CALCIUM CHANNELS 137

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 22: Canal Calcio T

from 17 to 13 ms; currents isolated by subtraction). Theseresults suggest that voltage directly affects channel activ-ity without the involvement of second messenger signals.The opposite conclusion was reached in studies of bull-frog atrial cells (8). Amphibian channels are stimulated by�-adrenergic agonists, and a prepulse enhances this stim-ulation. These effects were enhanced by GTP�S andATP�S, but blocked by either GDP�S or pertussis toxintreatment. From these studies it was concluded that Gproteins are involved and that phosphorylation of G pro-teins may play a role as well.

Facilitation of T-type currents from mouse sper-matogenic cells has also been reported (14). In this case,prepulses could only stimulate activity 50% above control.Prepulses to voltages greater than �30 mV were requiredto observe stimulation, and saturated around �60 mV.Short prepulses (�10 ms) were capable of stimulatingcurrents, and this effect wore off very slowly. Membrane-permeable inhibitors of tyrosine kinases (tyrphostins A47and A25) stimulated basal T-type currents but had noeffect on facilitated currents. This suggests that tyrosinekinases may tonically inhibit activity in this system andthat a prepulse somehow reverses their activity. Consis-tent with this hypothesis, inhibitors of tyrosine kinasephosphatases inhibited basal T-type currents and pre-vented voltage-induced potentiation.

In summary, T-type channels can be both stimulatedand inhibited by hormones and neurotransmitters that arecoupled to the Gq or Gi/Go families. Regulation appears tobe mediated by phosphorylation, with inhibition mediatedby PKC and stimulation by CaMKII. Additional studies arerequired to establish the location of the modulatory phos-phorylation site(s), whether G protein subunits are capa-ble of interacting directly with the channel, and the mech-anisms by which voltage regulates channel activity.

IV. MOLECULAR CLONING

OF T-TYPE CHANNELS

A. Cloning

Cloning of high voltage-gated Ca2� channels beganwith purification and microsequencing of the skeletalmuscle dihydropyridine receptor (396). Low-stringencyscreening of cDNA libraries with probes derived from the�1-subunit of skeletal muscle (�1S, or Cav1.1) led to thediscovery of cardiac (�1C, Cav1.2) and brain (�1D, Cav1.3)isoforms of the L-type family (Cav1; reviewed in Ref. 326).The more distantly related members of the Cav2 familywere cloned during the “brain era” (Fig. 4A) using evenlower stringency hybridization (374). PCR provided analternative approach, and primers based on highly con-served regions (IIIS5 and IVS5) were used successfully toisolate a fragment of Cav2.3 (357). Despite the efforts of

many groups, these approaches were not successful inisolating cDNAs encoding T-type channels.

Progress in the sequencing of human, yeast, and C.

elegans genomes provided a novel library that could bescreened with a computer, or in silico (Fig. 4B). Screeningwith the highly conserved P loop, or K� channel signaturesequence, allowed Ketchum et al. (205) to identify a novelK� channel sequence from Saccharomyces cerevisiae. Wedecided to try a similar approach using S5 segments. Thisled to the identification of two Ca2� channel-like proteinsthat were predicted from the C. elegans genome (325).These proteins were homologous to the skeletal muscleCa2� channel, and this similarity was noted in the Gen-Bank entry. We reasoned that there might be other se-quences in the GenBank that were similarly labeled andused a text-based search (Fig. 4B) to find sequences thatwere “similar to a calcium channel.” This led to the iden-tification of the human EST clone H06096 (324). Althoughthis clone is 1.4 kb long, only a part of the 5-sequence wasin the database. This fragment had a very low sequenceidentity to the S1 segment of other Ca2� channels, pre-cluding its direct identification by homology search. Thefull sequence of H06096 showed that it might encode aprotein with many of the hallmarks of voltage-gated chan-nels, notably the S4 voltage sensor sequence was wellconserved, and the pore loop sequence was highly homol-ogous to other Ca2� channels. Screening of the GenBankwith the H06096 sequence and the program BLAST iden-tified C54d2.5 as a C. elegans homolog. Similar to in vitrocloning where one “walks” through a cloning project us-ing probes to the ends, we screened the GenBank with the3-end of C54d2. This led to the identification of the ESTclone H19230. In vitro screening of a rat brain cDNAlibrary with H06096 allowed the cloning of Cav3.1 (324),whereas use of this clone to screen a human heart libraryled to the cloning of Cav3.2 (90). Screening of a rat brainlibrary with H19230 under low-stringency conditions al-lowed us to identify not only rat versions of Cav3.1 andCav3.2, but also Cav3.3 (228). A similar approach was usedby Monteil et al. (288) to identify C54d2.5 and the H06096and H19230 ESTs, which then allowed them to clone thefull-length human Cav3.1. Similar PCR approaches werealso used to clone a human Cav3.2 cDNA (438) and mouseand rat versions of Cav3.1 (211, 463). PCR approachesbased on the C. elegans sequence C27f2.5 led to thecloning of a distantly related channel gene whose func-tion remains unknown (227).

Sequencing of the human genome also played animportant role in the cloning of T-type channels. TheCav3.3 gene is located on chromosome 22, which was thefirst chromosome to be completely sequenced (105).BLAST search with rat Cav3.1 identified the coding re-gions of Cav3.3, which then allowed the design of PCRprimers to either clone its cDNA (228, 289) or to examine

138 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 23: Canal Calcio T

the splicing patterns of its gene (285). Proper identifica-tion of the 5- and 3-ends requires traditional cDNA clon-ing or specially designed PCR protocols. Recent studiesindicate that the human Cav3.3 mRNA contains an addi-tional exon that encodes 214 additional amino acids (140)than previously recognized (289). Studies comparing thefull-length to truncated Cav3.3 channels revealed differ-ences in channel expression and their ability to be mod-ulated by a strong prepulse (140).

Genetic mutations in various Ca2� channel subunitgenes have been associated with ataxic and epilepticphenotypes (187). As a first step in exploring such a link,we mapped the human and mouse genes for Cav3.1,CACNA1G (324), and Cav3.2, CACNA1H (90). The geneswere mapped using the Cambridge 4 radiation hybridpanel and fluorescent in situ hybridization. The locationsof the human genes are shown in Table 8. Mutations inhuman Cav3 genes have yet to be described.

All three T-type channel genes are alternativelyspliced, producing variants that differ in their intracellularloops (68, 284, 285, 463). Splicing produces Cav3.1 vari-ants that differ in at least two locations: 1) the II-III loop,where two variants are produced by insertion/deletion of

exon 16; and 2) the III-IV linker, where three (or actuallyfour, Ref. 229) variants are produced by a combination ofinsertion/deletion of exon 26 and use of an alternate5-splice site in exon 25 (288). These variants differ intheir electrophysiological properties (68) and may ac-count for the differences noted in the gating of T-typechannels between lateral geniculate (LGN) relay neuronsand interneurons (318). Splice variants of Cav3.3 havebeen detected by PCR that differ in the I-II loop andcarboxy terminus (285). Similarly, PCR has identifiedsplice variation in the III-IV linker of Cav3.2 in both rat(231) and human (182).

Cav3 cDNAs have been cloned from rat, mouse, andhuman sources (Table 8). Although all three have beencloned from rat and human, cloning from mouse is stillongoing. Comparisons of their nucleotide sequences re-veal a high level of conservation between species. Al-though species differences in the predicted amino acidsequence have been noted, some of these changes can beascribed to either frameshifts in the reading, cloning ar-tifacts, or differences in splicing. For example, it is likelythat a cloning artifact caused the deletion of three aminoacids from IIIS4 in a rat cDNA of Cav3.3 (279).

FIG. 4. A timeline of Ca2� channel cloning. A: graphillustrates the time of the first published reference to thecloning of each of the 25 subunits. Skeletal muscle Ca2�

channel subunits were purified, microsequenced, thencloned using degenerate oligonucleotide probes based onthe protein sequence. This initial period has been calledthe Muscle Era. Low-stringency screening of brain cDNAlibraries with the skeletal muscle probes led to the cloningof other �1- and �-subunits during the Brain Era. In thefinal epoch, or In Silico Era, subunits were first identifiedby database searching programs (BLAST) of Human andCaenorhabditis elegans genome sequencing projects, thencloned by standard in vitro methods. (Adapted from Ran-dall A and Benham C. Recent advances in the molecularunderstanding of voltage-gated Ca2� channels. Mol Cell

Neurosci 14: 255–272, 1999.) B: flow diagrams describingtwo ways of cloning in silico.

T-TYPE CALCIUM CHANNELS 139

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 24: Canal Calcio T

B. Structure

Conservation of amino acid sequence (Figs. 1 and 5)and predicted secondary structure (Fig. 6) indicates thatT-type channels are evolutionarily related to the �-sub-units of K�, Na�, and HVA Ca2� channels (184). Althoughlittle is known about the structure of Na� and Ca2�

channels, recent advances with K� channels provide aframework to interpret their structures. The most signif-icant advance was the determination of the KcsA K�

channel structure by X-ray crystallography (102). Thischannel contains two transmembrane segments sepa-rated by a pore loop (Fig. 6). The second transmembranesegment forms the barrel walls of the pore, while theselectivity filter is formed by the pore loops that dippartially into the barrel. In addition to this fundamentalmotif, voltage-gated channels also contain four additionaltransmembrane segments, with one (S4) acting as a volt-age sensor. In K� channels, four proteins assemble toform one channel, while in Na� and Ca2� channels asingle protein contains four of these modules, or repeats.Each transmembrane segment can be identified by a num-ber to indicate the repeat it is in (traditionally a Romannumeral) followed by its position within a repeat (Fig. 6).For example, the third transmembrane segment of thethird repeat can be abbreviated IIIS3. Among voltage-gated ion channels, the most highly conserved sequencebelongs to the S4 segments, where every third residue ispositively charged. This region is thought to move out-

ward when the plasma membrane is depolarized. Little isknown about how this movement is transduced into chan-nel opening. The pore loops are also well conserved,especially within the members of a particular ion channelclass. In voltage-gated Ca2� channels, the pore loops allcontain the following signature sequence: T-x-E/D-x-W. InHVA channels, all four repeats have a glutamate residue inthe pore, leading to the abbreviation EEEE. In repeats IIIand IV of Cav3 channels, the glutamate has been replacedby an aspartic acid residue, or EEDD. The importance ofthese residues in Ca2� permeation has been extensivelyestablished in HVA channels (449), and recently withCav3.1 (391).

Site-directed mutagenesis studies on the S6 region ofCav3.1 suggest that these segments play a similar struc-tural role in these channels as in KcsA (266). This pro-vides a theoretical basis for the use of the KcsA crystalstructure to guide studies of drug binding pockets, whichin the case of Cav1.2 appear to be located within thechannel walls on the cytoplasmic side of the pore loops(174).

As shown in Figure 6, most of the channel is thoughtto be intracellular. The extracellular loops are predictedto be quite short, with the notable exception of the loopconnecting IS5 to the pore loop (98–103 amino acids).This extracellular loop contains six cysteine residues thatare conserved in all three Cav3 channels and a number ofpossible sites for glycosylation. Neuraminidase treatmentof cardiac myocytes stimulates T-type (but not L-type)

TABLE 8. Comparison of the three cloned T-type channels

Cav3.1 Cav3.2 Cav3.3Reference

Nos.

Alias �1G, CavT.1 �1H, CavT.2 �1l

Gene name CACNA1G CACNA1H CACNA1l 249Chromosomal location 17q22 16p13.3 22q12.3–13.2 90, 228, 324Database entries*

Human NP_061496, O43497 AAC67239, O95180 XP_001125, Q9UNE6Rat AAC67372, O54898 AAG35187, Q9EQ60 NP_064469, Q970Y8Mouse NP_033913, Q9WUB8 NP_67390

Tissue expression Brain, ovary, placenta, heart Kidney, liver, adrenal cortex,brain, heart

Brain 90, 228, 288,289, 324

Electrophysiology†Threshold, mV �70 �70 �70 209Peak of I–V, mV �30 �30 �30 209Activation kinetics, ms 1 2 7 209Inactivation kinetics, ms 11 16 69 209Deactivation kinetics, ms 3.0 2.2 1.1 209h�, mV �72 �72 �72 209Recovery from inactivation, ms 117 395 352 209Conductance, pS; Ba 7.5 9 11 228, 324, 438

PharmacologyNickel (IC50, �M) 250 12 216 230Mibefradil (IC50, �M) in 10 Ba2� 1.2 1.1 1.5 267Mibefradil (IC50, �M) in 2 Ca2� 0.27 0.14 267

* Accession numbers are for a representative GenBank entry followed by the SWISS-PROT entry. † Results for conductance were obtainedin 110 mM Ba2�, while all other electrophysiological measurements were made in 1.25 mM Ca2�. Activation and inactivation kinetics were measuredwith exponential fits to the currents elicited by test potentials to �10 mV. Deactivation kinetics were measured at �90 mV. Pharmacology wasmeasured in 10 mM BaCl2, or as noted, in 2 mM CaCl2 solutions.

140 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 25: Canal Calcio T

currents threefold, suggesting that glycosylation of thechannel inhibits its activity (116, 451). Despite the factthat the sequences of intracellular loops differ consider-ably across Na� and Ca2� channels, they have a similaroverall size: long loops connect repeats 1–2 and 2–3(range 207–375 amino acids for Cav3 channels), whilerepeats 3–4 are connected by short loops (55–62 aminoacids). Similarly, the amino termini are usually short(�100 amino acids), whereas the carboxy termini arequite long (433–493 amino acids). The sequence of theseloops is poorly conserved, with the exception of se-quences close to the end of the S6 segments (Fig. 5). InHVA channels these loops are important for binding aux-iliary subunits such as the �-subunit, binding to regulatoryproteins such as calmodulin or the G protein ��-complex,and are sites for protein phosphorylation (424). Althoughthe roles of these loops in T-type channels are presentlyunknown, it is likely they contain important regulatorysites as well.

C. Distribution

The expression patterns of these genes in peripheraltissues and brain have been studied using Northern blots,RNA dot blots, and in situ hybridization (Table 8). Cav3.1mRNA is primarily expressed in human brain, but also inovary, placenta, and heart (288, 324). A wider distributionwas noted in dot blots prepared from human fetal tissues,with strong expression noted in kidney and lung (288).Cav3.2 mRNA is primarily expressed in kidney and liver,but also in heart, brain, pancreas, placenta, lung, skeletalmuscle, and adrenal cortex (90, 438). Cav3.3 mRNA isalmost exclusively expressed in brain (228); the only pe-ripheral tissues that showed expression on dot blots wereadrenal (but see Ref. 358) and thyroid (289).

The distribution of all three Cav3 channel transcriptsin rat brain has been studied in exquisite detail using insitu hybridization (393), and similar patterns of expres-sion have been obtained using Northern blots of humanbrain mRNA (288, 289, 438). Their patterns of expressionare complementary in many regions, with most brainregions expressing more than one isoform. In fact, someneurons may express all three genes as suggested by thelabeling of olfactory granule cells and hippocampal pyra-midal neurons. Many brain regions showed heavy expres-sion of Cav3.1 mRNA, including thalamic relay nuclei,olfactory bulb, amygdala, cerebral cortex, hippocampus,hypothalamus, cerebellum, and brain stem. Two separatestudies have confirmed this pattern of mRNA expressionand extended the findings by showing a similar distribu-tion of Cav3.1 protein (88, 197). Cav3.2 mRNA expressionwas detected in olfactory bulb, striatum, cerebral cortex,hippocampus, and reticular thalamic nucleus. Cav3.3mRNA expression is high in olfactory bulb, striatum, ce-

rebral cortex, hippocampus, reticular nucleus, lateral ha-benula, and cerebellum. DRG neurons express bothCav3.2 and Cav3.3 mRNA (393). One study found that thisexpression was restricted to small and medium-sized neu-rons (393), while a second study reported that Cav3.3transcripts were equally abundant in large DRG neurons(454).

D. Electrophysiology of Recombinant Channels

Expression of recombinant Cav3 channels leads tothe appearance of robust currents in a variety of heterol-ogous expression systems. Currents mediated by thecloned �1-subunit are nearly identical to those recordedfrom native channels: they both open and inactivate atpotentials near the typical resting membrane potential,recover rapidly from inactivation, and close slowly pro-ducing prominent tail currents (Fig. 7).

The currents mediated by rat and human isoforms ofthe three Cav3 channels have been compared under avariety of experimental conditions (209, 218, 228, 279). Ingeneral these results agree quite well with each other,with the differences most likely due to recording condi-tions or sequence variability. For example, Cav3.1 andCav3.2 currents inactivate (�1.7-fold) faster with Ba2�

than Ca2� (209, 211). Similarly, recording conditions suchas choice of charge carrier or length of depolarizingpulses can affect estimates of both deactivation kinetics(211, 216, 430) and recovery from inactivation (430).Therefore, comparisons under one set of conditions arethe most relevant; Table 8 presents the results obtained in1.25 mM Ca2� (209), whereas Figure 7 results were ob-tained in 5 mM Ca2�.

All three Cav3 clones form LVA channels. Depolariza-tion of the membrane to �70 mV is sufficient to triggerchannel opening, and the I-V curves for all three channelspeak around �30 mV (Fig. 7B). Some studies have foundthat Cav3.3 currents activate at slightly more positivepotentials than either Cav3.1 or Cav3.2 (127, 289). Part ofthis difference can be ascribed to the methods used toestimate the voltage dependence of activation. Of note,not all the recombinant channels within a given prepara-tion gate at the same low potentials, with �40% of thechannels requiring much stronger depolarizations (�10 to�50 mV) for opening (288, 438). Activation curves con-structed using tail currents that detect both channelsshow a lower slope and higher midpoint of activation thanother methods.

The kinetics of the currents are very voltage depen-dent between �70 and �20 mV, producing a typical criss-crossing pattern of traces (Fig. 7A; Ref. 334). Therefore, itis useful to compare the kinetics of the three channels at�10 mV or above (Table 8). Cav3.1 and Cav3.2 both acti-vate relatively quickly at this potential (� � 1–2 ms) and

T-TYPE CALCIUM CHANNELS 141

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 26: Canal Calcio T

FIG. 5. Alignment of the predicted amino acid sequence of human Cav3.1, 3.2, and 3.3 channels. Residues conservedin all three sequences are shown in the consensus sequence (CON). The approximate location of the membrane-spanningregions is indicated above the sequence. �-Subunits bind to the I-II loop of HVA channels through a sequence termed the�-interaction domain (AID). The consensus sequence of the AID is shown above the homologous region in the Cav3channels. The residues are colored according to the structural properties of the individual amino acids: red, positivelycharged; green, negatively charged; blue, polar; yellow, nonpolar.

142 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 27: Canal Calcio T

inactivate 10-fold slower (� � 11–16 ms). In stark con-trast, Cav3.3 currents activate and inactivate very muchslower. This difference is even greater when comparisons

are made using Xenopus oocytes (228), and it remains amystery why Cav3.3 channels behave so differently inamphibian eggs than in mammalian cells.

FIG. 5—Continued

T-TYPE CALCIUM CHANNELS 143

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 28: Canal Calcio T

Recombinant Cav3 channels, like their native coun-terparts, close slowly producing prominent tail currents.The voltage protocol used to measure tail currents in-cludes a short pulse to open channels, followed by repo-larization to a voltage where they close (Fig. 7C). Al-though all three Cav3 channels show slow tail currentsupon repolarization to �90 mV, Cav3.1 are the slowest(� � 3 ms), while Cav3.3 are the fastest (� � 1 ms). Incontrast, HVA channels close 10-fold faster.

Despite large differences in the rate at which theyinactivate, the voltage dependence of steady-state inacti-vation of the three recombinant Cav3 channels is remark-

ably similar. The midpoint of the h� curves are �72 mV(Table 8). Native T-type currents inactivate over a similarrange. These studies indicate that T-type channels, asobserved with other channels, can reach inactivatedstates without passing through open states (364). In fact,during a depolarizing pulse up to 30% of Cav3.3, channelsinactivate without opening (127).

Cav3 channels recover rapidly from short-term inac-tivation. Cav3.1 channels recover fastest, displaying amonoexponential recovery with a time constant of �100ms (Fig. 7, Table 8). In contrast, Cav3.3 channels recoverthreefold more slowly, while Cav3.2 channels recover

FIG. 6. Structural models. A: hydrop-athy plots of the KcsA and Shaker K�

channels and repeat III of Cav3.1. Thehydropathy index varies from 3 to �1.Letters above the plots indicate the rela-tive position of the putative transmem-brane segments (S1–S6) and pore loops(P). B: transmembrane segments are alsotypically displayed in two dimensions assnake diagrams, wherein the protein isshown to snake its way in and out of themembrane. C: a model of the three-di-mensional structure of four-repeat ionchannels. The repeats are assumed to beoriented in a clockwise manner based onwork with Na� channels. The S5-P-S6segments are thought to form a similarstructure as the S1-P-S2 segments ofKcsA, with the S6 segment forming thewalls of the channel. The S4 segment ishighly charged, so it is likely to be sur-rounded by the other segments with wa-ter filling the cracks. The S1–S3 segmentsare shown to form the exterior of thechannel. The transmembrane segmentsare most likely composed of �-helices,but we have no clue to their orientation.The intra- and extracellular loops are notshown for clarity. D: a snake diagramwhere each amino acid is represented bya circle. Residues that are conserved inall three Cav3 channels are shown with asolid circle. (Adapted from Perez-ReyesE. Three for T: molecular analysis of thelow voltage-activated calcium channelfamily. Cell Mol Life Sci 56: 660–669,1999.)

144 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 29: Canal Calcio T

even more slowly (209). The channels also differ in theirrecovery from steady-state inactivation, and again, Cav3.2was found to recover the slowest (209). These resultssuggest the existence of multiple inactivated states, asnoted previously for T-type currents from sensory neu-rons (53). These results also suggest that recovery kinet-ics can be used to differentiate currents carried by nativeCav3.1 from Cav3.2 (351). Fast recovery from inactivationis a critical property of native T-type channels that allowsthem to participate in rebound burst depolarizations.

The ability of T-type channels to open at similarpotentials at which they inactivate suggests that theymight generate current under steady-state conditions.This is commonly referred to as a window current and is

operationally defined as the overlap region between theactivation and steady-state inactivation curves (Fig. 7, G

and H). All three Cav3 channels are predicted to generatewindow currents supported by �1% of the channels. Re-cording conditions and methods of analysis have a dra-matic effect on estimates of the window current, so theseestimates must be interpreted with caution. For example,analysis of Cav3.3 results obtained in Xenopus oocytesindicates that up to 2% of the channels can be open in thesteady-state. Evidence for window currents have alsobeen obtained in thalamacortical neurons, where theyplay an important role in signal amplification (441). Win-dow currents also appear to be important in determiningintracellular Ca2� concentrations (18, 69, 264). Expres-

FIG. 7. Electrophysiological properties of recombinant Cav3 channels. Recombinant channels were expressed inhuman embryonic kidney cells (293) and recorded with the ruptured patch method using 5 mM Ca2� as the chargecarrier. A: the current-voltage (I-V) relationship is typically measured using a series of step depolarizations where the testpotential (�80 to �10 mV) is increased 10 mV between runs. After each run the cell is held at �100 mV for at least 10 sto allow recovery. Recordings were made using a stably transfected cell line expressing Cav3.2. B: I-V curve is a plot ofthe current as a function of test potential. In this example the currents for each cell were normalized to the peak currentobserved, and then averaged. Typical I-V relations are shown for human Cav3.1 (‚), 3.2 (ƒ), and 3.3 (E; same symbols forall panels). The data were fit with an equation: G � Gmax (Vm � Vrev)/[1 � exp(V1/2 � Vm)/k], where G is conductance,Vm is the test potential, Vrev is the apparent reversal potential, V1/2 is the midpoint of activation, and k is the slope factor.C: tail currents are measured with a two-step protocol where the first step is designed to open channels (e.g., 2 ms to�60 mV), and the second step measures how fast they close at different repolarization potentials. Traces were obtainedwith Cav3.1 and Cav3.3 after repolarization to �90 mV. Deactivation kinetics are measured by fitting tail currents withexponential functions. D: deactivation kinetics vary with the repolarization potential. E: recovery from inactivation canbe measured using a three-step protocol where step 1 is a long depolarizing pulse used to inactivate channels, step 2 isa repolarizing step to voltages where channels recover, and step 3 is a test pulse to measure the fraction of channels thatrecovered. The duration of step 2 is increased in subsequent runs. Recovery is defined as the ratio of the peak currentin step 3 divided by the current in step 1. F: recovery time courses of the three human Cav3 channels. Data were fit withan exponential to calculate the rate of recovery. Cav3.1 channels recover fastest with a � of �100 ms, Cav3.3 areintermediate (250–300 ms), while Cav3.2 recover slowest (�400 ms). G: steady-state inactivation curves of the threehuman Cav3 channels estimated with 15-s prepulses. Smooth curves represent fits to the data with a Boltzmann equation.The foot of the activation curve is also shown. There is a small voltage range where these two curves overlap, leadingto the prediction that channels can produce steady-state (window) currents. H: fraction of channels open in thesteady-state can be estimated by multiplying the Boltzmann functions that describe activation and inactivation.

T-TYPE CALCIUM CHANNELS 145

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 30: Canal Calcio T

sion of recombinant Cav3.1 or Cav3.2 was found to in-crease basal Ca2� in HEK-293 cells, and this increase wasblocked by appropriate concentrations of either mibe-fradil or nickel. Similarly, basal Ca2� is increased afterdifferentiation of the human prostatic cancer cell lineLNCaP, and this correlates with the expression of Cav3.2mRNA and currents (264). Hormonal regulation of theT-type window current provides cells with another mech-anism to regulate intracellular Ca2�. In adrenal glomeru-losa cells, ANG II increases T-type window currents byselectively shifting activation to more negative voltages(273, 443), an effect mediated by CaMKII phosphorylation(discussed in sect. IIID).

Another distinguishing property of T-type channels isthat they have a lower Ba2� conductance than HVA chan-nels. Therefore, studies of cloned T-type channels havefocused on the amplitude of single-channel currents inisotonic barium solutions. All three recombinant channelswere found to have small single-channel currents, corre-sponding to slope conductances in the 7–11 pS range(Table 8). In contrast, Cav1.2 channels display slope con-ductances of 20–30 pS, while Cav2 channels are in the13–16 pS range. Measurement of the single-channel am-plitude is complicated by the presence of subconductanceactivity (228). Failure to account for these subconduc-tance openings would result in lower estimates, and thismay explain why Cav3.2 was reported to have a conduc-tance of 5.3 pS (90). Subconductance activity has alsobeen observed with native T-type channels (64).

The effect of changing external pH has been studiedwith Cav3.1 (391) and Cav3.2 (95). In contrast to nativechannels, whose activity is affected by small deviationsfrom pH 7.2, the recombinant channels are largely unaf-fected by pH changes in the 8.2 to 6.9 range. Furtheracidification of the external solution to pH 5.5 decreasesthe currents measured during standard test pulses. Muta-tion of the aspartate in the dmain III pore loop to gluta-mate (EEDD to EEED) shifted the apparent pKa of Cav3.1from 6.0 to 7.0 (391). In addition,this mutation increasedthe sensitivity to Cd2� block. Therefore, each domaincontributes a negatively charged residue to form a high-affinity binding site for Ca2� (and Cd2�) in Cav3 channels(EEDD) as shown previously for Cav1 and Cav2 (bothEEEE) channels (75, 207, 319, 449). In addition to affect-ing permeation, shifting the pH from 8.2 to 5.5 produced adramatic 40-mV shift in the voltage dependence of activa-tion of Cav3.2, lowered its slope factor, and produced an11-mV shift in the h� curve (95). These results suggestedthat protonation of the channel affected the voltage de-pendence of transitions between deep closed states (R toC3 transitions in Fig. 3). Although it is unknown where onthe channel the affected residues lie, they may very wellbe in the pore. Consistent with this hypothesis, mutationsin the Cav3.1 pore also lowered its voltage dependenceand shifted activation to more positive potentials (391).

Studies using mock action potentials suggest thatmost of the Ca2� influx through T-type channels occursafter the peak voltage and corresponds to the tail current(218, 276, 334). This Ca2� influx might play a role in spikerepolarization and afterhyperpolarization mediated byapamin-sensitive Ca2�-dependent K� channels (440). Dueto differences in their rates of activation and inactivation,the three T-type channels respond quite differently totrains of mock action potentials (218). Both Cav3.1 andCav3.2 channels inactivate rapidly during trains deliveredat �20 Hz. In contrast, Cav3.3 continues to gate during100-Hz train frequencies. These differences suggest thatthe three channels play distinct roles in determining neu-ronal firing patterns and signal integration. Cav3.1 andCav3.2 channels are likely involved in short burst firing asseen in thalamacortical neurons, whereas Cav3.3 channelsare likely involved in sustained burst firing as seen inthalamic reticular neurons (70).

E. Auxiliary Subunits?

Auxiliary subunits of HVA Ca2� channels were firstidentified in purified preparations of skeletal musclechannels (see Ref. 326 for review). This channel complexwas found to contain at least four subunits: �1 (Cav1.1),�2� (�2�1), � (�1), and � (�1). Purified cardiac L-type andbrain N-type channels have a similar �1�2�� structure, butwith distinct �1- and �-subtypes. Additional members ofeach subunit class have been cloned (Fig. 4), such thatthere are now seven HVA �1-, four �-, three �2�-, and eight�-like subunits (reviewed in Ref. 171). The �-subunits playa critical role in channel function, controlling both chan-nel expression at the plasma membrane, voltage depen-dence, and pharmacology. The �2�- and �-subunits alsoplay a role in expression of functional channels. Muta-tions in Ca2� channel genes, or calcium channelopathies,have been linked to a variety of disease phenotypes inboth mice and humans (reviewed in Ref. 248).

Native T-type channels have yet to be purified, sotheir subunit structure is unknown. Two strategies havebeen used to infer its structure: to assay whether anti-sense oligonucleotides directed against HVA �-subunitscan alter native T-type channel function, and to testwhether the activity of the cloned T-type �1-subunit isaffected by coexpression with cloned HVA subunits. An-tisense depletion of �-subunits causes profound reduc-tions in HVA channel density; however, this treatment hasno effect on T-type channels recorded from the sameneurons (226, 234). Similarly, coexpression of cloned�-subunits has little or no effect on cloned T-type channelactivity (100). In contrast, coexpression with either �2�1

or �2�2 was found to double the size of Cav3.1-mediatedcurrents (100, 132). Concomitant immunolocalizationstudies supported the contention that this rise in currents

146 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 31: Canal Calcio T

was due to increased expression of �1-�2�1 complexes atthe plasma membrane (100). Similar results were ob-tained by coexpression of a �2�2 with mouse Cav3.1 (169),although the effects of �2�1 failed to reach statisticalsignificance (224). Coexpression with �2�1 causes no de-tectable changes in the biophysical properties of Cav3.1channels, while �2�2 has been reported to have minoreffects on inactivation (169). Mutations in the murine �2�2

gene have been linked to an absence of epilepsy pheno-type in the mouse strain ducky (23). Recordings fromPurkinje neurons isolated from ducky cerebella showedhalf the P-type current density observed in controls.These results, in combination with coexpression studies,suggest that �2�2 plays a greater role in HVA than LVAchannel expression.

Identification of the mutation in the mouse strainstargazer, which also exhibits an absence epilepsy phe-notype, led to the cloning of �2 (233). In silico cloning hasextended the �-like family to eight members (78). The�2-protein is only 25% identical to skeletal muscle � andmay play different physiological roles. Immunoprecipita-tion studies have shown that �2 may be a subunit of bothHVA Cav2.2 Ca2� channels and glutamate receptors of theAMPA class (74, 190, 366). Coexpression of �2 inhibits theexpression of Cav2.1 and Cav2.2 channels in the presenceof �2�1 and has minor effects on their biophysical prop-erties (190, 346). In contrast, �2 had little effect on Cav3.3expression, although it could slow the decay of the tailcurrent (145). The related isoforms, �3 and �4, have beenreported to have no effect on Cav3.3 currents (145). Incontrast, �2, �4, and �5 have been reported to accelerateinactivation of Cav3.1 currents and shift steady-state in-activation by �4 mV (210). To reiterate, the electrophys-iological activity of T-type �1-subunits expressed alone isvery similar to that observed with native channels. This isnot the case for HVA channels, which require �-subunitsfor proper gating (223), and �2�1- and �1-subunits forproper pharmacological activity (381). In conclusion,these studies indicate that HVA auxiliary subunits caninteract with Cav3 �1-subunits, but additional biochemicalstudies are required to establish this interaction in vivo.

V. PHARMACOLOGY

Cardiovascular L-type calcium channels are an estab-lished drug target in the control of blood pressure andcardiac rhythm. Selective Cav1 blockers have been devel-oped from over three drug classes, including dihydropy-ridines, phenylalkylamines, and benzothiazepines. In con-trast, there are no highly selective drugs or peptide toxins(369) that selectively block Cav3 channels. A comprehen-sive review of drugs tested on native T-type channels wasrecently published (158); therefore, this review focuseson drugs whose mechanism may involve block of T-typechannels.

A. Antihypertensives

Marketing of mibefradil (Posicor, Roche) as the firstselective T-type channel blocker stimulated considerableinterest in the pharmacology and physiology of thesechannels (412). Mibefradil was approved for use in essen-tial hypertension and stable angina pectoris until it waswithdrawn from the market due to drug-drug interactionsleading to irregular heart rhythms (219). Submicromolarconcentrations of mibefradil, formerly Ro-40–5967, pref-erentially block T-type channels, while higher concentra-tions are required to block cardiovascular L-type channels(282). Its 10- to 30-fold selectivity for T- over L-type chan-nels has been confirmed in a number of preparations,including recombinant channels expressed in mammaliancells (267, 323). The average IC50 from 11 studies formibefradil block of native T-type channels was 1.5 �M(see Table 1 in Ref. 267). Mibefradil potency is affected bythe concentration, and choice, of divalent cation, withless block in Ba2� than in Ca2� solutions (267). In 2 mMCa2� solutions, mibefradil blocked recombinant Cav3.2channels with an IC50 of 0.14 �M, whereas slightly higherconcentrations were required for block of Cav3.1 andCav3.3 (267). Mibefradil block of both HVA and LVA chan-nels is state dependent (41, 278). Single-channel measure-ments indicate the presence of open channel block, and inaddition, a longer lasting block that results in a reductionin active sweeps (280). Whole cell measurements indicatethat inhibition is enhanced by holding the membrane atpotentials where T-type channels inactivate, suggestingthe drug has higher affinity for inactivated states. Theapparent affinity for these states was estimated to be 83nM using native T-type currents (278), and similar resultswere obtained with recombinant channels (267). This pro-vides another mechanism for selectivity, since T-typechannels inactivate near the resting membrane potentialof most cells, while HVA channels require depolarization.A similar mechanism has been invoked to explain theselectivity of dihydropyridines for L-type channels in de-polarized vascular beds over L-type channels in heart.These results suggest that mibefradil selectivity in vivowas probably greater than estimated in vitro since moststudies used Ba2� solutions, and a significant fraction ofT-type channels is inactivated in vivo. Therapeutic plasmaconcentrations of mibefradil were found to be in the 0.5–1�M range, with 99.5% bound to plasma proteins such as�1-acid glycoprotein (433). Therefore, a significant frac-tion of T-type channels in vascular smooth muscle myo-cytes would have been blocked during treatment. Subse-quent studies have established that micromolarconcentrations of mibefradil are capable of blocking awide range of ion channels, including those for Na� (106),K� (138), and Cl� (304). It remains to be determinedwhether block of these other channels contributed to itsantihypertensive activity. This ability to block other ion

T-TYPE CALCIUM CHANNELS 147

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 32: Canal Calcio T

channels complicates the interpretation of mibefradil ef-fects, and additional studies are required to establish therole of T-type channels in smooth muscle proliferation (355),cardiac remodeling (113, 422), and renal protection (28).

Akaike et al. (4) have extensively studied the sensi-tivity of neuronal T-type currents to calcium channelblockers. These studies demonstrated that many drugsblocked neuronal T-type currents at micromolar concen-trations. The order of potency (IC50 in parentheses, in�M) was flunarizine (0.7) � nicardipine (3.5) � nifedipine(5) � nimodipine (7) � methoxyverapamil (50) and dilti-azem (70). A similar rank order was found in a study ofseptal neurons, except nicardipine was found to be morepotent (IC50 � 0. 77 �M; Ref. 6). Because early studiesusing 0.1–1 �M dihydropyridines had shown selectiveblock of L-type channels with little block of T-type chan-nels, these authors concluded that there was a subset ofdihydropyridine-sensitive T-type channels. A similar con-clusion was reached from a study of thalamic and mes-enteric SMC T-type currents (246, 398). Felodipine is alsoa potent T-type blocker, displaying an apparent affinity forinactivated channels of atrial myocytes of 13 nM, but only700 nM for those of pituitary GH3 cells (83). Such high-affinity block led these authors to conclude that block ofT-type channels might contribute to felodipine’s antihy-pertensive properties. Amlodipine (Norvasc, Pfizer),which is currently the most widely prescribed dihydro-pyridine, is �12-fold selective for cardiac L-type channels(IC50 � 0.5 �M), blocking T-type channels with an appar-ent IC50 of 5.7 �M (323). Recombinant Cav3.2 channelswere slightly less sensitive (IC50 � 31 �M). Because am-lodipine serum concentrations are �14 nM (432), at least3% of the T-type channels will be blocked during therapy,and this fraction might be higher in depolarized cells.

The dihydropyridine structure-activity relationshipsof T- and L-type channels are different. This is mostclearly demonstrated by the stereoisomers of BAY K 8644;T-type channels are weakly blocked by micromolar con-centrations of both (237), while for L-type channels the(�)-isomer is an agonist (198). Stimulation of LVA cur-rents by BAY K 8644 is difficult to interpret because thiscompound shifts gating of L-type channels to more nega-tive potentials. In contrast, the stereoselectivity and po-tency of niguldipine were similar for both L- and T-typechannels; the racemic mixture blocked T-type currents inguinea pig atrial myocytes with an IC50 of 0.18 �M (341).Although the binding site for dihydropyridines is differentbetween L- and T-type channels, the mechanism of blockappears similar. Dihydropyridines appear to stabilize in-activated states of both channels, causing a negative shiftin the steady-state availability curve (31, 83, 221, 336). Inconclusion, dihydropyridines are selective for L-typechannels, and some analogs are capable of blocking T-type channels in the 1–10 �M range.

B. Antiepileptics

Generalized, or petit mal, absence epilepsies arecharacterized by periods of synchronized neuronal activ-ity, generating a 3- to 4-Hz spike-wave pattern on theelectroencephalogram. Considerable evidence suggeststhat spike-wave discharges are produced by synchronizedoscillations of cortical, reticular thalamic, and corticotha-lamic neurons (92, 136, 378). Spike-wave discharges ofcorticothalamic neurons are in part mediated by T-typechannels that produce LTS (93, 277, 377, 387). Ethosux-imide is a first-line drug in the treatment of absenceepilepsy (193). In vitro studies found that ethosuximidecan block T-type channels with little effect of HVA chan-nels in thalamic neurons (86). Because block occurred attherapeutically relevant concentrations (0.3 mM), it wasconcluded that T-type channel block might explain itsmechanism of action. Support for this hypothesis camefrom the following studies with related analogs: 1) T-typecurrent block was also observed at therapeutically rele-vant concentrations of methyl-phenylsuccinimide (MPS),the active metabolite of the antiepileptic drug methsux-imide, and 2) no block was observed using either theinactive analog succinimide or the convulsant analogtetra-methylsuccinimide (87). With one exception (215),most studies have failed to confirm ethosuximide block ofT-type currents at therapeutically relevant concentrations(403), leading to the suggestion that block of other ionicchannels may be more relevant (232).

Ethosuximide and MPS are capable of blocking allthree recombinant T-type Ca2� channels (137). As ob-served with mibefradil, block is enhanced up to 10-fold byholding the membrane at potentials that partially inacti-vate T-type channels. In addition, ethosuximide hasgreater affinity for open than rested channels, suggestingthat it is both a gating modifier and an open channelblocker. The effect of ethosuximide on other recombinantCa2� channels has not been rigorously tested, althoughblock of Cav2.3 (IC50 � 20 mM) has been reported (295).In particular, its affinity for partially inactivated HVAchannels has not been studied, so its selectivity is un-known. Studies with MPS suggest that it is only abouttwofold selective for LVA over HVA channels in thalamicneurons (87). The affinity (Ki) of ethosuximide for inac-tivated states was estimated to be 2 mM. Because thera-peutic plasma concentrations are close to 0.5 mM (57),only a fraction of channels is expected to be inhibitedduring therapy. This partial block may be relevant due to“pharmacological amplification” (297). This is due to therole of thalamic T-type channels in gating Na� channels,such that nominal block of the T-type channel is sufficientto block all-or-none action potentials (178).

T-type channels are also blocked by antiepilepticsthat are thought to work by blocking Na� channels. Per-haps the best studied is phenytoin, which clearly blocks

148 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 33: Canal Calcio T

both channels at therapeutic concentrations (413). Phe-nytoin appears to be selective for LVA over HVA currents(413). The mechanisms of block are also similar, withhigher affinity for inactivated states as evidenced by itsability to shift the h� curve (413). From this shift it can becalculated that phenytoin blocks inactivated channelswith a Ki of 7 �M. Studies using recombinant T-typechannels indicated that Cav3.2 was the most sensitive,although block was variable even in a clonal cell line,suggesting that additional factors may be important forblock (404). Zonisamide has also been reported to par-tially block T-type channels (60% max, IC50 �70 �M) withlittle effect on HVA currents (386).

Additional support for the hypothesis that an in-crease in T-type currents might underlie epilepsy comesfrom animal models. One extensively studied model is thegenetic absence epilepsy rats from Strasbourg (GAERS;Ref. 419), whose T-type currents were larger than thoserecorded from a seizure-free rat strain (409). Consistentwith the enhanced currents, enhanced expression ofCav3.2 was detected in the reticular thalamic nuclei, al-though the changes in current density (50%) were largerthan the changes in channel mRNA (20%; Ref. 394). Inaddition, a small increase in Cav3.1 expression in theventrobasal nucleus of thalamus was detected at the mes-sage level, but not with currents (151). Another interest-ing result of this study was that T-type channel mRNA wasvery similar, and in fact rose slightly, between juvenile (15days old) and adult (�40 days old) rats. Enhanced expres-sion of Cav3.2 mRNA was noted at both time points.Because the absence epilepsy phenotype of these animalsis not manifested until after 30 days, these data suggestthat maturation of the thalamocortical circuit plays amore important role in determining the onset than theintrinsic properties of these neurons (419, 431). Increasedexpression of T-type currents has also been found inhippocampal CA1 pyramidal neurons in three differentmodels: a kindling model of focal epilepsy (112), a pilo-carpine model of temporal lobe epilepsy (380), and an invitro model (301). These currents, as well as intrinsicbursting of these neurons, could be blocked by 100 �MNi2� (380), suggesting the involvement of Cav3.2 channels(230).

C. Anesthetics

Although the precise mechanism of action of generalanesthetics is unknown, many studies have demonstratedthat they are capable of modulating the activity of ionchannels, including ligand-gated GABA and glycine chan-nels. A newly described mechanism for inhalation anes-thetics is that they hyperpolarize neurons by potentiatingthe activity of leak K� currents carried by the two-poredomain channels such as TASK-1 (371). Isoflurane, halo-

thane, and nitrous oxide block DRG T-type currents attherapeutically relevant concentrations of anesthetics(401, 403). Isoflurane block of either DRG or recombinantCav3.1 occurs at concentrations (271–303 �M) that arelower than those achieved during anesthesia (404). Blockis not totally selective, as block of HVA currents by halo-thane occurs at similar millimolar concentrations (160,390). In contrast, nitrous oxide selectively blocked DRGLVA currents with no effect on HVA currents (401). Inaddition, nitrous oxide was found to be a selectiveblocker of Cav3.2 currents, with little effect on Cav3.1currents.

Octanol (20 �M), an aliphatic alcohol, was reportedto block LVA currents totally from inferior olivary neu-rons (245). Subsequent studies required much higher con-centrations for block (IC50 �200 �M; reviewed in Ref.158). Recombinant T-type channels can be totally blockedby octanol, with 50% inhibition occurring at 160 �M forCav3.1 and 219 �M for Cav3.2 (404). Studies on hippocam-pal CA1 pyramidal neurons indicated that octanol is non-selective, blocking T-, N-, and L-type channels to a similarextent. In conclusion, T-type currents are blocked at ther-apeutically relevant concentrations of some anesthetics,and this block may contribute to their anesthetic andanalgesic properties.

D. Antipsychotics

Antipsychotics are used in the treatment of psychoticdisorders such as schizophrenia, the main phase of manic-depressive illness, and other acute idiopathic psychoticillnesses (20). Most antipsychotics (neuroleptics) act byblocking D2 class dopaminergic receptors, although drugsof diphenylbutylpiperidine class can also block T-typechannels.

Among diphenylbutylpiperidines, penfluridol was themost effective at blocking T-type currents of human med-ullary thyroid cancer (TT) cells (108). Penfluridol blockedwith an IC50 of 224 nM, and similar results were obtainedwith adrenal zona fasciculata cells (109). It was also�10-fold more selective for LVA currents over HVA cur-rents as 500 nM penfluridol inhibited 82% of the LVAcurrent but only 20% of the HVA. Block of other channelshas not been studied in detail, although it should be notedthat these drugs are also potent blockers of K� channels(139) and bind with subnanomolar affinity to L-type chan-nels (208). Despite strikingly different structures, thiorid-azine, clozapine, and haloperidol have comparable activ-ity in TT cells, but are still much less active thanfluspiriline or penfluridol (108). Similar results have beenobtained using recombinant rat Cav3 channels, althoughthe apparent affinity was higher than observed with nativechannels (349). Cav3.1 channels were blocked with thefollowing order of potency (IC50 in parentheses, in nM):

T-TYPE CALCIUM CHANNELS 149

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 34: Canal Calcio T

pimozide (43) � penfluridol (110) � flunarizine (530) �haloperidol (1,163). Similar results were obtained withCav3.2 and Cav3.3, although flunarizine block of Cav3.2required sevenfold higher concentrations. These com-pounds all shifted the steady-state inactivation curve tomore negative potentials, indicating that they bind pref-erentially to the inactivated state. Use-dependent blockhas been noted previously in studies of native channels,suggesting open channels may also have a higher affinitythan rested channels (15, 108). Pimozide binding to D2

receptors occurs over the same concentration range(KD � 29 nM) as block of the recombinant Cav3 channels,suggesting that T-type channels may be blocked duringtherapy (349).

VI. CONCLUSIONS

A. Physiological Roles

The expression of T-type channels in diverse celltypes suggests that they play a role in various physiolog-ical functions. The voltage dependence of activation andinactivation provides clues to these roles and places con-straints on when these channels will be active. In neuronswhere the resting membrane potential is in the �90 to�70 mV range, T-type channels can play a secondarypacemaker role; an excitatory postsynaptic potential(EPSP) opens T-type channels and generates a LTS,which in turn activates Na�-dependent action potentialsand HVA Ca2� channels. Therefore, T-type channels playan important role in the genesis of burst firing. In neuronswhere the resting membrane potential is relatively depo-larized (greater than �70 mV), T-type channels are inac-tivated and an EPSP directly activates Na� channels. Dueto their fast recovery from inactivation, T-type channelscan produce a rebound burst in depolarized neurons fol-lowing an IPSP. Electrophysiological recordings indicatethat these channels are preferentially localized to den-drites, and hence play a role in signal amplification (97,256, 265, 328). Large T-type currents have been found inthalamic neurons, where they play an important role inoscillatory behavior. The thalamus acts as a gateway tothe cerebral cortex, and inappropriate oscillations ofthese circuits, or thalamocortical dysrhythmias, havebeen implicated in a wide range of neurological disorders(242). T-type currents also appear to play a role in olfac-tion, vision, and pain reception (201, 317, 402).

Calcium influx not only depolarizes the plasma mem-brane but also acts as a second messenger, leading to theactivation of a plethora of enzyme and channel activities.T-type channels can cause robust increases in intracellu-lar Ca2�, especially in proximal dendrites (294, 460). Cal-cium and voltage synergistically open Ca2�-activated K�

channels, which contribute to spike repolarization and

afterhyperpolarizations (40, 244, 416). In addition to tran-sient increases in intracellular Ca2�, T-type window cur-rents may play an important role in slower increases inbasal Ca2�. This property appears to be important forhormone secretion from adrenal cortex and pituitary,myoblast fusion (45), and possibly smooth muscle con-traction.

B. Future Directions

Cloning of the T-type channel �1-subunits has pro-vided tools with which to study the distribution of thesechannels. As expected, mRNA for these channels wasfound to be distributed throughout the central nervoussystem; however, it was very surprising to find expressionin organs such as liver and kidney. Future studies arerequired to discern their role in these tissues. Such stud-ies would be aided by the development of high-affinityantibodies. Antibodies would also be useful in studyingthe distribution of these channels within neurons and intheir purification. HVA channels are multisubunit com-plexes; therefore, it seems likely that T-type channels willalso have auxiliary subunits. These hypothetical subunitsmay regulate surface expression or might be involved inhormonal regulation of channel activity.

Because mutations in many ion channel genes havebeen linked to inherited diseases, it seems likely thatmutations in T-type channel genes would lead to a phe-notype. Mutations that lead to enhanced expression ofchannels, or that reduce inactivation, might tip the bal-ance to overexcitability. Enhanced expression of T-typechannels have been observed in animal models of epi-lepsy (409), neuronal injury (79), cardiac hypertrophy(308), and heart failure (362). The opposite approach isalso being tested by creating transgenic mice that lackexpression of Cav3 genes. Studies with the Cav3.1 knock-out mouse have established the central role this isoformplays in generating burst firing and thalamic spike-wavedischarges (206). What will be the phenotype of otherCav3 knock-out mice?

Somewhat surprisingly the transcriptional regulationof Ca2� channel subunits has received scant attention.Preliminary studies indicate that the Cav3.1 promoter ismethylated in some tumors, which presumably leads todecreased expression (408). It has been noted that theCav3.2 gene contains a region with high homology to theconsensus sequence of neural restrictive silencer ele-ments, providing another avenue for future research(342). Transcriptional regulation also appears to be im-portant for hormonal upregulation of T channels (22, 447).Studies on gene regulation may also provide insights intothe control of T-channel expression during the cell cycle(220).

The pharmaceutical industry is using combinatorial

150 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 35: Canal Calcio T

libraries and high-throughput screening to search for newdrugs, and hopefully one of these will be a highly selectiveT-type channel blocker. Such a blocker might be useful inthe control of blood pressure, as evidenced by the utilityof mibefradil. If the drug penetrates the central nervoussystem, it might be useful in a variety of disorders thatinvolve thalamocortical dysrhythmias, such as epilepsyand neuropathic pain. Clearly, a highly selective blockerwould also be extremely useful in research and wouldlead to a greater understanding of the physiological rolesof voltage-gated Ca2� channels.

I thank my current and former fellows Drs. Jung-Ha Lee,Juan Carlos Gomora, and Janet Murbartian for help with record-ings of recombinant channels and proofreading. I thank JenniferDeLisle for countless trips to the library to photocopy articles. Ithank Drs. John Huguenard, David McCormick, Thierry Bal,Hee-Sup Shin, Andy Randall, and Chris Benham for permissionto use figures. I also acknowledge the valuable critiques andsuggestions provided by my collaborator Dr. Paula Barrett.

This work was supported in part by National Institutes ofHealth Grants HL-58728 and NS-38691 and an Established Inves-tigator Award from the American Heart Association.

Address for reprint requests and other correspondence: E.Perez-Reyes, Associate Professor of Pharmacology, Univ. ofVirginia, 1300 Jefferson Park Ave., Charlottesville, VA 22908-0735 (E-mail: [email protected]).

REFERENCES

1. ABDULLA FA AND SMITH PA. Nociceptin inhibits T-type Ca2� channelcurrent in rat sensory neurons by a G-protein-independent mech-anism. J Neurosci 17: 8721–8728, 1997.

2. AIZENMAN CD AND LINDEN DJ. Regulation of the rebound depolar-ization and spontaneous firing patterns of deep nuclear neurons inslices of rat cerebellum. J Neurophysiol 82: 1697–1709, 1999.

3. AKAIKE N, KANAIDE H, KUGA T, NAKAMURA M, SADOSHIMA J, AND

TOMOIKE H. Low voltage-activated calcium current in rat aortasmooth muscle cells in primary culture. J Physiol 416: 141–160,1989.

4. AKAIKE N, KOSTYUK PG, AND OSIPCHUK YV. Dihydropyridine-sensitivelow-threshold calcium channels in isolated rat hypothalamic neu-rones. J Physiol 412: 181–195, 1989.

5. ALLEN TG AND BROWN DA. M2 muscarinic receptor-mediated inhibi-tion of the Ca2� current in rat magnocellular cholinergic basalforebrain neurones. J Physiol 466: 173–189, 1993.

6. ALLEN TG, SIM JA, AND BROWN DA. The whole-cell calcium current inacutely dissociated magnocellular cholinergic basal forebrain neu-rones of the rat. J Physiol 460: 91–116, 1993.

7. ALONSO A AND LLINAS RR. Electrophysiology of the mammillarycomplex in vitro. II. Medial mammillary neurons. J Neurophysiol

68: 1321–1331, 1992.8. ALVAREZ JL, RUBIO LS, AND VASSORT G. Facilitation of T-type calcium

current in bullfrog atrial cells: voltage-dependent relief of a Gprotein inhibitory tone. J Physiol 491: 321–334, 1996.

9. ALVAREZ JL AND VASSORT G. Properties of the low threshold Cacurrent in single frog atrial cardiomyocytes. A comparison with thehigh threshold Ca current. J Gen Physiol 100: 519–545, 1992.

10. ALVAREZ DE TOLEDO G AND LOPEZ-BARNEO J. Ionic basis of the differ-ential neuronal activity of guinea-pig septal nucleus studied invitro. J Physiol 396: 399–415, 1988.

11. ANDREASEN D, JENSEN BL, HANSEN PB, KWON TH, NIELSEN S, AND

SKØTT O. The �1G-subunit of a voltage-dependent Ca2� channel islocalized in rat distal nephron and collecting duct. Am J Physiol

Renal Physiol 279: F997–F1005, 2000.

12. ARMSTRONG CM AND MATTESON DR. Two distinct populations ofcalcium channels in a clonal line of pituitary cells. Science 227:65–67, 1985.

13. ARNOULT C, CARDULLO RA, LEMOS JR, AND FLORMAN HM. Activation ofmouse sperm T-type Ca2� channels by adhesion of the egg zonapellucida. Proc Natl Acad Sci USA 93: 13004–13009, 1996.

14. ARNOULT C, LEMONS JR, AND FLORMAN HM. Voltage-dependent mod-ulation of T-type calcium channels by protein tyrosine phosphory-lation. EMBO J 16: 1593–1599, 1997.

15. ARNOULT C, VILLAZ M, AND FLORMAN HM. Pharmacological propertiesof the T-type Ca2� current of mouse spermatogenic cells. Mol

Pharmacol 53: 1104–1111, 1998.16. ARTALEJO CR, ADAMS ME, AND FOX AP. Three types of Ca2� channel

trigger secretion with different efficacies in chromaffin cells. Na-

ture 367: 72–76, 1994.17. ASHCROFT FM, KELLY RP, AND SMITH PA. Two types of Ca channel in

rat pancreatic beta-cells. Pflugers Arch 415: 504–506, 1990.18. ASSANDRI R, EGGER M, GASSMANN M, NIGGLI E, BAUER C, FORSTER I,

AND GORLACH A. Erythropoietin modulates intracellular calcium in ahuman neuroblastoma cell line. J Physiol 516: 343–352, 1999.

19. AVERY RB AND JOHNSTON D. Multiple channel types contribute to thelow-voltage-activated calcium current in hippocampal CA3 pyrami-dal neurons. J Neurosci 16: 5567–5582, 1996.

20. BALDESSARINI RJ. Drugs and the treatment of psychiatric disorders.In: The Pharmacological Basis of Therapeutics (9th ed.), edited byHardman JG, Limbird LE, Molinoff PB, Ruddon RW and Gilman AG.New York: McGraw-Hill, 1996, p. 399–430.

21. BALKE CW, ROSE WC, MARBAN E, AND WIER WG. Macroscopic andunitary properties of physiological ion flux through T-type Ca2�

channels in guinea-pig heart cells. J Physiol 456: 247–265, 1992.22. BARBARA JG AND TAKEDA K. Voltage-dependent currents and modu-

lation of calcium channel expression in zona fasciculata cells fromrat adrenal gland. J Physiol 488: 609–622, 1995.

23. BARCLAY J, BALAGUERO N, MIONE M, ACKERMAN SL, LETTS VA, BROD-BECK J, CANTI C, MEIR A, PAGE KM, KUSUMI K, PEREZ-REYES E, LANDER

ES, FRANKEL WN, GARDINER RM, DOLPHIN AC, AND REES M. Duckymouse phenotype of epilepsy and ataxia is associated with muta-tions in the Cacna2d2 gene and decreased calcium channel currentin cerebellar Purkinje cells. J Neurosci 21: 6095–6104, 2001.

24. BARNES S AND HAYNES LW. Low-voltage-activated calcium channelsin human retinoblastoma cells. Brain Res 598: 19–22, 1992.

25. BARNETT DW, PRESSEL DM, AND MISLER S. Voltage-dependent Na�

and Ca2� currents in human pancreatic islet �-cells: evidence forroles in the generation of action potentials and insulin secretion.Pflugers Arch 431: 272–282, 1995.

26. BARRETT PQ, ISALES CM, BOLLAG WB, AND MCCARTHY RT. Ca2� chan-nels and aldosterone secretion: modulation by K� and atrial natri-uretic peptide. Am J Physiol Renal Fluid Electrolyte Physiol 261:F706–F719, 1991.

27. BARRETT PQ, LU HK, COLBRAN R, CZERNIK A, AND PANCRAZIO JJ.Stimulation of unitary T-type Ca2� channel currents by calmodulin-dependent protein kinase II. Am J Physiol Cell Physiol 279: C1694–C1703, 2000.

28. BAYLIS C, QIU C, AND ENGELS K. Comparison of L-type and mixed L-and T-type calcium channel blockers on kidney injury caused bydeoxycorticosterone-salt hypertension in rats. Am J Kidney Dis

38: 1292–1297, 2001.29. BEAM KG AND KNUDSON CM. Effect of postnatal development on

calcium currents and slow charge movement in mammalian skel-etal muscle. J Gen Physiol 91: 799–815, 1988.

30. BEAM KG, KNUDSON CM, AND POWELL JA. A lethal mutation in miceeliminates the slow calcium current in skeletal muscle cells. Na-

ture 320: 168–170, 1986.31. BEAN BP. Nitrendipine block of cardiac calcium channels: high-

affinity binding to the inactivated state. Proc Natl Acad Sci USA 81:6388–6392, 1984.

32. BEAN BP. Two kinds of calcium channels in canine atrial cells.Differences in kinetics, selectivity, and pharmacology. J Gen

Physiol 86: 1–30, 1985.33. BEAN BP AND MCDONOUGH SI. Two for T. Neuron 20: 825–828, 1998.34. BENHAM CD, HESS P, AND TSIEN RW. Two types of calcium channels

in single smooth muscle cells from rabbit ear artery studied with

T-TYPE CALCIUM CHANNELS 151

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 36: Canal Calcio T

whole-cell and single-channel recordings. Circ Res 61: I10–I16,1987.

35. BENHAM CD AND TSIEN RW. Noradrenaline modulation of calciumchannels in single smooth muscle cells from rabbit ear artery.J Physiol 404: 767–784, 1988.

36. BERGER AJ AND TAKAHASHI T. Serotonin enhances a low-voltage-activated calcium current in rat spinal motoneurons. J Neurosci 10:1922–1928, 1990.

37. BERS DM AND PEREZ-REYES E. Ca channels in cardiac myocytes:structure and function in Ca influx and intracellular Ca release.Cardiovasc Res 42: 339–360, 1999.

38. BERTHIER C, MONTEIL A, LORY P, AND STRUBE C. �1H mRNA in singleskeletal muscle fibres accounts for T-type calcium current tran-sient expression during fetal development in mice. J Physiol 539:681–691, 2002.

39. BEUCKELMANN DJ, NABAUER M, AND ERDMANN E. Characteristics ofcalcium-current in isolated human ventricular myocytes from pa-tients with terminal heart failure. J Mol Cell Cardiol 23: 929–937,1991.

40. BEURRIER C, CONGAR P, BIOULAC B, AND HAMMOND C. Subthalamicnucleus neurons switch from single-spike activity to burst-firingmode. J Neurosci 19: 599–609, 1999.

41. BEZPROZVANNY I AND TSIEN RW. Voltage-dependent blockade of di-verse types of voltage-gated Ca2� channels expressed in Xenopus

oocytes by the Ca2� channel antagonist mibefradil (Ro 40–5967).Mol Pharmacol 48: 540–549, 1995.

42. BHATTACHARJEE A, WHITEHURST RM JR, ZHANG M, WANG L, AND LI M.T-type calcium channels facilitate insulin secretion by enhancinggeneral excitability in the insulin-secreting beta-cell line, INS-1.Endocrinology 138: 3735–3740, 1997.

43. BIAGI BA AND ENYEART JJ. Multiple calcium currents in a thyroidC-cell line: biophysical properties and pharmacology. Am J Physiol

Cell Physiol 260: C1253–C1263, 1991.44. BIAGI BA, MLINAR B, AND ENYEART JJ. Membrane currents in a

calcitonin-secreting human C cell line. Am J Physiol Cell Physiol

263: C986–C994, 1992.45. BIJLENGA P, LIU JH, ESPINOS E, HAENGGELI CA, FISCHER-LOUGHEED J,

BADER CR, AND BERNHEIM L. T-type �1H Ca2� channels are involvedin Ca2� signaling during terminal differentiation (fusion) of humanmyoblasts. Proc Natl Acad Sci USA 97: 7627–7632, 2000.

46. BKAILY G, SCULPTOREANU A, JACQUES D, AND JASMIN G. Increases ofT-type Ca2� current in heart cells of the cardiomyopathic hamster.Mol Cell Biochem 176: 199–204, 1997.

47. BLANKENFELD G, VERKHRATSKY AN, AND KETTENMANN H. Ca2� channelexpression in the oligodendrocyte lineage. Eur J Neurosci 4: 1035–1048, 1992.

48. BOGDANOV KY, ZIMAN BD, SPURGEON HA, AND LAKATTA EG. L- andT-type calcium currents differ in finch and rat ventricular cardio-myocytes. J Mol Cell Cardiol 27: 2581–2593, 1995.

49. BOHN G, MOOSMANG S, CONRAD H, LUDWIG A, HOFMANN F, AND KLUG-BAUER N. Expression of T- and L-type calcium channel mRNA inmurine sinoatrial node. FEBS Lett 481: 73–76, 2000.

50. BOOTMAN MD, COLLINS TJ, PEPPIATT CM, PROTHERO LS, MACKENZIE L,DE SMET P, TRAVERS M, TOVEY SC, SEO JT, BERRIDGE MJ, CICCOLINI F,AND LIPP P. Calcium signalling—an overview. Semin Cell Dev Biol

12: 3–10, 2001.51. BORGLAND SL, CONNOR M, AND CHRISTIE MJ. Nociceptin inhibits

calcium channel currents in a subpopulation of small nociceptivetrigeminal ganglion neurons in mouse. J Physiol 536: 35–47, 2001.

52. BOSSU JL, FAGNI L, AND FELTZ A. Voltage-activated calcium channelsin rat Purkinje cells maintained in culture. Pflugers Arch 414:92–94, 1989.

53. BOSSU JL AND FELTZ A. Inactivation of the low-threshold transientcalcium current in rat sensory neurones: evidence for a dual pro-cess. J Physiol 376: 341–357, 1986.

54. BOSSU JL, FELTZ A, AND THOMANN JM. Depolarization elicits twodistinct calcium currents in vertebrate sensory neurones. Pflugers

Arch 403: 360–368, 1985.55. BOURINET E, SOONG TW, SUTTON K, SLAYMAKER S, MATHEWS E, MON-

TEIL A, ZAMPONI GW, NARGEOT J, AND SNUTCH TP. Splicing of alpha 1Asubunit gene generates phenotypic variants of P- and Q-type cal-cium channels. Nat Neurosci 2: 407–415, 1999.

56. BOURINET E, ZAMPONI GW, STEA A, SOONG TW, LEWIS BA, JONES LP,

YUE DT, AND SNUTCH TP. The �1E calcium channel exhibits perme-ation properties similar to low-voltage-activated calcium channels.J Neurosci 16: 4983–4993, 1996.

57. BROWNE TR, DREIFUSS FE, DYKEN PR, GOODE DJ, PENRY JK, PORTER

RJ, WHITE BG, AND WHITE PT. Ethosuximide in the treatment ofabsence (petit mal) seizures. Neurology 25: 515–524, 1975.

58. BUISSON B, BOTTARI SP, DE GASPARO M, GALLO-PAYET N, AND PAYET

MD. The angiotensin AT2 receptor modulates T-type calcium cur-rent in non-differentiated NG108–15 cells. FEBS Lett 309: 161–164,1992.

59. BUISSON B, LAFLAMME L, BOTTARI SP, DE GASPARO M, GALLO-PAYET N,AND PAYET MD. A G protein is involved in the angiotensin AT2

receptor inhibition of the T-type calcium current in non-differenti-ated NG108–15 cells. J Biol Chem 270: 1670–1674, 1995.

60. BURLHIS TM AND AGHAJANIAN GK. Pacemaker potentials of seroto-nergic dorsal raphe neurons: contribution of a low-threshold Ca2�

conductance. Synapse 1: 582–588, 1987.61. CAMPBELL DC, RASMUSSON RL, AND STRAUSS HC. Ionic current mech-

anisms generating vertebrate primary cardiac pacemaker activityat the single cell level: an integrative view. Annu Rev Physiol 54:279–302, 1992.

62. CARBONE E AND LUX HD. A low voltage-activated calcium conduc-tance in embryonic chick sensory neurons. Biophys J 46: 413–418,1984.

63. CARBONE E AND LUX HD. A low voltage-activated, fully inactivatingCa channel in vertebrate sensory neurones. Nature 310: 501–502,1984.

64. CARBONE E AND LUX HD. Single low-voltage-activated calcium chan-nels in chick and rat sensory neurones. J Physiol 386: 571–601,1987.

65. CARBONE E, SHER E, AND CLEMENTI F. Ca currents in human neuro-blastoma IMR32 cells: kinetics, permeability, and pharmacology.Pflugers Arch 416: 170–179, 1990.

66. CHAPMAN ML, VANDONGEN HM, AND VANDONGEN AM. Activation-dependent subconductance levels in the drk1 K channel suggest asubunit basis for ion permeation and gating. Biophys J 72: 708–719,1997.

67. CHAUDHARI N. A single nucleotide deletion in the skeletal muscle-specific calcium channel transcript of muscular dysgenesis (mdg)mice. J Biol Chem 267: 25636–25639, 1992.

68. CHEMIN J, MONTEIL A, BOURINET E, NARGEOT J, AND LORY P. Alterna-tively spliced �1G (Cav3.1) intracellular loops promote specificT-type Ca2� channel gating properties. Biophys J 80: 1238–1250,2001.

69. CHEMIN J, MONTEIL A, BRIQUAIRE C, RICHARD S, PEREZ-REYES E, NAR-GEOT J, AND LORY P. Overexpression of T-type calcium channels inHEK-293 cells increases intracellular calcium without affectingcellular proliferation. FEBS Lett 478: 166–172, 2000.

70. CHEMIN J, MONTEIL A, PEREZ-REYES E, BOURINET E, NARGEOT J, AND

LORY P. Specific contribution of human T-type calcium channelisotypes (�1G, �1H and �1I) to neuronal excitability. J Physiol 540:3–14, 2002.

71. CHEMIN J, MONTEIL A, PEREZ-REYES E, NARGEOT J, AND LORY P. Directinhibition of T-type calcium channels by the endogenous cannabi-noid anandamide. EMBO J 20: 7033–7040, 2001.

72. CHEN C, ZHANG J, VINCENT JD, AND ISRAEL JM. Two types of voltage-dependent calcium current in rat somatotrophs are reduced bysomatostatin. J Physiol 425: 29–42, 1990.

73. CHEN CF AND HESS P. Mechanism of gating of T-type calciumchannels. J Gen Physiol 96: 603–630, 1990.

74. CHEN L, CHETKOVICH DM, PETRALIA RS, SWEENEY NT, KAWASAKI Y,WENTHOLD RJ, BREDT DS, AND NICOLL RA. Stargazin regulates syn-aptic targeting of AMPA receptors by two distinct mechanisms.Nature 408: 936–943, 2000.

75. CHEN XH, BEZPROZVANNY I, AND TSIEN RW. Molecular basis of protonblock of L-type Ca2� channels. J Gen Physiol 108: 363–374, 1996.

76. CHEN XL, BAYLISS DA, FERN RJ, AND BARRETT PQ. A role for T-typeCa2� channels in the synergistic control of aldosterone productionby ANG II and K�. Am J Physiol Renal Physiol 276: F674–F683,1999.

77. CHRISTIE BR, ELIOT LS, ITO K, MIYAKAWA H, AND JOHNSTON D. DifferentCa2� channels in soma and dendrites of hippocampal pyramidal

152 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 37: Canal Calcio T

neurons mediate spike-induced Ca2� influx. J Neurophysiol 73:2553–2557, 1995.

78. CHU PJ, ROBERTSON HM, AND BEST PM. Calcium channel � subunitsprovide insights into the evolution of this gene family. Gene 280:37–48, 2001.

79. CHUNG JM, HUGUENARD JR, AND PRINCE DA. Transient enhancementof low-threshold calcium current in thalamic relay neurons aftercorticectomy. J Neurophysiol 70: 20–27, 1993.

80. CHURCHILL D AND MACVICAR BA. Biophysical and pharmacologicalcharacterization of voltage-dependent Ca2� channels in neuronsisolated from rat nucleus accumbens. J Neurophysiol 79: 635–647,1998.

81. COGNARD C, LAZDUNSKI M, AND ROMEY G. Different types of Ca2�

channels in mammalian skeletal muscle cells in culture. Proc Natl

Acad Sci USA 83: 517–521, 1986.82. COHEN CJ, MCCARTHY RT, BARRETT PQ, AND RASMUSSEN H. Ca chan-

nels in adrenal glomerulosa cells: K� and angiotensin II increaseT-type Ca channel current. Proc Natl Acad Sci USA 85: 2412–2416,1988.

83. COHEN CJ, SPIRES S, AND VAN SKIVER D. Block of T-type Ca channelsin guinea pig atrial cells by antiarrhythmic agents and Ca channelantagonists. J Gen Physiol 100: 703–728, 1992.

84. COTA G. Calcium channel currents in pars intermedia cells of the ratpituitary gland. Kinetic properties and washout during intracellulardialysis. J Gen Physiol 88: 83–105, 1986.

85. COULTER DA, HUGUENARD JR, AND PRINCE DA. Calcium currents in ratthalamocortical relay neurones: kinetic properties of the transient,low-threshold current. J Physiol 414: 587–604, 1989.

86. COULTER DA, HUGUENARD JR, AND PRINCE DA. Characterization ofethosuximide reduction of low-threshold calcium current in tha-lamic neurons. Ann Neurol 25: 582–593, 1989.

87. COULTER DA, HUGUENARD JR, AND PRINCE DA. Differential effects ofpetit mal anticonvulsants and convulsants on thalamic neurones:calcium current reduction. Br J Pharmacol 100: 800–806, 1990.

88. CRAIG PJ, BEATTIE RE, FOLLY EA, BANERJEE MD, REEVES MB, PRIEST-LEY JV, CARNEY SL, SHER E, PEREZ-REYES E, AND VOLSEN SG. Distri-bution of the voltage-dependent calcium channel �1G subunitmRNA and protein throughout the mature rat brain. Eur J Neurosci

11: 2949–2964, 1999.89. CRIBBS LL. Vascular smooth muscle calcium channels: could “T” be

a target? Circ Res 89: 560–562, 2001.90. CRIBBS LL, LEE JH, YANG J, SATIN J, ZHANG Y, DAUD A, BARCLAY J,

WILLIAMSON MP, FOX M, REES M, AND PEREZ-REYES E. Cloning andcharacterization of �1H from human heart, a member of the T-typeCa2� channel gene family. Circ Res 83: 103–109, 1998.

91. CRIBBS LL, MARTIN BL, SCHRODER EA, KELLER BB, DELISLE BP, AND

SATIN J. Identification of the T-type calcium channel Cav3.1d indeveloping mouse heart. Circ Res 88: 403–407, 2001.

92. CRUNELLI V AND LERESCHE N. Childhood absence epilepsy: genes,channels, neurons and networks. Nat Rev Neurosci 3: 371–382,2002.

93. CRUNELLI V, LIGHTOWLER S, AND POLLARD CE. A T-type Ca2� currentunderlies low-threshold Ca2� potentials in cells of the cat and ratlateral geniculate nucleus. J Physiol 413: 543–561, 1989.

94. DARSZON A, LABARCA P, NISHIGAKI T, AND ESPINOSA F. Ion channels insperm physiology. Physiol Rev 79: 481–510, 1999.

95. DELISLE BP AND SATIN J. pH modification of human T-type calciumchannel gating. Biophys J 78: 1895–1905, 2000.

96. DESCHENES M, PARADIS M, ROY JP, AND STERIADE M. Electrophysiol-ogy of neurons of lateral thalamic nuclei in cat: resting propertiesand burst discharges. J Neurophysiol 51: 1196–1219, 1984.

97. DESTEXHE A, NEUBIG M, ULRICH D, AND HUGUENARD J. Dendriticlow-threshold calcium currents in thalamic relay cells. J Neurosci

18: 3574–3588, 1998.98. DOERR T, DENGER R, AND TRAUTWEIN W. Calcium currents in single

SA nodal cells of the rabbit heart studied with action potentialclamp. Pflugers Arch 413: 599–603, 1989.

99. DOLPHIN AC. Properties and modulation of T-type currents in dorsalroot ganglia and ND7–23 cells: comparison with �1E currentsexpressed in COS-7 cells. In: Low Voltage-Activated T-type Cal-

cium Channels, edited by Tsien RW, Clozel J-P, and NargeotJ. Chester, UK: Adis International, 1998, p. 269–278.

100. DOLPHIN AC, WYATT CN, RICHARDS J, BEATTIE RE, CRAIG P, LEE JH,

CRIBBS LL, VOLSEN SG, AND PEREZ-REYES E. The effect of �2� andother accessory subunits on expression and properties of the cal-cium channel �1G. J Physiol 519: 35–45, 1999.

101. DOUGHTY JM, BARNES-DAVIES M, RUSZNAK Z, HARASZTOSI C, AND FOR-SYTHE ID. Contrasting Ca2� channel subtypes at cell bodies andsynaptic terminals of rat anterioventral cochlear bushy neurones.J Physiol 512: 365–376, 1998.

102. DOYLE DA, CABRAL JM, PFUETZNER RA, KUO A, GULBIS JM, COHEN SL,CHAIT BT, AND MACKINNON R. The structure of the potassium chan-nel: molecular basis of K� conduction and selectivity. Science 280:69–77, 1998.

103. DROLET P, BILODEAU L, CHORVATOVA A, LAFLAMME L, GALLO-PAYET N,AND PAYET MD. Inhibition of the T-type Ca2� current by the dopa-mine D1 receptor in rat adrenal glomerulosa cells: requirement ofthe combined action of the G �� protein subunit and cyclic aden-osine 3,5-monophosphate. Mol Endocrinol 11: 503–514, 1997.

104. DROOGMANS G AND NILIUS B. Kinetic properties of the cardiac T-typecalcium channel in the guinea-pig. J Physiol 419: 627–650, 1989.

105. DUNHAM I, HUNT AR, COLLINS JE, BRUSKIEWICH R, BEARE DM, CLAMP M,SMINK LJ, AINSCOUGH R, ALMEIDA JP, BABBAGE A, BAGGULEY C, BAILEY J,BARLOW K, BATES KN, BEASLEY O, BIRD CP, BLAKEY S, BRIDGEMAN AM,BUCK D, BURGESS J, BURRILL WD, BURTON J, CARDER C, CARTER NP, CHEN

Y, CLARK G, CLEGG SM, COBLEY V, COLE CG, COLLIER RE, CONNOR RE,CONROY D, CORBY N, COVILLE GJ, COX AV, DAVIS J, DAWSON E, DHAMI PD,DOCKREE C, DODSWORTH SJ, DURBIN RM, ELLINGTON A, EVANS KL, FEY

JM, FLEMING K, FRENCH L, GARNER AA, GILBERT JGR, GOWAND ME,GRAFHAM D, GRIFFITHS MN, HALL C, HALL R, HALL-TAMLYN G, HEATHCOTT

RW, HO S, HOLMES S, HUNT SE, JONES MC, KERSHAW J, KIMBERLEY A,KING A, LAIRD GK, LANGFORD CF, LEVERSHA MA, LLOYD C, LLOYD DM,MARTYN ID, MASHREGHI-MOHAMMADI M, MATTHEWS L, MCCANN OT, MC-CLAY J, MCLAREN S, MCMURRAY AA, MILNE SA, MORTIMORE BJ, ODELL

CN, PAVITT R, PEARCE AV, PEARSON D, PHILLIMORE BJ, PHILLIPS SH,PLUMB RW, RAMSAY H, RAMSEY Y, ROGERS L, ROSS MT, SCOTT CE, SEHRA

HK, SKUCE CD, SMALLEY S, SMITH ML, SODERLUND C, SPRAGON L, STEWARD

CA, SULSTON JE, SWANN RM, VAUDIN M, WALL M, WALLIS JM, WHITELEY

MN, WILLEY D, WILLIAMS L, WILLIAMS S, WILLIAMSON H, WILMER TE,WILMING L, WRIGHT CL, HUBBARD T, BENTLEY DR, BECK S, ROGERS J,SHIMIZU N, MINOSHIMA S, KAWASAKI K, SASAKI T, ASAKAWA S, KUDOH J,SHINTANI A, SHIBUYA K, YOSHIZAKI Y, AOKI N, MITSUYAMA S, ROE BA, CHEN

F, CHU L, CRABTREE J, DESCHAMPS S, DO A, DO T, DORMAN A, FANG F, FU

Y, HU P, HUA A, KENTON S, LAI H, LAO HI, LEWIS J, LEWIS S, LIN S-P, LOH

P, MALAJ E, NGUYEN T, PAN H, PHAN S, QI S, QIAN Y, RAY L, REN Q,SHAULL S, SLOAN D, SONG L, WANG Q, WANG Y, WANG Z, WHITE J,WILLINGHAM D, WU H, YAO Z, ZHAN M, ZHANG G, CHISSOE S, MURRAY J,MILLER N, MINX, P, FULTON R, JOHNSON D, BEMIS G, BENTLEY D, BRAD-SHAW H, BOURNE S, CORDES M, DU Z, FULTON L, GOELA D, GRAVES T,HAWKINS J, HINDS K, KEMP K, LATREILLE P, LAYMAN D, OZERSKY P, ROHLF-ING T, SCHEET P, WALKER C, WAMSLEY A, WOHLDMANN P, PEPIN K, NELSON

J, KORF I, BEDELL JA, HILLIER L, MARDIS E, WATERSTON R, WILSON R,EMANUEL BS, SHAIKH T, KURAHASHI H, SAITTA S, BUDARF ML, MCDERMID

HE, JOHNSON A, WONG ACC, MORROW BE, EDELMANN L, KIM UJ, SHIZUYA

H, SIMON MI, DUMANSKI JP, PEYRARD M, KEDRA D, SEROUSSI E, FRANSSON

I, TAPIA I, BRUDER CE, AND O’BRIEN KP. The DNA sequence of humanchromosome 22. Nature 402: 489–495, 1999.

106. ELLER P, BERJUKOV S, WANNER S, HUBER I, HERING S, KNAUS HG, TOTH

G, KIMBALL SD, AND STRIESSNIG J. High affinity interaction of mibe-fradil with voltage-gated calcium and sodium channels. Br J Phar-

macol 130: 669–677, 2000.107. ELSEN FP AND RAMIREZ JM. Calcium currents of rhythmic neurons

recorded in the isolated respiratory network of neonatal mice.J Neurosci 18: 10652–10662, 1998.

108. ENYEART JJ, BIAGI BA, AND MLINAR B. Preferential block of T-typecalcium channels by neuroleptics in neural crest-derived rat andhuman C cell lines. Mol Pharmacol 42: 364–372, 1992.

109. ENYEART JJ, MLINAR B, AND ENYEART JA. T-type Ca2� channels arerequired for adrenocorticotropin-stimulated cortisol production bybovine adrenal zona fasciculata cells. Mol Endocrinol 7: 1031–1040,1993.

110. ERICKSON KR, RONNEKLEIV OK, AND KELLY MJ. Role of a T-typecalcium current in supporting a depolarizing potential, dampedoscillations, and phasic firing in vasopressinergic guinea pig su-praoptic neurons. Neuroendocrinology 57: 789–800, 1993.

111. ERICKSON MG, ALSEIKHAN BA, PETERSON BZ, AND YUE DT. Preasso-

T-TYPE CALCIUM CHANNELS 153

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 38: Canal Calcio T

ciation of calmodulin with voltage-gated Ca2� channels revealed byFRET in single living cells. Neuron 31: 973–985, 2001.

112. FAAS GC, VREUGDENHIL M, AND WADMAN WJ. Calcium currents inpyramidal CA1 neurons in vitro after kindling epileptogenesis inthe hippocampus of the rat. Neuroscience 75: 57–67, 1996.

113. FAREH S, BENARDEAU A, AND NATTEL S. Differential efficacy of L- andT-type calcium channel blockers in preventing tachycardia-inducedatrial remodeling in dogs. Cardiovasc Res 49: 762–770, 2001.

114. FARES N, GOMEZ JP, AND POTREAU D. T-type calcium current isexpressed in dedifferentiated adult rat ventricular cells in primaryculture. Comptes Rendus Acad Sci 319: 569–576, 1996.

115. FEDULOVA SA, KOSTYUK PG, AND VESELOVSKY NS. Two types of cal-cium channels in the somatic membrane of new-born rat dorsalroot ganglion neurones. J Physiol 359: 431–446, 1985.

116. FERMINI B AND NATHAN RD. Removal of sialic acid alters both T- andL-type calcium currents in cardiac myocytes. Am J Physiol Heart

Circ Physiol 260: H735–H743, 1991.117. FISHER R AND JOHNSTON D. Differential modulation of single voltage-

gated calcium channels by cholinergic and adrenergic agonists inadult hippocampal neurons. J Neurophysiol 64: 1291–1302, 1990.

118. FISHER RE, GRAY R, AND JOHNSTON D. Properties and distribution ofsingle voltage-gated calcium channels in adult hippocampal neu-rons. J Neurophysiol 64: 91–104, 1990.

119. FISHER TE AND BOURQUE CW. Voltage-gated calcium currents in themagnocellular neurosecretory cells of the rat supraoptic nucleus.J Physiol 486: 571–580, 1995.

120. FISHMAN MC AND SPECTOR I. Potassium current suppression byquinidine reveals additional calcium currents in neuroblastomacells. Proc Natl Acad Sci USA 78: 5245–5249, 1981.

121. FLORMAN HM, CORRON ME, KIM TD, AND BABCOCK DF. Activation ofvoltage-dependent calcium channels of mammalian sperm is re-quired for zona pellucida-induced acrosomal exocytosis. Dev Biol

152: 304–314, 1992.122. FOEHRING RC, LORENZON NM, HERRON P, AND WILSON CJ. Correlation

of physiologically and morphologically identified neuronal types inhuman association cortex in vitro. J Neurophysiol 66: 1825–1837,1991.

123. FOX AP AND KRASNE S. Two calcium currents in Neanthes arena-

ceodentatus egg cell membranes. J Physiol 356: 491–505, 1984.124. FOX AP, NOWYCKY MC, AND TSIEN RW. Kinetic and pharmacological

properties distinguishing three types of calcium currents in chicksensory neurones. J Physiol 394: 149–172, 1987.

125. FOX AP, NOWYCKY MC, AND TSIEN RW. Single-channel recordings ofthree types of calcium channels in chick sensory neurones.J Physiol 394: 173–200, 1987.

126. FRASER DD AND MACVICAR BA. Low-threshold transient calciumcurrent in rat hippocampal lacunosum-moleculare interneurons:kinetics and modulation by neurotransmitters. J Neurosci 11:2812–2820, 1991.

127. FRAZIER CJ, SERRANO JR, GEORGE EG, YU X, VISWANATHAN A, PEREZ-REYES E, AND JONES SW. Gating kinetics of the �1I T-type calciumchannel. J Gen Physiol 118: 457–470, 2001.

128. FUKUSHIMA Y AND HAGIWARA S. Currents carried by monovalentcations through calcium channels in mouse neoplastic B lympho-cytes. J Physiol 358: 255–284, 1985.

129. FURUKAWA T, ITO H, NITTA J, TSUJINO M, ADACHI S, HIROE M, MARUMO

F, SAWANOBORI T, AND HIRAOKA M. Endothelin-1 enhances calciumentry through T-type calcium channels in cultured neonatal ratventricular myocytes. Circ Res 71: 1242–1253, 1992.

130. GANITKEVICH V AND ISENBERG G. Contribution of two types of calciumchannels to membrane conductance of single myocytes fromguinea-pig coronary artery. J Physiol 426: 19–42, 1990.

131. GANITKEVICH V AND ISENBERG G. Stimulation-induced potentiation ofT-type Ca2� channel currents in myocytes from guinea-pig coro-nary artery. J Physiol 443: 703–725, 1991.

132. GAO B, SEKIDO Y, MAXIMOV A, SAAD M, FORGACS E, LATIF F, LERMAN M,LEE JH, PEREZ-REYES E, BEZPROZVANNY I, AND MINNA JD. Functionalproperties of a new voltage-dependent calcium channel �2� auxil-iary subunit gene (CACNA2D2). J Biol Chem 275: 12237–12242,2000.

133. GARCıA J AND BEAM KG. Calcium transients associated with the Ttype calcium current in myotubes. J Gen Physiol 104: 1113–1128,1994.

134. GARCIA MA AND MEIZEL S. Progesterone-mediated calcium influxand acrosome reaction of human spermatozoa: pharmacologicalinvestigation of T-type calcium channels. Biol Reprod 60: 102–109,1999.

135. GAUCK V, THOMANN M, JAEGER D, AND BORST A. Spatial distribution oflow- and high-voltage-activated calcium currents in neurons of thedeep cerebellar nuclei. J Neurosci 21: 151–154, 2001.

136. GLOOR P AND FARIELLO RG. Generalized epilepsy: some of its cellularmechanisms differ from those of focal epilesy. Trends Neurosci 11:63–68, 1988.

137. GOMORA JC, DAUD AN, WEIERGRABER M, AND PEREZ-REYES E. Block ofcloned human T-type calcium channels by succinimide antiepilep-tic drugs. Mol Pharmacol 60: 1121–1132, 2001.

138. GOMORA JC, ENYEART JA, AND ENYEART JJ. Mibefradil potently blocksATP-activated K� channels in adrenal cells. Mol Pharmacol 56:1192–1197, 1999.

139. GOMORA JC AND ENYEART JJ. Dual pharmacological properties of acyclic AMP-sensitive potassium channel. J Pharmacol Exp Ther

290: 266–275, 1999.140. GOMORA JC, MURBARTIAN J, ARIAS JM, LEE JH, AND PEREZ-REYES E.

Cloning and expression of the human T-type channel Cav33: in-sights into prepulse facilitation. Biophys J 83: 229–241, 2002.

141. GOMORA JC, XU L, ENYEART JA, AND ENYEART JJ. Effect of mibefradilon voltage-dependent gating and kinetics of T-type Ca2� channelsin cortisol-secreting cells. J Pharmacol Exp Ther 292: 96–103, 2000.

142. GONOI T AND HASEGAWA S. Post-natal disappearance of transientcalcium channels in mouse skeletal muscle: effects of denervationand culture. J Physiol 401: 617–637, 1988.

143. GORDIENKO DV, CLAUSEN C, AND GOLIGORSKY MS. Ionic currents andendothelin signaling in smooth muscle cells from rat renal resis-tance arteries. Am J Physiol Renal Fluid Electrolyte Physiol 266:F325–F341, 1994.

144. GOTTSCHALK W, KIM DS, CHIN H, AND STANLEY EF. High-voltage-activated calcium channel messenger RNA expression in the 140–3neuroblastoma-glioma cell line. Neuroscience 94: 975–983, 1999.

145. GREEN PJ, WARRE R, HAYES PD, MCNAUGHTON NC, MEDHURST AD,PANGALOS M, DUCKWORTH DM, AND RANDALL AD. Kinetic modificationof the �1I subunit-mediated T-type Ca2� channel by a humanneuronal Ca2� channel � subunit. J Physiol 533: 467–478, 2001.

146. GREENE RW, HAAS HL, AND MCCARLEY RW. A low threshold calciumspike mediates firing pattern alterations in pontine reticular neu-rons. Science 234: 738–740, 1986.

147. GRIFFITH WH, TAYLOR L, AND DAVIS MJ. Whole-cell and single-chan-nel calcium currents in guinea pig basal forebrain neurons. J Neu-

rophysiol 71: 2359–2376, 1994.148. GROSS RA, WILEY JW, RYAN-JASTROW T, AND MACDONALD RL. Regula-

tion by GTP and its stable thiol derivatives of calcium currentcomponents in rat nodose ganglion neurons. Mol Pharmacol 37:546–553, 1990.

149. GUIA A, STERN MD, LAKATTA EG, AND JOSEPHSON IR. Ion concentra-tion-dependence of rat cardiac unitary L-type calcium channelconductance. Biophys J 80: 2742–2750, 2001.

150. GUSTAFSSON F, ANDREASEN D, SALOMONSSON M, JENSEN BL, AND HOL-STEIN-RATHLOU N. Conducted vasoconstriction in rat mesentericarterioles: role for dihydropyridine-insensitive Ca2� channels.Am J Physiol Heart Circ Physiol 280: H582–H590, 2001.

151. GUYON A, VERGNES M, AND LERESCHE N. Thalamic low thresholdcalcium current in a genetic model of absence epilepsy. Neurore-

port 4: 1231–1234, 1993.152. HAGIWARA N, IRISAWA H, AND KAMEYAMA M. Contribution of two types

of calcium currents to the pacemaker potentials of rabbit sino-atrial node cells. J Physiol 395: 233–253, 1988.

153. HAGIWARA S AND BYERLY L. Calcium channel. Annu Rev Neurosci 4:69–125, 1981.

154. HAGIWARA S, OZAWA S, AND SAND O. Voltage clamp analysis of twoinward current mechanisms in the egg cell membrane of a starfish.J Gen Physiol 65: 617–644, 1975.

155. HANSEN PB, JENSEN BL, ANDREASEN D, FRIIS UG, AND SKØTT O.Vascular smooth muscle cells express the �1A subunit of a P-/Q-type voltage-dependent Ca2� channel, and it is functionally impor-tant in renal afferent arterioles. Circ Res 87: 896–902, 2000.

156. HANSEN PB, JENSEN BL, ANDREASEN D, AND SKØTT O. Differential

154 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 39: Canal Calcio T

expression of T- and L-type voltage-dependent calcium channels inrenal resistance vessels. Circ Res 89: 630–638, 2001.

157. HAY M, HASSER EM, AND LINDSLEY KA. Area postrema voltage-acti-vated calcium currents. J Neurophysiol 75: 133–141, 1996.

158. HEADY TN, GOMORA JC, MACDONALD TL, AND PEREZ-REYES E. Molec-ular pharmacology of T-type Ca2� channels. Jpn J Pharmacol 85:339–350, 2001.

159. HERNANDEZ-CRUZ A AND PAPE HC. Identification of two calciumcurrents in acutely dissociated neurons from the rat lateral genic-ulate nucleus. J Neurophysiol 61: 1270–1283, 1989.

160. HERRINGTON J AND LINGLE CJ. Kinetic and pharmacological proper-ties of low voltage-activated Ca2� current in rat clonal (GH3)pituitary cells. J Neurophysiol 68: 213–232, 1992.

161. HESS P, LANSMAN JB, AND TSIEN RW. Calcium channel selectivity fordivalent and monovalent cations. Voltage and concentration depen-dence of single channel current in ventricular heart cells. J Gen

Physiol 88: 293–319, 1986.162. HEUBACH JF, KOHLER A, WETTWER E, AND RAVENS U. T-type and

tetrodotoxin-sensitive Ca2� currents coexist in guinea pig ventric-ular myocytes and are both blocked by mibefradil. Circ Res 86:628–635, 2000.

163. HIGASHIMA M, KINOSHITA H, AND KOSHINO Y. Contribution of T-typecalcium channels to afterdischarge generation in rat hippocampalslices. Brain Res 781: 127–134, 1998.

164. HILLE B. Ionic Channels of Excitable Membranes. Sunderland, MA:Sinauer, 2001. p. 814.

165. HIRANO Y, FOZZARD HA, AND JANUARY CT. Characteristics of L- andT-type Ca2� currents in canine cardiac Purkinje cells. Am J Physiol

Heart Circ Physiol 256: H1478–H1492, 1989.166. HIRANO Y, FOZZARD HA, AND JANUARY CT. Inactivation properties of

T-type calcium current in canine cardiac Purkinje cells. Biophys J

56: 1007–1016, 1989.167. HIRST GD, SILVERBERG GD, AND VAN HELDEN DF. The action potential

and underlying ionic currents in proximal rat middle cerebral ar-terioles. J Physiol 371: 289–304, 1986.

168. HIRST GDS AND EDWARDS FR. Sympathetic neuroeffector transmis-sion in arteries and arterioles. Physiol Rev 69: 546–604, 1989.

169. HOBOM M, DAI S, MARAIS E, LACINOVA L, HOFMANN F, AND KLUGBAUER

N. Neuronal distribution and functional characterization of thecalcium channel �2�2 subunit. Eur J Neurosci 12: 1217–1226, 2000.

170. HOEHN K, WATSON TW, AND MACVICAR BA. Multiple types of calciumchannels in acutely isolated rat neostriatal neurons. J Neurosci 13:1244–1257, 1993.

171. HOFMANN F, LACINOVA L, AND KLUGBAUER N. Voltage-dependent cal-cium channels: from structure to function. Rev Physiol Biochem

Pharmacol 139: 33–87, 1999.172. HONDA M, HAYASHI K, MATSUDA H, KUBOTA E, TOKUYAMA H, OKUBO K,

OZAWA Y, AND SARUTA T. Divergent natriuretic action of calciumchannel antagonists in mongrel dogs: renal haemodynamics as adeterminant of natriuresis. Clin Sci 101: 421–427, 2001.

173. HUANG B, QIN D, DENG L, BOUTJDIR M, AND EL-SHERIF N. Reexpres-sion of T-type Ca2� channel gene and current in post-infarctionremodeled rat left ventricle. Cardiovasc Res 46: 442–449, 2000.

174. HUBER I, WAPPL E, HERZOG A, MITTERDORFER J, GLOSSMANN H, LANGER

T, AND STRIESSNIG J. Conserved Ca2�-antagonist-binding propertiesand putative folding structure of a recombinant high-affinity dihy-dropyridine-binding domain. Biochem J 347: 829–836, 2000.

175. HUGUENARD JR. Low threshold calcium currents in central nervoussystem neurons. Annu Rev Physiol 58: 329–348, 1996.

176. HUGUENARD JR. Neuronal circuitry of thalamocortical epilepsy andmechanisms of antiabsence drug action. Adv Neurol 79: 991–999,1999.

177. HUGUENARD JR, GUTNICK MJ, AND PRINCE DA. Transient Ca2� cur-rents in neurons isolated from rat lateral habenula. J Neurophysiol

70: 158–166, 1993.178. HUGUENARD JR AND PRINCE DA. Intrathalamic rhythmicity studied in

vitro: nominal T-current modulation causes robust antioscillatoryeffects. J Neurosci 14: 5485–5502, 1994.

179. HUGUENARD JR AND PRINCE DA. A novel T-type current underliesprolonged Ca2�-dependent burst firing in GABAergic neurons of ratthalamic reticular nucleus. J Neurosci 12: 3804–3817, 1992.

180. HUSER J, BLATTER LA, AND LIPSIUS SL. Intracellular Ca2� release

contributes to automaticity in cat atrial pacemaker cells. J Physiol

524: 415–422, 2000.181. INOUE Y, OIKE M, NAKAO K, KITAMURA K, AND KURIYAMA H. Endothelin

augments unitary calcium channel currents on the smooth musclecell membrane of guinea-pig portal vein. J Physiol 423: 171–191,1990.

182. JAGANNATHAN S, PUNT EL, GU Y, ARNOULT C, SAKKAS D, BARRATT CLR,AND PUBLICOVER SJ. Identification and localization of T-type voltage-operated calcium channel subunits in human male germ cells.Expression of multiple isoforms. J Biol Chem 277: 8449–8456,2002.

183. JAHNSEN H AND LLINAS R. Electrophysiological properties of guinea-pig thalamic neurones: an in vitro study. J Physiol 349: 227–247,1984.

184. JAN LY AND JAN YN. A superfamily of ion channels. Nature 345: 672,1990.

185. JANSSEN LJ. T-type and L-type Ca2� currents in canine bronchialsmooth muscle: characterization and physiological roles. Am J

Physiol Cell Physiol 272: C1757–C1765, 1997.186. JEANMONOD D, MAGNIN M, MOREL A, SIEGEMUND M, CANCRO A, LANZ M,

LLINAS R, RIBARY U, KRONBERG E, SCHULMAN J, AND ZONENSHAYN M.Thalamocortical dysrhythmia. II. Clinical and surgical aspects.Thalamus Related Syst 1: 245–254, 2001.

187. JEN J. Calcium channelopathies in the central nervous system. Curr

Opin Neurobiol 9: 274–280, 1999.188. KANEDA M AND AKAIKE N. The low-threshold Ca current in isolated

amygdaloid neurons in the rat. Brain Res 497: 187–190, 1989.189. KANEDA M, WAKAMORI M, ITO C, AND AKAIKE N. Low-threshold cal-

cium current in isolated Purkinje cell bodies of rat cerebellum.J Neurophysiol 63: 1046–1051, 1990.

190. KANG MG, CHEN CC, FELIX R, LETTS VA, FRANKEL WN, MORI Y, AND

CAMPBELL KP. Biochemical and biophysical evidence for �2 subunitassociation with neuronal voltage-activated Ca2� channels. J Biol

Chem 276: 32917–32924, 2001.191. KANG Y AND KITAI ST. Calcium spike underlying rhythmic firing in

dopaminergic neurons of the rat substantia nigra. Neurosci Res 18:195–207, 1993.

192. KARAM H, CLOZEL JP, BRUNEVAL P, GONZALEZ MF, AND MENARD J.Contrasting effects of selective T- and L-type calcium channelblockade on glomerular damage in DOCA hypertensive rats. Hy-

pertension 34: 673–678, 1999.193. KARCESKI S, MORRELL M, AND CARPENTER D. The expert consensus

guideline series: treatment of epilepsy. Epilepsy Behav 2: A1–A50,2001.

194. KARSCHIN A AND LIPTON SA. Calcium channels in solitary retinalganglion cells from post-natal rat. J Physiol 418: 379–396, 1989.

195. KARST H, JOELS M, AND WADMAN WJ. Low-threshold calcium currentin dendrites of the adult rat hippocampus. Neurosci Lett 164:154–158, 1993.

196. KASAI H. Voltage- and time-dependent inhibition of neuronal cal-cium channels by a GTP-binding protein in a mammalian cell line.J Physiol 448: 189–209, 1992.

197. KASE M, KAKIMOTO S, SAKUMA S, HOUTANI T, OHISHI H, UEYAMA T, AND

SUGIMOTO T. Distribution of neurons expressing �1G subunit mRNAof T-type voltage-dependent calcium channel in adult rat centralnervous system. Neurosci Lett 268: 77–80, 1999.

198. KASS RS. Voltage-dependent modulation of cardiac calcium channelcurrent by optical isomers of Bay K 8644: implications for channelgating. Circ Res 61: I1–I5, 1987.

199. KAVALALI ET, ZHUO M, BITO H, AND TSIEN RW. Dendritic Ca2� chan-nels characterized by recordings from isolated hippocampal den-dritic segments. Neuron 18: 651–663, 1997.

200. KAWAI F. Odorants suppress T- and L-type Ca2� currents in olfac-tory receptor cells by shifting their inactivation curves to a negativevoltage. Neurosci Res 35: 253–263, 1999.

201. KAWAI F AND MIYACHI E. Enhancement by T-type Ca2� currents ofodor sensitivity in olfactory receptor cells. J Neurosci 21: 141–145,2001.

202. KAWANO S AND DEHAAN RL. Low-threshold current is major calciumcurrent in chick ventricle cells. Am J Physiol Heart Circ Physiol

256: H1505–1508, 1989.203. KEJA JA AND KITS KS. Single-channel properties of high- and low-

T-TYPE CALCIUM CHANNELS 155

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 40: Canal Calcio T

voltage-activated calcium channels in rat pituitary melanotropiccells. J Neurophysiol 71: 840–855, 1994.

204. KEJA JA, STOOF JC, AND KITS KS. Dopamine D2 receptor stimulationdifferentially affects voltage-activated calcium channels in rat pi-tuitary melanotropic cells. J Physiol 450: 409–435, 1992.

205. KETCHUM KA, JOINER WJ, SELLERS AJ, KACZMAREK LK, AND GOLDSTEIN

SA. A new family of outwardly rectifying potassium channel pro-teins with two pore domains in tandem. Nature 376: 690–695, 1995.

206. KIM D, SONG I, KEUM S, LEE T, JEONG MJ, KIM SS, MCENERY MW, AND

SHIN HS. Lack of the burst firing of thalamocortical relay neuronsand resistance to absence seizures in mice lacking �1G T-type Ca2�

channels. Neuron 31: 35–45, 2001.207. KIM MS, MORII T, SUN LX, IMOTO K, AND MORI Y. Structural determi-

nants of ion selectivity in brain calcium channel. FEBS Lett 318:145–148, 1993.

208. KING VF, GARCIA ML, SHEVELL JL, SLAUGHTER RS, AND KACZOROWSKI

GJ. Substituted diphenylbutylpiperidines bind to a unique highaffinity site on the L-type calcium channel. Evidence for a fourthsite in the cardiac calcium entry blocker receptor complex. J Biol

Chem 264: 5633–5641, 1989.209. KLOCKNER U, LEE JH, CRIBBS LL, DAUD A, HESCHELER J, PEREVERZEV A,

PEREZ-REYES E, AND SCHNEIDER T. Comparison of the Ca2� currentsinduced by expression of three cloned �1 subunits, �1G, �1H and�1I, of low-voltage-activated T-type Ca2� channels. Eur J Neurosci

11: 4171–4178, 1999.210. KLUGBAUER N, DAI S, SPECHT V, LACINOVA L, MARAIS E, BOHN G, AND

HOFMANN F. A family of �-like calcium channel subunits. FEBS Lett

470: 189–197, 2000.211. KLUGBAUER N, MARAIS E, LACINOVA L, AND HOFMANN F. A T-type

calcium channel from mouse brain. Pflugers Arch 437: 710–715,1999.

212. KNUDSON CM, CHAUDHARI N, SHARP AH, POWELL JA, BEAM KG, AND

CAMPBELL KP. Specific absence of the alpha 1 subunit of the dihy-dropyridine receptor in mice with muscular dysgenesis. J Biol

Chem 264: 1345–1348, 1989.213. KOBRINSKY EM, PEARSON HA, AND DOLPHIN AC. Low- and high-volt-

age-activated calcium channel currents and their modulation in thedorsal root ganglion cell line ND7–23. Neuroscience 58: 539–552,1994.

214. KOH SD, MONAGHAN K, RO S, MASON HS, KENYON JL, AND SANDERS KM.Novel voltage-dependent non-selective cation conductance in mu-rine colonic myocytes. J Physiol 533: 341–355, 2001.

215. KOSTYUK PG, MOLOKANOVA EA, PRONCHUK NF, SAVCHENKO AN, AND

VERKHRATSKY AN. Different action of ethosuximide on low- andhigh-threshold calcium currents in rat sensory neurons. Neuro-

science 51: 755–758, 1992.216. KOSTYUK PG AND SHIROKOV RE. Deactivation kinetics of different

components of calcium inward current in the membrane of micesensory neurones. J Physiol 409: 343–355, 1989.

217. KOSTYUK PG, SHUBA YA M, AND SAVCHENKO AN. Three types ofcalcium channels in the membrane of mouse sensory neurons.Pflugers Arch 411: 661–669, 1988.

218. KOZLOV AS, MCKENNA F, LEE JH, CRIBBS LL, PEREZ-REYES E, FELTZ A,AND LAMBERT RC. Distinct kinetics of cloned T-type Ca2� channelslead to differential Ca2� entry and frequency-dependence duringmock action potentials. Eur J Neurosci 11: 4149–4158, 1999.

219. KRAYENBUHL JC, VOZEH S, KONDO-OESTREICHER M, AND DAYER P. Drug-drug interactions of new active substances: mibefradil example.Eur J Clin Pharmacol 55: 559–565, 1999.

220. KUGA T, KOBAYASHI S, HIRAKAWA Y, KANAIDE H, AND TAKESHITA A. Cellcycle-dependent expression of L- and T-type Ca2� currents in rataortic smooth muscle cells in primary culture. Circ Res 79: 14–19,1996.

221. KUGA T, SADOSHIMA J, TOMOIKE H, KANAIDE H, AKAIKE N, AND NAKA-MURA M. Actions of Ca2� antagonists on two types of Ca2� channelsin rat aorta smooth muscle cells in primary culture. Circ Res 67:469–480, 1990.

222. KUO CC AND YANG S. Recovery from inactivation of T-type Ca2�

channels in rat thalamic neurons. J Neurosci 21: 1884–1892, 2001.223. LACERDA AE, KIM HS, RUTH P, PEREZ-REYES E, FLOCKERZI V, HOFMANN

F, BIRNBAUMER L, AND BROWN AM. Normalization of current kineticsby interaction between the �1 and � subunits of the skeletal muscledihydropyridine-sensitive Ca2� channel. Nature 352: 527–530, 1991.

224. LACINOVA L, KLUGBAUER N, AND HOFMANN F. Absence of modulationof the expressed calcium channel �1G subunit by �2� subunits.J Physiol 516: 639–645, 1999.

225. LALEVEE N, PLUCIENNIK F, AND JOFFRE M. Voltage-dependent calciumcurrent with properties of T-type current in Sertoli cells fromimmature rat testis in primary cultures. Biol Reprod 56: 680–687,1997.

226. LAMBERT RC, MAULET Y, MOUTON J, BEATTIE R, VOLSEN S, DE WAARD

M, AND FELTZ A. T-type Ca2� current properties are not modified byCa2� channel � subunit depletion in nodosus ganglion neurons.J Neurosci 17: 6621–6628, 1997.

227. LEE JH, CRIBBS LL, AND PEREZ-REYES E. Cloning of a novel fourrepeat protein related to voltage-gated sodium and calcium chan-nels. FEBS Lett 445: 231–236, 1999.

228. LEE JH, DAUD AN, CRIBBS LL, LACERDA AE, PEREVERZEV A, KLOCKNER

U, SCHNEIDER T, AND PEREZ-REYES E. Cloning and expression of anovel member of the low voltage-activated T-type calcium channelfamily. J Neurosci 19: 1912–1921, 1999.

229. LEE JH, FREIHAGE J, CRIBBS LL, AND PEREZ-REYES E. Tissue distribu-tion of four CavT 1 (�1G) splice variants of domain III-IV loop(Abstract). Biophys J 76: A408, 1999.

230. LEE JH, GOMORA JC, CRIBBS LL, AND PEREZ-REYES E. Nickel block ofthree cloned T-type calcium channels: low concentrations selec-tively block �1H. Biophys J 77: 3034–3042, 1999.

231. LEE JH, KIM EG, PARK BG, KIM KH, CHA SK, KONG ID, LEE JW, AND

JEONG SW. Identification of T-type �1H Ca2� channels (Cav3.2) inmajor pelvic ganglion neurons. J Neurophysiol 87: 2844–2850,2002.

232. LERESCHE N, PARRI HR, ERDEMLI G, GUYON A, TURNER JP, WILLIAMS

SR, ASPRODINI E, AND CRUNELLI V. On the action of the anti-absencedrug ethosuximide in the rat and cat thalamus. J Neurosci 18:4842–4853, 1998.

233. LETTS VA, FELIX R, BIDDLECOME GH, ARIKKATH J, MAHAFFEY CL,VALENZUELA A, BARTLETT FS II, MORI Y, CAMPBELL KP, AND FRANKEL

WN. The mouse stargazer gene encodes a neuronal Ca2�-channel �subunit. Nat Genet 19: 340–347, 1998.

234. LEURANGUER V, BOURINET E, LORY P, AND NARGEOT J. Antisensedepletion of �-subunits fails to affect T-type calcium channel prop-erties in a neuroblastoma cell line. Neuropharmacology 37: 701–708, 1998.

235. LEURANGUER V, MANGONI ME, NARGEOT J, AND RICHARD S. Inhibitionof T-type and L-type calcium channels by mibefradil: physiologicand pharmacologic bases of cardiovascular effects. J Cardiovasc

Pharmacol 37: 649–661, 2001.236. LEURANGUER V, MONTEIL A, BOURINET E, DAYANITHI G, AND NARGEOT J.

T-type calcium currents in rat cardiomyocytes during postnataldevelopment: contribution to hormone secretion. Am J Physiol

Heart Circ Physiol 279: H2540–H2548, 2000.237. LIEVANO A, BOLDEN A, AND HORN R. Calcium channels in excitable

cells: divergent genotypic and phenotypic expression of alpha1-subunits. Am J Physiol Cell Physiol 267: C411–C424, 1994.

238. LIGON B, BOYD AE III, AND DUNLAP K. Class A calcium channelvariants in pancreatic islets and their role in insulin secretion.J Biol Chem 273: 13905–13911, 1998.

239. LLEDO PM, HOMBURGER V, BOCKAERT J, AND VINCENT JD. DifferentialG protein-mediated coupling of D2 dopamine receptors to K� andCa2� currents in rat anterior pituitary cells. Neuron 8: 455–463,1992.

240. LLEDO PM, LEGENDRE P, ISRAEL JM, AND VINCENT JD. Dopamineinhibits two characterized voltage-dependent calcium currents inidentified rat lactotroph cells. Endocrinology 127: 990–1001, 1990.

241. LLINAS R AND MUHLETHALER M. Electrophysiology of guinea-pig cer-ebellar nuclear cells in the in vitro brain stem-cerebellar prepara-tion. J Physiol 404: 241–258, 1988.

242. LLINAS R, RIBARY U, JEANMONOD D, CANCRO R, KRONBERG E, SCHULMAN

J, ZONENSHAYN M, MAGNIN M, MOREL A, AND SIEGMUND M. Thalamo-cortical dysrhythmia. I. Functional and imaging aspects. Thalamus

Related Syst 1: 237–244, 2001.243. LLINAS R AND YAROM Y. Electrophysiology of mammalian inferior

olivary neurones in vitro. Different types of voltage-dependentionic conductances. J Physiol 315: 549–567, 1981.

244. LLINAS R AND YAROM Y. Properties and distribution of ionic conduc-

156 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 41: Canal Calcio T

tances generating electroresponsiveness of mammalian inferiorolivary neurones in vitro. J Physiol 315: 569–584, 1981.

245. LLINAS RR. The intrinsic electrophysiological properties of mamma-lian neurons: insights into central nervous system function. Science

242: 1654–1664, 1988.246. LOIRAND G, MIRONNEAU C, MIRONNEAU J, AND PACAUD P. Two types of

calcium currents in single smooth muscle cells from rat portal vein.J Physiol 412: 333–349, 1989.

247. LOPEZ-GONZALEZ I, DE LA VEGA-BELTRAN JL, SANTI CM, FLORMAN HM,FELIX R, AND DARSZON A. Calmodulin antagonists inhibit T-type Ca2�

currents in mouse spermatogenic cells and the zona pellucida-induced sperm acrosome reaction. Dev Biol 236: 210–219, 2001.

248. LORENZON NM AND BEAM KG. Calcium channelopathies. Kidney Int

57: 794–802, 2000.249. LORY P, OPHOFF RA, AND NAHMIAS J. Towards a unified nomenclature

describing voltage-gated calcium channel genes. Hum Genet 100:149–150, 1997.

250. LOTSHAW DP. Role of membrane depolarization and T-type Ca2�

channels in angiotensin II and K� stimulated aldosterone secretion.Mol Cell Endocrinol 175: 157–171, 2001.

251. LU HK, FERN RJ, LUTHIN D, LINDEN J, LIU LP, COHEN CJ, AND BARRETT

PQ. Angiotensin II stimulates T-type Ca2� channel currents viaactivation of a G protein, Gi. Am J Physiol Cell Physiol 271:C1340–C1349, 1996.

252. LU HK, FERN RJ, NEE JJ, AND BARRETT PQ. Ca2�-dependent activa-tion of T-type Ca2� channels by calmodulin-dependent proteinkinase II. Am J Physiol Renal Fluid Electrolyte Physiol 267: F183–F189, 1994.

253. LUTHER JA AND TASKER JG. Voltage-gated currents distinguish par-vocellular from magnocellular neurones in the rat hypothalamicparaventricular nucleus. J Physiol 523: 193–209, 2000.

254. LUTHI A AND MCCORMICK DA. H-current: properties of a neuronal andnetwork pacemaker. Neuron 21: 9–12, 1998.

255. LUX HD, CARBONE E, AND ZUCKER H. Na� currents through low-voltage-activated Ca2� channels of chick sensory neurones: blockby external Ca2� and Mg2�. J Physiol 430: 159–188, 1990.

256. MAGEE JC AND JOHNSTON D. Synaptic activation of voltage-gatedchannels in the dendrites of hippocampal pyramidal neurons. Sci-

ence 268: 301–304, 1995.257. MAGISTRETTI J AND DE CURTIS M. Low-voltage activated T-type cal-

cium currents are differently expressed in superficial and deeplayers of guinea pig piriform cortex. J Neurophysiol 79: 808–816,1998.

258. MANSVELDER HD AND KITS KS. All classes of calcium channel couplewith equal efficiency to exocytosis in rat melanotropes, inducinglinear stimulus-secretion coupling. J Physiol 526: 327–339, 2000.

259. MARCHETTI C, AMICO C, PODESTA D, AND ROBELLO M. Inactivation ofvoltage-dependent calcium current in an insulinoma cell line. Eur

Biophys J 23: 51–58, 1994.260. MARCHETTI C, CARBONE E, AND LUX HD. Effects of dopamine and

noradrenaline on Ca channels of cultured sensory and sympatheticneurons of chick. Pflugers Arch 406: 104–111, 1986.

261. MARCHETTI C, CHILDS GV, AND BROWN AM. Membrane currents ofidentified isolated rat corticotropes and gonadotropes. Am J

Physiol Endocrinol Metab 252: E340–E346, 1987.262. MARCHETTI C, CHILDS GV, AND BROWN AM. Voltage-dependent cal-

cium currents in rat gonadotropes separated by centrifugal elutria-tion. Am J Physiol Endocrinol Metab 258: E589–E596, 1990.

263. MARGETA-MITROVIC M, GRIGG JJ, KOYANO K, NAKAJIMA Y, AND NAKAJIMA

S. Neurotensin and substance P inhibit low- and high-voltage-activated Ca2� channels in cultured newborn rat nucleus basalisneurons. J Neurophysiol 78: 1341–1352, 1997.

264. MARIOT P, VANOVERBERGHE K, LALEVEE N, ROSSIER MF, AND PREVAR-SKAYA N. Overexpression of an �1H (Cav3.2) T-type calcium channelduring neuroendocrine differentiation of human prostate cancercells. J Biol Chem 277: 10824–10833, 2002.

265. MARKRAM H AND SAKMANN B. Calcium transients in dendrites ofneocortical neurons evoked by single subthreshold excitatorypostsynaptic potentials via low-voltage-activated calcium channels.Proc Natl Acad Sci USA 91: 5207–5211, 1994.

266. MARKSTEINER R, SCHURR P, BERJUKOW S, MARGREITER E, PEREZ-REYES

E, AND HERING S. Inactivation determinants in segment IIIS6 ofCav3.1. J Physiol 537: 27–34, 2001.

267. MARTIN RL, LEE JH, CRIBBS LL, PEREZ-REYES E, AND HANCK DA.Mibefradil block of cloned T-type calcium channels. J Pharmacol

Exp Ther 295: 302–308, 2000.268. MARTıNEZ ML, HEREDIA MP, AND DELGADO C. Expression of T-type

Ca2� channels in ventricular cells from hypertrophied rat hearts. J

Mol Cell Cardiol 31: 1617–1625, 1999.269. MATSUDA JJ, VOLK KA, AND SHIBATA EF. Calcium currents in isolated

rabbit coronary arterial smooth muscle myocytes. J Physiol 427:657–680, 1990.

270. MATTESON DR AND ARMSTRONG CM. Properties of two types of cal-cium channels in clonal pituitary cells. J Gen Physiol 87: 161–182,1986.

271. MAYLIE J AND MORAD M. Evaluation of T- and L-type Ca2� currentsin shark ventricular myocytes. Am J Physiol Heart Circ Physiol

269: H1695–H1703, 1995.272. MCCARTHY RT AND COHEN CJ. Nimodipine block of calcium channels

in rat vascular smooth muscle cell lines. Exceptionally high-affinitybinding in A7r5 and A10 cells. J Gen Physiol 94: 669–692, 1989.

273. MCCARTHY RT, ISALES C, AND RASMUSSEN H. T-type calcium channelsin adrenal glomerulosa cells: GTP-dependent modulation by angio-tensin II. Proc Natl Acad Sci USA 90: 3260–3264, 1993.

274. MCCARTHY RT, ISALES CM, BOLLAG WB, RASMUSSEN H, AND BARRETT

PQ. Atrial natriuretic peptide differentially modulates T- and L-typecalcium channels. Am J Physiol Renal Fluid Electrolyte Physiol

258: F473–F478, 1990.275. MCCLESKEY EW. Calcium channel permeation: a field in flux. J Gen

Physiol 113: 765–772, 1999.276. MCCOBB DP AND BEAM KG. Action potential waveform voltage-

clamp commands reveal striking differences in calcium entry vialow and high voltage-activated calcium channels. Neuron 7: 119–127, 1991.

277. MCCORMICK DA AND BAL T. Sleep and arousal: thalamocorticalmechanisms. Annu Rev Neurosci 20: 185–215, 1997.

278. MCDONOUGH SI AND BEAN BP. Mibefradil inhibition of T-type cal-cium channels in cerebellar Purkinje neurons. Mol Pharmacol 54:1080–1087, 1998.

279. MCRORY JE, SANTI CM, HAMMING KS, MEZEYOVA J, SUTTON KG, BAILLIE

DL, STEA A, AND SNUTCH TP. Molecular and functional characteriza-tion of a family of rat brain T-type calcium channels. J Biol Chem

276: 3999–4011, 2001.280. MICHELS G, MATTHES J, HANDROCK R, KUCHINKE U, GRONER F, CRIBBS

LL, PEREVERZEV A, SCHNEIDER T, PEREZ-REYES E, AND HERZIG S. Single-channel pharmacology of mibefradil in human native T-type andrecombinant Cav3.2 calcium channels. Mol Pharmacol 61: 682–694,2002.

281. MISHRA SK AND HERMSMEYER K. Inhibition of signal Ca2� in dogcoronary arterial vascular muscle cells by Ro 40–5967. J Cardio-

vasc Pharmacol 24: 1–7, 1994.282. MISHRA SK AND HERMSMEYER K. Selective inhibition of T-type Ca2�

channels by Ro 40–5967. Circ Res 75: 144–148, 1994.283. MITRA R AND MORAD M. Two types of calcium channels in guinea pig

ventricular myocytes. Proc Natl Acad Sci USA 83: 5340–5344, 1986.284. MITTMAN S, GUO J, AND AGNEW WS. Structure and alternative splicing

of the gene encoding �1G, a human brain T calcium channel �1subunit. Neurosci Lett 274: 143–146, 1999.

285. MITTMAN S, GUO J, EMERICK MC, AND AGNEW WS. Structure andalternative splicing of the gene encoding �1I, a human brain Tcalcium channel �1 subunit. Neurosci Lett 269: 121–124, 1999.

286. MLINAR B, BIAGI BA, AND ENYEART JJ. A novel K� current inhibited byadrenocorticotropic hormone and angiotensin II in adrenal corticalcells. J Biol Chem 268: 8640–8644, 1993.

287. MLINAR B, BIAGI BA, AND ENYEART JJ. Voltage-gated transient cur-rents in bovine adrenal fasciculata cells. I. T-type Ca2� current.J Gen Physiol 102: 217–237, 1993.

288. MONTEIL A, CHEMIN J, BOURINET E, MENNESSIER G, LORY P, AND

NARGEOT J. Molecular and functional properties of the human �1Gsubunit that forms T-type calcium channels. J Biol Chem 275:6090–6100, 2000.

289. MONTEIL A, CHEMIN J, LEURANGUER V, ALTIER C, MENNESSIER G, BOU-RINET E, LORY P, AND NARGEOT J. Specific properties of T-typecalcium channels generated by the human �1I subunit. J Biol Chem

275: 16530–16535, 2000.290. MOOLENAAR WH AND SPECTOR I. Ionic currents in cultured mouse

T-TYPE CALCIUM CHANNELS 157

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 42: Canal Calcio T

neuroblastoma cells under voltage-clamp conditions. J Physiol 278:265–286, 1978.

291. MORITA H, COUSINS H, ONOUE H, ITO Y, AND INOUE R. Predominantdistribution of nifedipine-insensitive, high voltage-activated Ca2�

channels in the terminal mesenteric artery of guinea pig. Circ Res

85: 596–605, 1999.292. MOUGINOT D, BOSSU JL, AND GAHWILER BH. Low-threshold Ca2�

currents in dendritic recordings from Purkinje cells in rat cerebel-lar slice cultures. J Neurosci 17: 160–170, 1997.

293. MULLER TH, MISGELD U, AND SWANDULLA D. Ionic currents in culturedrat hypothalamic neurones. J Physiol 450: 341–362, 1992.

294. MUNSCH T, BUDDE T, AND PAPE HC. Voltage-activated intracellularcalcium transients in thalamic relay cells and interneurons. Neu-

roreport 8: 2411–2418, 1997.295. NAKASHIMA YM, TODOROVIC SM, PEREVERZEV A, HESCHELER J, SCHNEI-

DER T, AND LINGLE CJ. Properties of Ba2� currents arising fromhuman �E and �1E�3 constructs expressed in HEK293 cells: phys-iology, pharmacology, and comparison to native T-type Ba2� cur-rents. Neuropharmacology 37: 957–972, 1998.

296. NAMBU A AND LLINAS R. Electrophysiology of globus pallidus neu-rons in vitro. J Neurophysiol 72: 1127–1139, 1994.

297. NARAHASHI T. Neuroreceptors and ion channels as the basis for drugaction: past, present, and future. J Pharmacol Exp Ther 294: 1–26,2000.

298. NARAHASHI T, TSUNOO A, AND YOSHII M. Characterization of two typesof calcium channels in mouse neuroblastoma cells. J Physiol 383:231–249, 1987.

299. NEWCOMB R, SZOKE B, PALMA A, WANG G, CHEN X, HOPKINS W, CONG

R, MILLER J, URGE L, TARCZY-HORNOCH K, LOO JA, DOOLEY DJ, NADASDI

L, TSIEN RW, LEMOS J, AND MILJANICH G. Selective peptide antagonistof the class E calcium channel from the venom of the tarantulaHysterocrates gigas. Biochemistry 37: 15353–15362, 1998.

300. NICHOLS CG, SHYNG SL, NESTOROWICZ A, GLASER B, CLEMENT JP,GONZALEZ G, AGUILAR-BRYAN L, PERMUTT MA, AND BRYAN J. Adenosinediphosphate as an intracellular regulator of insulin secretion. Sci-

ence 272: 1785–1787, 1996.301. NIESEN CE AND GE S. Chronic epilepsy in developing hippocampal

neurons: electrophysiologic and morphologic features. Dev Neuro-

sci 21: 328–338, 1999.302. NIESPODZIANY I, DERAMBURE P, AND POULAIN P. Properties of T-type

calcium current in enkephalinergic neurones in guinea-pig hypo-thalamic slices. Pflugers Arch 437: 871–880, 1999.

303. NILIUS B, HESS P, LANSMAN JB, AND TSIEN RW. A novel type of cardiaccalcium channel in ventricular cells. Nature 316: 443–446, 1985.

304. NILIUS B, PRENEN J, KAMOUCHI M, VIANA F, VOETS T, AND DROOGMANS

G. Inhibition by mibefradil, a novel calcium channel antagonist, ofCa2�- and volume-activated Cl� channels in macrovascular endo-thelial cells. Br J Pharmacol 121: 547–555, 1997.

305. NOBILE M, CARBONE E, LUX HD, AND ZUCKER H. Temperature sensi-tivity of Ca currents in chick sensory neurones. Pflugers Arch 415:658–663, 1990.

306. NOWYCKY MC, FOX AP, AND TSIEN RW. Three types of neuronalcalcium channel with different calcium agonist sensitivity. Nature

316: 440–443, 1985.307. NUNEZ A, GARCıA-AUSTT E, AND BUNO W. Slow intrinsic spikes re-

corded in vivo in rat CA1-CA3 hippocampal pyramidal neurons.Exp Neurol 109: 294–299, 1990.

308. NUSS HB AND HOUSER SR. T-type Ca2� current is expressed inhypertrophied adult feline left ventricular myocytes. Circ Res 73:777–782, 1993.

309. NUSSINOVITCH I AND KLEINHAUS AL. Dopamine inhibits voltage-acti-vated calcium channel currents in rat pars intermedia pituitarycells. Brain Res 574: 49–55, 1992.

310. O’DELL TJ AND ALGER BE. Single calcium channels in rat and guinea-pig hippocampal neurons. J Physiol 436: 739–767, 1991.

311. OHYA Y, ABE I, FUJII K, TAKATA Y, AND FUJISHIMA M. Voltage-depen-dent Ca2� channels in resistance arteries from spontaneously hy-pertensive rats. Circ Res 73: 1090–1099, 1993.

312. OSIPENKO ON, VARNAI P, MIKE A, SPAT A, AND VIZI ES. Dopamineblocks T-type calcium channels in cultured rat adrenal glomerulosacells. Endocrinology 134: 511–514, 1994.

313. OUADID H, SEGUIN J, RICHARD S, CHAPTAL PA, AND NARGEOT J. Prop-

erties and modulation of Ca channels in adult human atrial cells. J

Mol Cell Cardiol 23: 41–54, 1991.314. OZAWA Y, HAYASHI K, NAGAHAMA T, FUJIWARA K, AND SARUTA T. Effect

of T-type selective calcium antagonist on renal microcirculation:studies in the isolated perfused hydronephrotic kidney. Hyperten-

sion 38: 343–347, 2001.315. PACAUD P, LOIRAND G, MIRONNEAU C, AND MIRONNEAU J. Opposing

effects of noradrenaline on the two classes of voltage-dependentcalcium channels of single vascular smooth muscle cells in short-term primary culture. Pflugers Arch 410: 557–559, 1987.

316. PAN ZH. Differential expression of high- and two types of low-voltage-activated calcium currents in rod and cone bipolar cells ofthe rat retina. J Neurophysiol 83: 513–527, 2000.

317. PAN ZH, HU HJ, PERRING P, AND ANDRADE R. T-type Ca2� channelsmediate neurotransmitter release in retinal bipolar cells. Neuron

32: 89–98, 2001.318. PAPE HC, BUDDE T, MAGER R, AND KISVARDAY ZF. Prevention of

Ca2�-mediated action potentials in GABAergic local circuit neu-rones of rat thalamus by a transient K� current. J Physiol 478:403–422, 1994.

319. PARENT L AND GOPALAKRISHNAN M. Glutamate substitution in repeatIV alters divalent and monovalent cation permeation in the heartCa2� channel. Biophys J 69: 1801–1813, 1995.

320. PARSEY RV AND MATTESON DR. Ascorbic acid modulation of calciumchannels in pancreatic beta cells. J Gen Physiol 102: 503–523, 1993.

321. PATIL PG, BRODY DL, AND YUE DT. Preferential closed-state inacti-vation of neuronal calcium channels. Neuron 20: 1027–1038, 1998.

322. PEMBERTON KE, HILL-EUBANKS LJ, AND JONES SV. Modulation of low-threshold T-type calcium channels by the five muscarinic receptorsubtypes in NIH 3T3 cells. Pflugers Arch 440: 452–461, 2000.

323. PERCHENET L, BENARDEAU A, AND ERTEL EA. Pharmacological prop-erties of Cav3.2, a low voltage-activated Ca2� channel cloned fromhuman heart. Naunyn-Schmiedeberg’s Arch Pharmacol 361: 590–599, 2000.

324. PEREZ-REYES E, CRIBBS LL, DAUD A, LACERDA AE, BARCLAY J, WILLIAM-SON MP, FOX M, REES M, AND LEE JH. Molecular characterization ofa neuronal low voltage-activated T-type calcium channel. Nature

391: 896–900, 1998.325. PEREZ-REYES E, LEE JH, DAUD A, AND CRIBBS LL. Cloning of calcium

channels representing two new subfamilies distinct from high volt-age-activated types (Abstract). Biophys J 72: A146, 1997.

326. PEREZ-REYES E AND SCHNEIDER T. Calcium channels: structure, func-tion, and classification. Drug Dev Res 33: 295–318, 1994.

327. PETKOV GV, FUSI F, SAPONARA S, GAGOV HS, SGARAGLI GP, AND BOEV

KK. Characterization of voltage-gated calcium currents in freshlyisolated smooth muscle cells from rat tail main artery. Acta Physiol

Scand 173: 257–265, 2001.328. POUILLE F, CAVELIER P, DESPLANTEZ T, BEEKENKAMP H, CRAIG PJ,

BEATTIE RE, VOLSEN SG, AND BOSSU JL. Dendro-somatic distributionof calcium-mediated electrogenesis in Purkinje cells from rat cer-ebellar slice cultures. J Physiol 527: 265–282, 2000.

329. PROTAS L, DIFRANCESCO D, AND ROBINSON RB. L-type but not T-typecalcium current changes during postnatal development in rabbitsinoatrial node. Am J Physiol Heart Circ Physiol 281: H1252–H1259, 2001.

330. PUBLICOVER SJ, PRESTON MR, AND EL HAJ AJ. Voltage-dependentpotentiation of low-voltage-activated Ca2� channel currents in cul-tured rat bone marrow cells. J Physiol 489: 649–661, 1995.

331. QU Y AND BOUTJDIR M. Gene expression of SERCA2a and L- andT-type Ca channels during human heart development. Pediatr Res

50: 569–574, 2001.332. QUIGNARD JF, FRAPIER JM, HARRICANE MC, ALBAT B, NARGEOT J, AND

RICHARD S. Voltage-gated calcium channel currents in human cor-onary myocytes. Regulation by cyclic GMP and nitric oxide. J Clin

Invest 99: 185–193, 1997.333. QUIGNARD JF, HARRICANE MC, MENARD C, LORY P, NARGEOT J, CAPRON

L, MORNET D, AND RICHARD S. Transient down-regulation of L-typeCa2� channel and dystrophin expression after balloon injury in rataortic cells. Cardiovasc Res 49: 177–188, 2001.

334. RANDALL AD AND TSIEN RW. Contrasting biophysical and pharmaco-logical properties of T-type and R-type calcium channels. Neuro-

pharmacology 36: 879–893, 1997.335. REN D, NAVARRO B, PEREZ G, JACKSON AC, HSU S, SHI Q, TILLY JL, AND

158 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 43: Canal Calcio T

CLAPHAM DE. A sperm ion channel required for sperm motility andmale fertility. Nature 413: 603–609, 2001.

336. RICHARD S, DIOCHOT S, NARGEOT J, BALDY-MOULINIER M, AND VALMIER

J. Inhibition of T-type calcium currents by dihydropyridines inmouse embryonic dorsal root ganglion neurons. Neurosci Lett 132:229–234, 1991.

337. RICHARD S, LECLERCQ F, LEMAIRE S, PIOT C, AND NARGEOT J. Ca2�

currents in compensated hypertrophy and heart failure. Cardio-

vasc Res 37: 300–311, 1998.338. RICHARD S AND NARGEOT J. T-type calcium currents in vascular

smooth muscle cells: a role in cellular proliferation? In: Low Volt-

age-activated T-type Calcium Channels, edited by Tsien RW, Clo-zel JP, and Nargeot J. Chester, UK: Adis International, 1998, p.123–132.

339. RICHARD S, NEVEU D, CARNAC G, BODIN P, TRAVO P, AND NARGEOT J.Differential expression of voltage-gated Ca2�-currents in cultivatedaortic myocytes. Biochim Biophys Acta 1160: 95–104, 1992.

340. RIOS E AND PIZARRO G. Voltage sensor of excitation-contractioncoupling in skeletal muscle. Physiol Rev 71: 849–908, 1991.

341. ROMANIN C, SEYDL K, GLOSSMANN H, AND SCHINDLER H. The dihydro-pyridine niguldipine inhibits T-type Ca2� currents in atrial myo-cytes. Pflugers Arch 420: 410–412, 1992.

342. ROOPRA A, HUANG Y, AND DINGLEDINE R. Neurological disease: listen-ing to gene silencers. Mol Interventions 1: 219–228, 2001.

343. ROSEN AD. Temperature modulation of calcium channel function inGH3 cells. Am J Physiol Cell Physiol 271: C863–C868, 1996.

344. ROSSIER MF, APTEL HB, PYTHON CP, BURNAY MM, VALLOTTON MB, AND

CAPPONI AM. Inhibition of low threshold calcium channels by an-giotensin II in adrenal glomerulosa cells through activation ofprotein kinase C. J Biol Chem 270: 15137–15142, 1995.

345. ROSSIER MF, ERTEL EA, VALLOTTON MB, AND CAPPONI AM. Inhibitoryaction of mibefradil on calcium signaling and aldosterone synthesisin bovine adrenal glomerulosa cells. J Pharmacol Exp Ther 287:824–831, 1998.

346. ROUSSET M, CENS T, RESTITUITO S, BARRERE C, BLACK JL, MCENERY

MW, AND CHARNET P. Functional roles of �2, �3, and �4, three newCa2� channel subunits, in P/Q-type Ca2� channel expressed inXenopus oocyte. J Physiol 532: 583–593, 2001.

347. RYU PD AND RANDIC M. Low- and high-voltage-activated calciumcurrents in rat spinal dorsal horn neurons. J Neurophysiol 63:273–285, 1990.

348. SALA S AND MATTESON DR. Single-channel recordings of two types ofcalcium channels in rat pancreatic beta-cells. Biophys J 58: 567–571, 1990.

349. SANTI CM, CAYABYAB FS, SUTTON KG, MCRORY JE, MEZEYOVA J, HAM-MING KS, PARKER D, STEA A, AND SNUTCH TP. Differential inhibition ofT-type calcium channels by neuroleptics. J Neurosci 22: 396–403,2002.

350. SANTI CM, DARSZON A, AND HERNANDEZ-CRUZ A. A dihydropyridine-sensitive T-type Ca2� current is the main Ca2� current carrier inmouse primary spermatocytes. Am J Physiol Cell Physiol 271:C1583–C1593, 1996.

351. SATIN J AND CRIBBS LL. Identification of a T-type Ca2� channelisoform in murine atrial myocytes (AT-1 cells). Circ Res 86: 636–642, 2000.

352. SATIN LS AND SMOLEN PD. Electrical bursting in �-cells of the pan-creatic islets of Langerhans. Endocrine 2: 677–687, 1994.

353. SATOH H. Role of T-type Ca2� channel inhibitors in the pacemakerdepolarization in rabbit sino-atrial nodal cells. Gen Pharmacol 26:581–587, 1995.

354. SAYER RJ, BROWN AM, SCHWINDT PC, AND CRILL WE. Calcium currentsin acutely isolated human neocortical neurons. J Neurophysiol 69:1596–1606, 1993.

355. SCHMITT R, CLOZEL JP, IBERG N, AND BUHLER FR. Mibefradil preventsneointima formation after vascular injury in rats. Possible role ofthe blockade of the T-type voltage-operated calcium channel. Ar-

terioscler Thromb Vasc Biol 15: 1161–1165, 1995.356. SCHMITT R, KLEINBLOESEM CH, BELZ GG, SCHROETER V, FEIFEL U,

POZENEL H, KIRCH W, HALABI A, WOITTIEZ AJ, WELKER HA, AND VAN

BRUMMELEN P. Hemodynamic and humoral effects of the novelcalcium antagonist Ro 40–5967 in patients with hypertension. Clin

Pharmacol Ther 52: 314–323, 1992.357. SCHNEIDER T, WEI X, OLCESE R, COSTANTIN JL, NEELY A, PALADE P,

PEREZ-REYES E, QIN N, ZHOU J, CRAWFORD GD, SMITH RG, APPEL SH,STEFANI E, AND BIRNBAUMER L. Molecular analysis and functionalexpression of the human type E �1 subunit. Receptors Channels 2:255–270, 1994.

358. SCHRIER AD, WANG H, TALLEY EM, PEREZ-REYES E, AND BARRETT PQ.�1H T-type Ca2� channel is the predominant subtype expressed inbovine and rat zona glomerulosa. Am J Physiol Cell Physiol 280:C265–C272, 2001.

359. SCHROEDER JE, FISCHBACH PS, AND MCCLESKEY EW. T-type calciumchannels: heterogeneous expression in rat sensory neurons andselective modulation by phorbol esters. J Neurosci 10: 947–951,1990.

360. SCHROEDER JE, FISCHBACH PS, ZHENG D, AND MCCLESKEY EW. Activa-tion of � opioid receptors inhibits transient high- and low-thresholdCa2� currents, but spares a sustained current. Neuron 6: 13–20,1991.

361. SCOTT RH, WOOTTON JF, AND DOLPHIN AC. Modulation of neuronalT-type calcium channel currents by photoactivation of intracellularguanosine 5-O-(3-thio)triphosphate. Neuroscience 38: 285–294,1990.

362. SEN L AND SMITH TW. T-type Ca2� channels are abnormal in genet-ically determined cardiomyopathic hamster hearts. Circ Res 75:149–155, 1994.

363. SERRANO JR, DASHTI SR, PEREZ-REYES E, AND JONES SW. Mg2� blockunmasks Ca2�/Ba2� selectivity of �1G T-type calcium channels.Biophys J 79: 3052–3062, 2000.

364. SERRANO JR, PEREZ-REYES E, AND JONES SW. State-dependent inacti-vation of the �1G T-type calcium channel. J Gen Physiol 114:185–201, 1999.

365. SHAH MJ, MEIS S, MUNSCH T, AND PAPE HC. Modulation by extracel-lular pH of low- and high-voltage-activated calcium currents of ratthalamic relay neurons. J Neurophysiol 85: 1051–1058, 2001.

366. SHARP AH, BLACK JL III, DUBEL SJ, SUNDARRAJ S, SHEN JP, YUNKER

AM, COPELAND TD, AND MCENERY MW. Biochemical and anatomicalevidence for specialized voltage-dependent calcium channel � iso-form expression in the epileptic and ataxic mouse, stargazer. Neu-

roscience 105: 599–617, 2001.367. SHOROFSKY SR AND JANUARY CT. L- and T-type Ca2� channels in

canine cardiac Purkinje cells. Single-channel demonstration of L-type Ca2� window current. Circ Res 70: 456–464, 1992.

368. SHUBA YM, TESLENKO VI, SAVCHENKO AN, AND POGORELAYA NH. Theeffect of permeant ions on single calcium channel activation inmouse neuroblastoma cells: ion-channel interaction. J Physiol 443:25–44, 1991.

369. SIDACH SS AND MINTZ IM. Kurtoxin, a gating modifier of neuronalhigh- and low-threshold Ca channels. J Neurosci 22: 2023–2034,2002.

370. SIPIDO KR, CARMELIET E, AND VAN DE WERF F. T-type Ca2� current asa trigger for Ca2� release from the sarcoplasmic reticulum inguinea-pig ventricular myocytes. J Physiol 508: 439–451, 1998.

371. SIROIS JE, LEI Q, TALLEY EM, LYNCH C III, AND BAYLISS DA. TheTASK-1 two-pore domain K� channel is a molecular substrate forneuronal effects of inhalation anesthetics. J Neurosci 20: 6347–6354, 2000.

372. SMIRNOV SV AND AARONSON PI. Ca2� currents in single myocytesfrom human mesenteric arteries: evidence for a physiological roleof L-type channels. J Physiol 457: 455–475, 1992.

373. SMIRNOV SV, ZHOLOS AV, AND SHUBA MF. Potential-dependent inwardcurrents in single isolated smooth muscle cells of the rat ileum.J Physiol 454: 549–571, 1992.

374. SNUTCH TP, LEONARD JP, GILBERT MM, LESTER HA, AND DAVIDSON N.Rat brain expresses a heterogeneous family of calcium channels.Proc Natl Acad Sci USA 87: 3391–3395, 1990.

375. SON WY, LEE JH, AND HAN CT. Acrosome reaction of human sper-matozoa is mainly mediated by �1H T-type calcium channels. Mol

Hum Reprod 6: 893–897, 2000.376. SOONG TW, STEA A, HODSON CD, DUBEL SJ, VINCENT SR, AND SNUTCH

TP. Structure and functional expression of a member of the lowvoltage-activated calcium channel family. Science 260: 1133–1136,1993.

377. STERIADE M AND LLINAS RR. The functional states of the thalamusand the associated neuronal interplay. Physiol Rev 68: 649–742,1988.

T-TYPE CALCIUM CHANNELS 159

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 44: Canal Calcio T

378. STERIADE M, MCCORMICK DA, AND SEJNOWSKI TJ. Thalamocorticaloscillations in the sleeping and aroused brain. Science 262: 679–685, 1993.

379. STUTZIN A, STOJILKOVIC SS, CATT KJ, AND ROJAS E. Characteristics oftwo types of calcium channels in rat pituitary gonadotrophs. Am J

Physiol Cell Physiol 257: C865–C874, 1989.380. SU H, SOCHIVKO D, BECKER A, CHEN J, JIANG Y, YAARI Y, AND BECK H.

Upregulation of a T-type Ca2� channel causes a long-lasting mod-ification of neuronal firing mode after status epilepticus. J Neurosci

22: 3645–3655, 2002.381. SUH-KIM H, WEI X, KLOS A, PAN S, RUTH P, FLOCKERZI V, HOFMANN F,

PEREZ-REYES E, AND BIRNBAUMER L. Reconstitution of the skeletalmuscle dihydropyridine receptor. Functional interaction among��� and �� subunits. Receptors Channels 4: 217–225, 1996.

382. SUI GP, WU C, AND FRY CH. Inward calcium currents in cultured andfreshly isolated detrusor muscle cells: evidence of a T-type calciumcurrent. J Urol 165: 621–626, 2001.

383. SUN QQ AND DALE N. Serotonergic inhibition of the T-type and highvoltage-activated Ca2� currents in the primary sensory neurons ofXenopus larvae. J Neurosci 17: 6839–6849, 1997.

384. SUN QQ AND DALE N. G-proteins are involved in 5-HT receptor-mediated modulation of N- and P/Q- but not T-type Ca2� channels.J Neurosci 19: 890–899, 1999.

385. SUNDGREN-ANDERSSON AK AND JOHANSSON S. Calcium spikes andcalcium currents in neurons from the medial preoptic nucleus ofrat. Brain Res 783: 194–209, 1998.

386. SUZUKI S, KAWAKAMI K, NISHIMURA S, WATANABE Y, YAGI K, SEINO M,AND MIYAMOTO K. Zonisamide blocks T-type calcium channel incultured neurons of rat cerebral cortex. Epilepsy Res 12: 21–27,1992.

387. SUZUKI S AND ROGAWSKI MA. T-type calcium channels mediate thetransition between tonic and phasic firing in thalamic neurons.Proc Natl Acad Sci USA 86: 7228–7232, 1989.

388. SWYNGHEDAUW B. Molecular mechanisms of myocardial remodeling.Physiol Rev 79: 215–262, 1999.

389. TAKAHASHI K, UENO S, AND AKAIKE N. Kinetic properties of T-typeCa2� currents in isolated rat hippocampal CA1 pyramidal neurons.J Neurophysiol 65: 148–155, 1991.

390. TAKENOSHITA M AND STEINBACH JH. Halothane blocks low-voltage-activated calcium currents in rat sensory neurons. J Neurosci 11:1404–1412, 1991.

391. TALAVERA K, STAES M, JANSSENS A, KLUGBAUER N, DROOGMANS G,HOFMANN F, AND NILIUS B. Aspartate residues of the glu-glu-asp-asp(EEDD) pore locus control selectivity and permeation of the T-typeCa2� channel �1G. J Biol Chem 276: 45628–45635, 2001.

392. TALEB O, TROUSLARD J, DEMENEIX BA, AND FELTZ P. Characterizationof calcium and sodium currents in porcine pars intermedia cells.Neurosci Lett 66: 55–60, 1986.

393. TALLEY EM, CRIBBS LL, LEE JH, DAUD A, PEREZ-REYES E, AND BAYLISS

DA. Differential distribution of three members of a gene familyencoding low voltage-activated (T-type) calcium channels. J Neu-

rosci 19: 1895–1911, 1999.394. TALLEY EM, SOLORZANO G, DEPAULIS A, PEREZ-REYES E, AND BAYLISS

DA. Low-voltage-activated calcium channel subunit expression in agenetic model of absence epilepsy in the rat. Mol Brain Res 75:159–165, 2000.

395. TANABE T, BEAM KG, POWELL JA, AND NUMA S. Restoration of exci-tation-contraction coupling and slow calcium current in dysgenicmuscle by dihydropyridine receptor complementary DNA. Nature

336: 134–139, 1988.396. TANABE T, TAKESHIMA H, MIKAMI A, FLOCKERZI V, TAKAHASHI H, KAN-

GAWA K, KOJIMA M, MATSUO H, HIROSE T, AND NUMA S. Primarystructure of the receptor for calcium channel blockers from skel-etal muscle. Nature 328: 313–318, 1987.

397. TANG CM, PRESSER F, AND MORAD M. Amiloride selectively blocks thelow threshold (T) calcium channel. Science 240: 213–215, 1988.

398. TARASENKO AN, KOSTYUK PG, EREMIN AV, AND ISAEV DS. Two types oflow-voltage-activated Ca2� channels in neurones of rat laterodorsalthalamic nucleus. J Physiol 499: 77–86, 1997.

399. TASKER JG AND DUDEK FE. Electrophysiological properties of neu-rones in the region of the paraventricular nucleus in slices of rathypothalamus. J Physiol 434: 271–293, 1991.

400. THOMPSON SM AND WONG RK. Development of calcium current

subtypes in isolated rat hippocampal pyramidal cells. J Physiol

439: 671–689, 1991.401. TODOROVIC SM, JEVTOVIC-TODOROVIC V, MENNERICK S, PEREZ-REYES E,

AND ZORUMSKI CF. Cav3.2 channel is a molecular substrate forinhibition of T-type calcium currents in rat sensory neurons bynitrous oxide. Mol Pharmacol 60: 603–610, 2001.

402. TODOROVIC SM, JEVTOVIC-TODOROVIC V, MEYENBURG A, MENNERICK S,PEREZ-REYES E, ROMANO C, OLNEY JW, AND ZORUMSKI CF. Redoxmodulation of T-type calcium channels in rat peripheral nocicep-tors. Neuron 31: 75–85, 2001.

403. TODOROVIC SM AND LINGLE CJ. Pharmacological properties of T-typeCa2� current in adult rat sensory neurons: effects of anticonvulsantand anesthetic agents. J Neurophysiol 79: 240–252, 1998.

404. TODOROVIC SM, PEREZ-REYES E, AND LINGLE CJ. Anticonvulsants butnot general anesthetics have differential blocking effects on differ-ent T-type current variants. Mol Pharmacol 58: 98–108, 2000.

405. TOMBAUGH GC AND SOMJEN GG. Differential sensitivity to intracellu-lar pH among high- and low-threshold Ca2� currents in isolated ratCA1 neurons. J Neurophysiol 77: 639–653, 1997.

406. TOMIC M, KOSHIMIZU T, YUAN D, ANDRIC SA, ZIVADINOVIC D, AND

STOJILKOVIC SS. Characterization of a plasma membrane calciumoscillator in rat pituitary somatotrophs. J Biol Chem 274: 35693–35702, 1999.

407. TOSELLI M AND LUX HD. Opposing effects of acetylcholine on thetwo classes of voltage-dependent calcium channels in hippocampalneurons. Exp Suppl 57: 97–103, 1989.

408. TOYOTA M, HO C, OHE-TOYOTA M, BAYLIN SB, AND ISSA JP. Inactivationof CACNA1G, a T-type calcium channel gene, by aberrant methyl-ation of its 5 CpG island in human tumors. Cancer Res 59: 4535–4541, 1999.

409. TSAKIRIDOU E, BERTOLLINI L, DE CURTIS M, AVANZINI G, AND PAPE HC.Selective increase in T-type calcium conductance of reticular tha-lamic neurons in a rat model of absence epilepsy. J Neurosci 15:3110–3117, 1995.

410. TSENG GN AND BOYDEN PA. Multiple types of Ca2� currents in singlecanine Purkinje cells. Circ Res 65: 1735–1750, 1989.

411. TSENG GN AND BOYDEN PA. Different effects of intracellular Ca andprotein kinase C on cardiac T and L Ca currents. Am J Physiol

Heart Circ Physiol 261: H364–H379, 1991.412. TSIEN RW, CLOZEL JP, AND NARGEOT J. Low Voltage-activated T-type

Calcium Channels. Chester, UK: Adis International Limited, 1998,p. 394.

413. TWOMBLY DA, YOSHII M, AND NARAHASHI T. Mechanisms of calciumchannel block by phenytoin. J Pharmacol Exp Ther 246: 189–195,1988.

414. TYTGAT J, NILIUS B, AND CARMELIET E. Modulation of the T-typecardiac Ca channel by changes in proton concentration. J Gen

Physiol 96: 973–990, 1990.415. TYTGAT J, NILIUS B, VEREECKE J, AND CARMELIET E. The T-type Ca

channel in guinea-pig ventricular myocytes is insensitive to isopro-terenol. Pflugers Arch 411: 704–706, 1988.

416. UMEMIYA M AND BERGER AJ. Properties and function of low- andhigh-voltage-activated Ca2� channels in hypoglossal motoneurons.J Neurosci 14: 5652–5660, 1994.

417. UMEMIYA M AND BERGER AJ. Single-channel properties of four cal-cium channel types in rat motoneurons. J Neurosci 15: 2218–2224,1995.

418. VAJNA R, KLOCKNER U, PEREVERZEV A, WEIERGRABER M, CHEN XH,MILJANICH G, KLUGBAUER N, HESCHELER J, PEREZ-REYES E, AND SCHNEI-DER T. Functional coupling between “R-type” Ca2� channels andinsulin secretion in the insulinoma cell line INS-1. Eur J Biochem

268: 1066–1075, 2001.419. VERGNES M AND MARESCAUX C. Pathophysiological mechanisms un-

derlying genetic absence epilepsy in rats. In: Idiopathic General-

ized Epilepsies: Clinical, Experimental, and Genetic Aspects, ed-ited by Malafosse A, Genton P, Hirsch E, Marescaux C, Broglin D,and Bernasconi R. London: Libbey, 1994, p. 151–168.

420. VESELOVSKII NS AND FEDULOVA SA. Two types of calcium channels inthe somatic membrane of spinal ganglion neurons in the rat. Dokl

Akad Nauk SSSR 268: 747–750, 1983.421. VIANA F, BAYLISS DA, AND BERGER AJ. Calcium conductances and

their role in the firing behavior of neonatal rat hypoglossal mo-toneurons. J Neurophysiol 69: 2137–2149, 1993.

160 EDWARD PEREZ-REYES

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 45: Canal Calcio T

422. VILLAME J, MASSICOTTE J, JASMIN G, AND DUMONT L. Effects of mibe-fradil, a T- and L-type calcium channel blocker, on cardiac remod-eling in the UM-X7.1 cardiomyopathic hamster. Cardiovasc Drugs

Ther 15: 41–48, 2001.423. WAGNER C, KRAMER BK, HINDER M, KIENINGER M, AND KURTZ A. T-type

and L-type calcium channel blockers exert opposite effects onrenin secretion and renin gene expression in conscious rats. Br J

Pharmacol 124: 579–585, 1998.424. WALKER D AND DE WAARD M. Subunit interaction sites in voltage-

dependent Ca2� channels: role in channel function. Trends Neuro-

sci 21: 148–154, 1998.425. WAN X, DESILETS M, SOBOLOFF J, MORRIS C, AND TSANG BK. Musca-

rinic activation inhibits T-type Ca2� current in hen granulosa cells.Endocrinology 137: 2514–2521, 1996.

426. WANG L, BHATTACHARJEE A, FU J, AND LI M. Abnormally expressedlow-voltage-activated calcium channels in �-cells from NOD miceand a related clonal cell line. Diabetes 45: 1678–1683, 1996.

427. WANG L, BHATTACHARJEE A, ZUO Z, HU F, HONKANEN RE, BERGGREN

PO, AND LI M. A low voltage-activated Ca2� current mediates cyto-kine-induced pancreatic �-cell death. Endocrinology 140: 1200–1204, 1999.

428. WANG R, KARPINSKI E, AND PANG PK. Two types of calcium channelsin isolated smooth muscle cells from rat tail artery. Am J Physiol

Heart Circ Physiol 256: H1361–H1368, 1989.429. WANG X, MCKENZIE JS, AND KEMM RE. Whole cell calcium currents in

acutely isolated olfactory bulb output neurons of the rat. J Neuro-

physiol 75: 1138–1151, 1996.430. WARRE R AND RANDALL A. Modulation of the deactivation kinetics of

a recombinant rat T-type Ca2� channel by prior inactivation. Neu-

rosci Lett 293: 216–220, 2000.431. WARREN RA AND JONES EG. Maturation of neuronal form and func-

tion in a mouse thalamo-cortical circuit. J Neurosci 17: 277–295,1997.

432. WATANABE K, OCHIAI Y, WASHIZUKA T, INOMATA T, MIYAKITA Y, SHIBA M,IZUMI T, SHIBATA A, QU YL, AND NAGATOMO T. Clinical evaluation ofserum amlodipine level in patients with angina pectoris. Gen Phar-

macol 27: 205–209, 1996.433. WELKER HA, WILTSHIRE H, AND BULLINGHAM R. Clinical pharmacoki-

netics of mibefradil. Clin Pharmacokinet 35: 405–423, 1998.434. WEN JF, CUI X, AHN JS, KIM SH, SEUL KH, KIM SZ, PARK YK, LEE HS,

AND CHO KW. Distinct roles for L- and T-type Ca2� channels inregulation of atrial ANP release. Am J Physiol Heart Circ Physiol

279: H2879–H2888, 2000.435. WENNEMUTH G, WESTENBROEK RE, XU T, HILLE B, AND BABCOCK DF.

Cav2.2 and Cav2.3 (N- and R-type) Ca2� channels in depolarization-evoked entry of Ca2� into mouse sperm. J Biol Chem 275: 21210–21217, 2000.

436. WHITE G, LOVINGER DM, AND WEIGHT FF. Transient low-thresholdCa2� current triggers burst firing through an afterdepolarizing po-tential in an adult mammalian neuron. Proc Natl Acad Sci USA 86:6802–6806, 1989.

437. WILCOX KS, GUTNICK MJ, AND CHRISTOPH GR. Electrophysiologicalproperties of neurons in the lateral habenula nucleus: an in vitrostudy. J Neurophysiol 59: 212–225, 1988.

438. WILLIAMS ME, WASHBURN MS, HANS M, URRUTIA A, BRUST PF, PRO-DANOVICH P, HARPOLD MM, AND STAUDERMAN KA. Structure and func-tional characterization of a novel human low-voltage activatedcalcium channel. J Neurochem 72: 791–799, 1999.

439. WILLIAMS PJ, MACVICAR BA, AND PITTMAN QJ. Electrophysiologicalproperties of neuroendocrine cells of the intact rat pars intermedia:multiple calcium currents. J Neurosci 10: 748–756, 1990.

440. WILLIAMS S, SERAFIN M, MUHLETHALER M, AND BERNHEIM L. Distinctcontributions of high- and low-voltage-activated calcium currentsto afterhyperpolarizations in cholinergic nucleus basalis neurons ofthe guinea pig. J Neurosci 17: 7307–7315, 1997.

441. WILLIAMS SR, TOTH TI, TURNER JP, HUGHES SW, AND CRUNELLI V. The“window” component of the low threshold Ca2� current producesinput signal amplification and bistability in cat and rat thalamocor-tical neurones. J Physiol 505: 689–705, 1997.

442. WILLIAMSON AV AND SATHER WA. Nonglutamate pore residues in ion

selection and conduction in voltage-gated Ca2� channels. Biophys

J 77: 2575–2589, 1999.443. WOLFE JT, WANG H, PEREZ-REYES E, AND BARRETT PQ. Stimulation of

recombinant Cav3.2, T-type, Ca2� channel currents by CaMKII�c.J Physiol 538: 343–355, 2002.

444. WU JY AND LIPSIUS SL. Effects of extracellular Mg2� on T- and L-typeCa2� currents in single atrial myocytes. Am J Physiol Heart Circ

Physiol 259: H1842–H1850, 1990.445. XIONG Z, SPERELAKIS N, NOFFSINGER A, AND FENOGLIO-PREISER C.

Changes in calcium channel current densities in rat colonic smoothmuscle cells during development and aging. Am J Physiol Cell

Physiol 265: C617–C625, 1993.446. XU X AND BEST PM. Postnatal changes in T-type calcium current

density in rat atrial myocytes. J Physiol 454: 657–672, 1992.447. XU XP AND BEST PM. Increase in T-type calcium current in atrial

myocytes from adult rats with growth hormone-secreting tumors.Proc Natl Acad Sci USA 87: 4655–4659, 1990.

448. YAMAKAGE M, CHEN X, TSUJIGUCHI N, KAMADA Y, AND NAMIKI A. Dif-ferent inhibitory effects of volatile anesthetics on T- and L-typevoltage-dependent Ca2� channels in porcine tracheal and bronchialsmooth muscles. Anesthesiology 94: 683–693, 2001.

449. YANG J, ELLINOR PT, SATHER WA, ZHANG JF, AND TSIEN RW. Moleculardeterminants of Ca2� selectivity and ion permeation in L-type Ca2�

channels. Nature 366: 158–161, 1993.450. YATANI A, SEIDEL CL, ALLEN J, AND BROWN AM. Whole-cell and

single-channel calcium currents of isolated smooth muscle cellsfrom saphenous vein. Circ Res 60: 523–533, 1987.

451. YEE HF JR, WEISS JN, AND LANGER GA. Neuraminidase selectivelyenhances transient Ca2� current in cardiac myocytes. Am J Physiol

Cell Physiol 256: C1267–C1272, 1989.452. YEOMAN MS, BREZDEN BL, AND BENJAMIN PR. LVA and HVA Ca2�

currents in ventricular muscle cells of the Lymnaea heart. J Neu-

rophysiol 82: 2428–2440, 1999.453. YOUNG RC, SMITH LH, AND MCLAREN MD. T-type and L-type calcium

currents in freshly dispersed human uterine smooth muscle cells.Am J Obstet Gynecol 169: 785–792, 1993.

454. YUSAF SP, GOODMAN J, PINNOCK RD, DIXON AK, AND LEE K. Expres-sion of voltage-gated calcium channel subunits in rat dorsal rootganglion neurons. Neurosci Lett 311: 137–141, 2001.

455. ZAMPONI GW, BOURINET E, AND SNUTCH TP. Nickel block of a familyof neuronal calcium channels: subtype- and subunit-dependentaction at multiple sites. J Membr Biol 151: 77–90, 1996.

456. ZHAN XJ, COX CL, RINZEL J, AND SHERMAN SM. Current clamp andmodeling studies of low-threshold calcium spikes in cells of thecat’s lateral geniculate nucleus. J Neurophysiol 81: 2360–2373,1999.

457. ZHAN XJ, COX CL, AND SHERMAN SM. Dendritic depolarization effi-ciently attenuates low-threshold calcium spikes in thalamic relaycells. J Neurosci 20: 3909–3914, 2000.

458. ZHANG L, VALIANTE TA, AND CARLEN PL. Contribution of the low-threshold T-type calcium current in generating the post-spike de-polarizing afterpotential in dentate granule neurons of immaturerats. J Neurophysiol 70: 223–231, 1993.

459. ZHANG Y, CRIBBS LL, AND SATIN J. Arachidonic acid modulation of�1H, a cloned human T-type calcium channel. Am J Physiol Heart

Circ Physiol 278: H184–H193, 2000.460. ZHOU Q, GODWIN DW, O’MALLEY DM, AND ADAMS PR. Visualization of

calcium influx through channels that shape the burst and tonicfiring modes of thalamic relay cells. J Neurophysiol 77: 2816–2825,1997.

461. ZHOU Z AND JANUARY CT. Both T- and L-type Ca2� channels cancontribute to excitation-contraction coupling in cardiac Purkinjecells. Biophys J 74: 1830–1839, 1998.

462. ZHOU Z AND LIPSIUS SL. T-type calcium current in latent pacemakercells isolated from cat right atrium. J Mol Cell Cardiol 26: 1211–1219, 1994.

463. ZHUANG H, BHATTACHARJEE A, HU F, ZHANG M, GOSWAMI T, WANG L,WU S, BERGGREN PO, AND LI M. Cloning of a T-type Ca2� channelisoform in insulin-secreting cells. Diabetes 49: 59–64, 2000.

T-TYPE CALCIUM CHANNELS 161

Physiol Rev • VOL 83 • JANUARY 2003 • www.prv.org

on October 7, 2011

physrev.physiology.orgD

ownloaded from

Page 46: Canal Calcio T

doi:10.1152/physrev.00018.2002 83:117-161, 2003.Physiol RevEdward Perez-ReyesT-type Calcium ChannelsMolecular Physiology of Low-Voltage-Activated

You might find this additional info useful...

457 articles, 286 of which can be accessed free at:This article cites http://physrev.physiology.org/content/83/1/117.full.html#ref-list-1

100 other HighWire hosted articles, the first 5 are:This article has been cited by

  [PDF] [Full Text] [Abstract]

, August 24, 2011; 31 (34): 12218-12228.J. Neurosci.Gabe J. Murphy and Fred RiekeAbsolute Sensitivity of Parallel Retinal CircuitsElectrical Synaptic Input to Ganglion Cells Underlies Differences in the Output and 

[PDF] [Full Text] [Abstract], August 31, 2011; 31 (35): 12566-12578.J. Neurosci.

Richard A. Felix II, Anders Fridberger, Sara Leijon, Albert S. Berrebi and Anna K. MagnussonParaolivary NucleusSound Rhythms Are Encoded by Postinhibitory Rebound Spiking in the Superior 

[PDF] [Full Text] [Abstract], September , 2011; 3 (9): .Cold Spring Harb Perspect Biol

David C. Hill-Eubanks, Matthias E. Werner, Thomas J. Heppner and Mark T. NelsonCalcium Signaling in Smooth Muscle 

[PDF] [Full Text] [Abstract], September 28, 2011; 31 (39): 13991-14004.J. Neurosci.

Shunbing Zhao, Amir Farzad Sheibanie, Myongkeun Oh, Pascale Rabbah and Farzan NadimPeptide Neuromodulation of Synaptic Dynamics in an Oscillatory Network 

[PDF] [Full Text] [Abstract], October 13, 2011; 369 (1952): 3820-3839.Phil. Trans. R. Soc. A

Vincenzo Crunelli, Adam C. Errington, Stuart W. Hughes and Tibor I. Tóthdynamics of the slow (<1 Hz) sleep oscillation in thalamocortical networks

potential: a key determinant of the local and global2+The thalamic low-threshold Ca

including high resolution figures, can be found at:Updated information and services http://physrev.physiology.org/content/83/1/117.full.html

can be found at:Physiological Reviewsabout Additional material and information http://www.the-aps.org/publications/prv

This infomation is current as of October 7, 2011. 

website at http://www.the-aps.org/.MD 20814-3991. Copyright © 2003 by the American Physiological Society. ISSN: 0031-9333, ESSN: 1522-1210. Visit ourpublished quarterly in January, April, July, and October by the American Physiological Society, 9650 Rockville Pike, Bethesda

provides state of the art coverage of timely issues in the physiological and biomedical sciences. It isPhysiological Reviews

on October 7, 2011

physrev.physiology.orgD

ownloaded from