In Vitro Characterization of T-Type Calcium Channel ...

2 downloads 0 Views 846KB Size Report
Jul 1, 2011 - Steven V. Fox, Scott M. Doran, James C. Barrow, Zhi-Qiang Yang, Thomas S. ...... Luján R, Emmenegger Y, Watanabe M, Bond CT, Franken.
0022-3565/10/3352-409–417$20.00 THE JOURNAL OF PHARMACOLOGY AND EXPERIMENTAL THERAPEUTICS Copyright © 2010 by The American Society for Pharmacology and Experimental Therapeutics JPET 335:409–417, 2010

Vol. 335, No. 2 171058/3632190 Printed in U.S.A.

In Vitro Characterization of T-Type Calcium Channel Antagonist TTA-A2 and In Vivo Effects on Arousal in Mice Richard L. Kraus, Yuxing Li, Yun Gregan, Anthony L. Gotter, Victor N. Uebele, Steven V. Fox, Scott M. Doran, James C. Barrow, Zhi-Qiang Yang, Thomas S. Reger, Kenneth S. Koblan, and John J. Renger Departments of Neuroscience (R.L.K., Y.L., Y.G., A.L.G., V.N.U., S.V.F., S.M.D., K.S.K., J.J.R.) and Medicinal Chemistry (J.C.B., Z-Q.Y., T.S.R.), Merck Research Laboratories, West Point, Pennsylvania Received June 4, 2010; accepted August 3, 2010

Introduction Calcium, the most common signal transduction element in cells, enters the cytosol through highly selective voltagegated calcium channels. Electrophysiological studies have identified two major classes of calcium channels, high-voltage-activated (L-type: Cav1.1, Cav1.2, Cav1.3, Cav1.4; P/Qtype: Cav2.1; N-type: Cav2.2; R-type: Cav2.3) and low-voltage-activated (T-type: Cav3) calcium channels. In contrast to high-voltage-activated calcium channels, T-type channels were defined by activation at depolarizations near the resting membrane potential, fast gating kinetics, slow deactivation kinetics, and small single channels conductance (Catterall et Article, publication date, and citation information can be found at http://jpet.aspetjournals.org. doi:10.1124/jpet.110.171058.

inactivated channels. The compound showed a ⬃300-fold selectivity for Cav3 channels over high-voltage activated calcium channels. Inhibitory effects on native T-type currents were confirmed in brain slice recordings from the dorsal lateral geniculate nucleus and the subthalamic nucleus. Furthermore, we demonstrate that in vivo T-type channel inhibition by TTA-A2 suppresses active wake and promotes slow-wave sleep in wild-type mice but not in mice lacking both Cav3.1 and Cav3.3, suggesting the selective effect of TTA-A2 on recurrent thalamocortical network activity. The discovery of the potent and selective T-type channel antagonist TTA-A2 has enabled us to study the in vivo effects of pharmacological T-channel inhibition on arousal in mice, and it will help to explore the validity of these channels as potential drug targets for sleep-related and other neurological diseases.

al., 2005). To date, three T-type channels, Cav3.1, Cav3.2, and Cav3.3, have been cloned in mice, rats, and humans, showing all the properties of native neuronal T-type calcium channels when expressed heterologously (Perez-Reyes, 2003). Low-voltage-activated T-type calcium channels are expressed throughout the body. They have been implicated in many physiological processes as diverse as hormone secretion, smooth muscle contraction, fertilization, cell growth and proliferation, pacemaker activity, pain processing (PerezReyes, 2003; Zamponi et al., 2009), low-threshold spikes (LTS), and thalamocortical rhythms during sleep and epilepsy (Steriade et al., 1993; McCormick and Bal, 1997; Crunelli et al., 2006; Beenhakker and Huguenard, 2009). Calcium-dependent low-threshold currents mediated by Ttype calcium channels underlie low-threshold spikes that have been demonstrated as intrinsic firing properties of tha-

ABBREVIATIONS: TTA-A2, 2-(4-cyclopropylphenyl)-N-((1R)-1-{5-[(2,2,2-trifluoroethyl)oxo]-pyridin-2-yl}ethyl)acetamide; LTS, low-threshold spikes; EEG, electroencephalogram; EMG, electromyogram; ECoG, electrocorticogram; REM; rapid eye movement; dLGN, dorsal lateral geniculate nucleus; STN, subthalamic nucleus; DMSO, dimethyl sulfoxide; CRS, choline replacement solution; aCSF, artificial cerebrospinal fluid; HEK, human embryonic kidney; CI, confidence interval. 409

Downloaded from jpet.aspetjournals.org at Merck & Co., Inc. on July 1, 2011

ABSTRACT T-type calcium channels have been implicated in many behaviorally important neurophysiological processes, and altered channel activity has been linked to the pathophysiology of neurological disorders such as insomnia, epilepsy, Parkinson’s disease, depression, schizophrenia, and pain. We have previously identified a number of potent and selective T-type channel antagonists (Barrow et al., 2007; Shipe et al., 2008; Yang et al., 2008). Here we describe the properties of the antagonist TTA-A2 [2-(4-cyclopropylphenyl)-N-((1R)-1-{5-[(2,2,2-trifluoroethyl)oxo]-pyridin-2-yl}ethyl)acetamide], assessed in patchclamp experiments. TTA-A2 blocks T-type channels (Cav3.1, 3.2, 3.3) voltage dependently and with high potency (IC50 ⬃100 nM). Stimulation at 3 Hz revealed additional use dependence of inhibition. A hyperpolarized shift of the channel availability curve and delayed channel recovery from inactivation suggest that the compound preferentially interacts with and stabilizes

410

Kraus et al.

Materials and Methods Cell Culture Low Voltage-Activated T-Type Calcium Channels. For whole-cell voltage-clamp recordings, cell lines expressing human Cav3.1, Cav3.2, or Cav3.3 channels were used. Tetracycline-inducible HEK-293 cells that stably express the tetracycline repressor protein were transfected with Cav3.1, Cav3.2, and Cav3.3 in pcDNA 4/TO. Cells were maintained in DMEM (Invitrogen, Carlsbad, CA) and supplemented with 10% fetal bovine serum (Clontech, Mountain View, CA) 2 mM L-glutamine, 100 units/ml penicillin/streptomycin (Invitrogen), 150 ␮g/ml Zeocin (phleomycin; Invitrogen), and 5 ␮g/ml blasticidin (Invitrogen). Cells were plated on poly-D-lysine-coated cover slips, and Cav3 channel expression was induced by exposure to 0.1 mM tetracycline (Sigma-Aldrich, St. Louis, MO) for at least 24 h before recording. Cells were then washed with mammalian Ringer’s solution immediately before recording. High Voltage-Activated Calcium Channels. L-type currents (Cav1.2) were recorded from HEK-293 cell coexpressing ␣1C, ␤2a, ␣2␦ and Kir2.3 (Xia et al., 2004). Cells were grown in DMEM (Invitrogen) and supplemented with 10% fetal calf serum (Sigma F-2442), 40 ␮g/ml phleomycin, 100 ␮g/ml hygromycin (Invitrogen), and 100 ␮g/ml G-418 (Invitrogen). P-type currents (Cav2.1) were recorded from HEK-293 cells coexpressing ␣1A, ␤3, ␣2␦, and Kir2.3. Cells were grown in DMEM, 10% fetal bovine serum, penicillin/streptomycin, Geneticin (G-418; Invitrogen), phleomycin, and 6 ␮g/ml blasticidin. N-type currents (Cav2.2) were recorded from HEK-293 cells coexpressing ␣1B, ␤3, ␣2␦, and Kir2.3. Cells were grown in DMEM, 10% fetal bovine serum, penicillin/streptomycin, G-418, phleomycin, and 10 ␮g/ml blasticidin. R-type currents (Cav2.3) were recorded from HEK-293 cells coexpressing ␣1E, ␤3, ␣2␦, and Kir2.3. Cells were grown in DMEM, 10% fetal bovine serum, penicillin/streptomycin, G-418, and phleomycin. All cell lines were maintained at 5% CO2 at 37°C (Dai et al., 2008). Cav2.1 expressing cell line was transferred to 30°C/5% CO2 for 48 h after cells achieved 80% confluence. Cells were plated on poly-D-lysine-coated cover slips for at least 24 h before recording.

Recombinant Cell Electrophysiology Single-cell recordings of currents through low-voltage-activated CaV3.1, Cav3.2, and Cav3.3 calcium channels and high-voltage-activated CaV1.2, CaV2.1, CaV2.2, and CaV2.3 calcium channels were performed at room temperature (20 –22°C) using the whole-cell patch-clamp technique. Currents were recorded using a HEKA EPC-9 patch-clamp amplifier. Data were stored on a personal computer equipped with HEKA Pulse 8.5 and analyzed using Pulsefit and Fitmaster (HEKA, Lambrecht/Pfalz, Germany), Igor Pro 4.0 (Wavemetrics, Lake Oswego, OR), Origin 6.0 (MicroCal LLC, Northampton, MA), and Prism 4 (GraphPad Software Inc., San Diego, CA). Patch pipettes were made from borosilicate glass tubing (World Precision Instruments, Inc., Sarasota, FL), fire polished, and had a resistance of 1 to 4.5 M⍀ when filled with internal solution measured in the recording medium. Voltage errors were minimized by series resistance compensation (typically 70%). Leak resistance and capacitance were corrected by subtracting the scaled current observed with a P/4 protocol. Data were acquired at 10 to 50 kHz and filtered at 5 to 10 kHz. Salts for recording solutions were purchased from Sigma-Aldrich. For T-type calcium current measurements, the following solutions were used. Pipette solution consisted of 125 mM CsCl, 10 mM TEACl4, 10 mM HEPES, 8 mM NaCl, 0.06 mM CaCl2, 0.6 mM EGTA, 4 mM ATP-Mg, 0.3 mM GTP; and 280 mOsm, pH 7.2 adjusted with CsOH. The bath solution consisted of 130 mM NaCl, 30 mM D-(⫹)-glucose, 20 mM HEPES, 4 mM KCl, 2 mM CaCl2, and 1 mM MgCl2, pH 7.4. For high-voltage-activated channel recordings, the internal solution used was 135 mM CsCl, 10 mM HEPES, 1 mM MgCl2, 10 mM EGTA, 5 mM ATP-Mg, and 0.1 mM GTP; pH 7.2. The bath solution was 150 mM Chol-Cl, 10 mM HEPES, 1 mM MgCl2, 15

Downloaded from jpet.aspetjournals.org at Merck & Co., Inc. on July 1, 2011

lamic and cortical neurons producing network oscillations and generating characteristic EEG patterns during wakefulness, non-REM sleep (delta slow-wave sleep), and REM sleep. LTS are predominantly observed in the thalamus during non-REM sleep and are generally absent during wakefulness and REM sleep, suggesting a key role of T-type calcium channels in the regulation of sleep (Steriade et al., 1993; McCormick and Bal, 1997; Crunelli et al., 2006). Abnormal neuronal hyperpolarization that results in aberrant thalamocortical oscillations that depend on T-type channel activity has been linked to the pathophysiology of a variety of neurological disorders. It is believed that dysfunction at the level of thalamic relay cells induces thalamocortical dysrhythmia linked to insomnia, absence epilepsy, Parkinson’s disease, tinnitus, depression, schizophrenia, and neurogenic pain (Huguenard and Prince, 1994; Huguenard, 1999; Llina´s et al., 1999). Consistently, data from Cav3 transgenic, gene knockout, and knockdown animal models suggest the involvement of T-type channels in epilepsy, pain signaling (Kim et al., 2001, 2003; Bourinet et al., 2005; Choi et al., 2007), and sleep regulation (Kim et al., 2001; Lee et al., 2004; Anderson et al., 2005). A recent study has also suggested a role of T-type channels in body weight maintenance (Uebele et al., 2009a). Many drugs in clinical use, including antipsychotic (penfluridol, clozapine, haloperidol, pimozide), antidepressants (fluoxetine, trazodone), and antiepileptics (ethosuximide, phenytoin, zonisamde), have been shown to have activity on T-type channels (Heady et al., 2001; Santi et al., 2002; Traboulsie et al., 2006; Kraus et al., 2007). Although most of those molecules are neither selective nor potent on T-type channels, their activity provides evidence for the potential of T-type channels as therapeutic targets (Shin et al., 2008). Mibefradil (Posicor), the first T-type calcium channel antagonist identified, was introduced to the clinic as an antihypertensive, and later withdrawn from the U.S. market because of interactions with liver enzymes (Van der Vring et al., 1999). Despite its activity on other highvoltage-activated ion channels, mibefradil is frequently used as a pharmacological research tool to study T-type channels (Mehrke et al., 1994; Heady et al., 2001; Leuranguer et al., 2001). To more convincingly understand the true role of T-type channels, potent and selective agents are needed to study T-type channel function in neurological disease states and to validate T-type channels as efficacious drug targets. Consequently, numerous efforts to identify such T-type channel antagonists have been initiated in several laboratories. The discovery of a number of potent and selective T-type calcium channel antagonists has recently been reported (Barrow et al., 2007; Shipe et al., 2008; Yang et al., 2008; Uebele et al., 2009b). Here, we describe the activity of the pyridyl amide T-type calcium channel antagonist TTA-A2 ([2-(4-cyclopropylphenyl)-N-((1R)-1-{5-[(2,2,2-trifluoroethyl)oxo]pyridin-2-yl}ethyl) acetamide] on recombinant human T-type calcium channels expressed in HEK-293 cells, including its potency, voltage and use dependence, and effects on channel gating. In addition, we confirmed compound activity on native T-type currents from thalamic relay neurons in the dorsal lateral geniculate nucleus (dLGN) and the subthalamic nucleus (STN), and demonstrate T-type channel-selective pharmacodynamic effects of TTA-A2 in vivo.

T-Type Calcium Channel Inhibition by TTA-A2 mM BaCl2, 5 mM TEA-OH; pH 7.2. TTA-A2 was dissolved in DMSO, stored at ⫺20°C, and diluted in the bath solutions.

Brain Slice Preparation

Brain Slice Recordings Whole-cell patch clamp recordings were performed from thalamic neurons in 350 ␮m-thick midbrain slices at a temperature of 32°C. During recordings, slices were maintained fully submerged and continuously perfused (2–3 ml/min) with 95% O2/5% CO2 equilibrated aCSF. Cells in the subthalamic nucleus and the dorsal lateral geniculate nucleus were visualized using a differential interference contrast microscope and an infrared video system. Salts for recording solutions and tetrodotoxin were purchased from Sigma-Aldrich. Recording electrodes were pulled from borosilicate glass on a two-stage puller and had resistance in the range of 4 to 7 M⍀ when filled with Cs methanesulfonate internal solution containing 140 mM Cs methanesulfonate, 15 mM NaF, 4 mM MgCl2, 40 mM HEPES, 0.1 mM Cs BAPTA, 1 mM QX-314, 2 mM Mg ATP, 0.2 mM Na GTP, 12 mM phosphocreatine, 0.1 mM leupeptin. All recordings were done using HEKA EPC 10 patch clamp amplifiers. Recordings were sampled at 20 kHz and filtered at 6.67 kHz.

Electrophysiology Analysis Concentration-response curves were fitted using a sigmoidal Hill function, I ⫽ 1/(1 ⫹ exp[EC50 ⫺ Vm]/Hill slope). 95% Confidence intervals are listed for IC50 values and Hill coefficients. Channel activation and inactivation curves resulting from the voltage command protocols were plotted to the Boltzmann equation: I/Imax ⫽ 1/ (1 ⫹ exp[(V0.5 ⫺ V)/k]) to derive half-maximal voltage (V0.5) and slope factor (k) for comparison. Time course of recovery from inactivation was best fitted by a single exponential decay-yielding time constant (␶) in the absence and presence of drug. Student’s t test, ANOVA one-way post hoc Dunnett’s test, was used for statistical analysis, and values were considered significant at p ⬍ 0.05. Data are given as mean ⫾ S.E.M.

homozygous mice, as well as wild-type animals. Before the study, animals were acclimated to a 12:12-light/dark schedule (lights on at 2:00 PM, lights off at 2:00 AM). All animal procedures were conducted in accordance with Merck Institutional Animal Care and Use Committee. Double Cav3.1/Cav3.3 knockout (n ⫽ 6) and wild-type (n ⫽ 6) mice were surgically implanted with TL11M2-F20-EET radio telemetry transmitters (Data Sciences International [DSI], St. Paul, MN). Electrocorticogram (ECoG) electrode placement includes one electrode positioned 1 mm anterior to bregma, 2 mm lateral of the central suture, and a second at 1 mm anterior to lambda and 2.5 mm lateral fixed in predrilled holes with stainless steel screws. Electromyogram (EMG) electrodes were placed below the trapezius muscle. Radiotelemetry ECoG and EMG signals were acquired at a 500 Hz rate from RPC-1 receivers using DataQuest A.R.T. 3.1 Acquisition software (DSI). Automated raw data upload was performed daily to a remote server, and sleep scoring was performed using mouse scoring parameters on the Somnologica Data analysis package (Embla, ON, Canada). Surgical procedures and the accuracy of automated sleep scoring parameters have been confirmed by hand scoring of ECoG and EMG traces. The effects of 10 mg/kg TTA-A2 was assessed in both wild-type and double Cav3.1/Cav3.3 knockout animals in a 5-day crossover dosing paradigm in which animals received a once-daily treatment 1 h before lights on (ZT 23:00; 13:00 DST). Daily vehicle and compound treatments occurred for the first 5 days, followed by 2 days of washout and subsequently a reversal of the vehicle and compound treatment groups, such that all animals ultimately received both vehicle and compound. The time spent in active wake, REM, and delta (slow-wave) sleep for each 30-min period from each animal for all 5 days of the treatment arm was averaged according to the treatment condition and was plotted over a single 24-h period. Significant differences between vehicle and compound treatment groups at each 30-min interval were determined using the linear effects model for repeated measures.

Results The mechanism of T-type calcium channel block by TTA-A2 (chemical structure shown in Fig. 1) was studied in wholecell patch clamp experiments using three stable HEK-293 cell lines expressing human Cav3.1, Cav3.2, or Cav3.3 channels. First we assessed Cav3.1 channel activity (as previously reported in Uebele et al., 2009a). To evoke maximal Cav3.1 currents, 30-ms test pulses to the peak of the I/V curve, ⫺20 mV, were given every 10 s from a holding potential of ⫺100 mV. TTA-A2 caused a concentration-dependent decrease of the measured Cav3.1 peak current amplitude. Inhibition was reversible upon washout. Normalized steady-state current amplitudes in the presence of drug were plotted against drug concentrations, and the resulting curve was fitted to the Hill equation. Under these experimental conditions, the concen-

F

TTA-A2 N N

Mouse Polysomnography Mice homozygous for targeted mutation of both Cacna1g (Cav3.1) and Cacna1i (Cav3.3) genes along with wild-type controls were generated on a C57BL6/Sv129 background and bred identically. Heterozygous Cacna1g and Cacna1i mutant lines obtained from Deltagen were intercrossed at Taconic Farms to generate double mutant

F F

O

O

Fig. 1. Chemical structure of TTA-A2.

Downloaded from jpet.aspetjournals.org at Merck & Co., Inc. on July 1, 2011

All animals used in these studies were cared for in accordance with the National Institutes of Health Guide for the Care and Use of Laboratory Animals. The Merck Research Laboratories Institutional Animal Care and Use Committee approved all studies described in this manuscript, and experimental protocols were in accordance with all applicable guidelines regarding the care and use of animals. Animals were housed in an Association for Assessment and Accreditation of Laboratory Animal Care International-approved facility with free access to food and water. Electrophysiology experiments were performed on normal sagittal slices from 15- to 20-day-old Sprague-Dawley rats (Taconic Farms, Germantown, NY). Animals were decapitated and brains were rapidly removed and submerged in an ice-cold choline replacement solution (CRS) containing 126 mM choline chloride, 5 mM KCl2, 1.2 mM NaH2PO4, 1.3 mM MgCl2, 8 mM MgSO4, 10 mM glucose, and 26 mM NaHCO3, equilibrated with 95% O2/5% CO2. Slices 350 ␮m-thick were obtained using a vibrating blade microtome (Leica Microsystems, Inc., Deerfield, IL). Slices were immediately transferred to a holding chamber containing normal artificial cerebrospinal fluid (aCSF); 124 mM NaCl, 2.5 mM KCl, 1.3 mM MgSO4, 1.0 mM NaH2PO4, 2 mM CaCl2, 20 mM glucose, and 26 mM NaHCO3, equilibrated with 95% O2/5% CO2 that was maintained at room temperature. In all experiments, 5 ␮M glutathione, 500 ␮M pyruvate, and 250 ␮M kynurenic acid were included in the choline chloride buffer and in the holding chamber aCSF.

411

412

Kraus et al.

tration required for 50% current inhibition (IC50) at ⫺100 mV was 4.1 ␮M (Fig. 2B). According to the modulated receptor hypothesis, drugs can

B

-100 mV

10 3

0.4 nA

1 M C

-80 mV 300

0.2 nA

100

Remaining Current (%)

A

30 nM C

have varying affinities for channels in the closed/resting, open, and inactivated state (Hille, 1977; Hondeghem and Katzung, 1984). Based on T-type calcium channel availabil-

100 75

IC 50 = 4.1 M

50 25 0

IC 50 = 89 nM Cav3.1 -9

-8

-7

-6

-5

-4

TTA-A2 [M]

-100 mV

D

10 3 1 M C

0.4 nA

-80 mV

300 100

Remaining Current (%)

100

0.2 nA

75 50 25 0

30 nM

3 1 M

0.4 nA

C

-80 mV 300 100

0.4 nA 10 ms

Remaining Current (%)

10

C

Cav3.2 -9

-8

-7

-6

-5

-4

TTA-A2 [M]

F

-100 mV

30 nM

IC 50 = 92 nM

10 ms

C

E

IC 50 = 5.6 M

100

IC 50 = 3.7 M

75 50

IC 50 = 98nM

25 0

Cav3.3 -9

-8

-7

-6

-5

-4

TTA-A2 [M] Fig. 2. Inhibition of Cav3 T-type calcium channels by TTA-A2. A, top, Cav3.1 current traces evoked by 30 ms depolarizations steps from ⫺100 to ⫺20 mV under control conditions and at steady-state in the presence of 1, 3, and 10 ␮M TTA-A2. A, bottom, Cav3.1 current traces evoked by 30-ms depolarizations steps from ⫺80 to ⫺20 mV under control conditions and at steady-state in the presence of 30, 100, and 300 nM TTA-A2 (Uebele et al., 2009a). B, concentration-response curve of TTA-A2 for Cav3.1 channels at holding potentials of ⫺100 and ⫺80 mV. Data points reflect mean ⫾ S.E. of three to four determinations. Error bars are hidden behind symbols. IC50 values of 4.1 ␮M (95% CI 3.7 to 4.4 ␮M) and 89 nM (95% CI 77 to 102 nM) and Hill coefficients of 1.2 (95% CI 1.3 to 1.1) and 1.4 (95% CI 1.7 to 1.1) were determined at ⫺100 and ⫺80 mV, respectively, using the equation of a sigmoidal dose-response curve (Uebele et al., 2009a). C, top, Cav3.2 current traces evoked by 30-ms depolarizations steps from ⫺100 to ⫺20 mV under control conditions and in the presence of 1, 3, and 10 ␮M TTA-A2. C, bottom, Cav3.2 current traces evoked by 30-ms depolarizations steps from ⫺80 to ⫺20 mV under control conditions and in the presence of 30, 100, and 300 nM TTA-A2. D, concentration-response curve of TTA-A2 for Cav3.2 channels at holding potentials of ⫺100 and ⫺80 mV. Data points reflect mean ⫾ S.E. of three to four determinations. IC50 values of 5.6 ␮M (95% CI 3 to 10 ␮M) and 92 nM (95% CI 67 to 128 nM) and Hill coefficients of 1.1 (95% CI 2 to 0.3) and 1.4 (95% CI 1.6 to 0.7) were determined at ⫺100 and ⫺80 mV, respectively, using the equation of a sigmoidal dose-response curve. E, top, Cav3.3 current traces evoked by 70-ms depolarizations steps from ⫺100 to ⫺20 mV under control conditions and in the presence of 1, 3, and 10 ␮M TTA-A2. E, bottom, Cav3.3 current traces evoked by 70-ms depolarizations steps from ⫺80 to ⫺20 mV under control conditions and in the presence of 30, 100, and 300 nM TTA-A2. F, concentration-response curve of TTA-A2 for Cav3.3 channels at holding potentials of ⫺100 and ⫺80 mV. Data points reflect mean ⫾ S.E.M. of three to nine determinations. Error bars are hidden behind symbols. IC50 values of 3.7 ␮M (95% CI 3.1 to 4.3 ␮M) and 98 nM (95% CI 83 to 114 nM) and Hill coefficients of 1.2 (95% CI 1.5 to 1.0) and 1.4 (95% CI 1.3 to 0.9) were determined at ⫺100 and ⫺80 mV, respectively, using the equation of a sigmoidal dose-response curve.

Downloaded from jpet.aspetjournals.org at Merck & Co., Inc. on July 1, 2011

C

T-Type Calcium Channel Inhibition by TTA-A2

B 1.0

1.0

0.8

0.8

Normalized Tail Current

Normalized Tail Current

A

age-dependent manner, resulting in slowly decaying tail currents during a repolarization phase after the test pulse. Cav3.1 tail current amplitude at ⫺80 mV was measured after brief depolarizations (8 ms) to different test pulses from ⫺80 to ⫹ 50 mV in the absence or presence of TTA-A2 and plotted against test pulse potential. Activation curves were well described by a single Boltzmann function (Fig. 3A). TTA-A2 (300 nm) caused robust current inhibition at ⫺80 mV (82.7 ⫾ 1.8%, n ⫽ 3) but did not significantly shift midpoint of the Cav3.1 activation curve (V0.5act-control ⫽ ⫺40.5 ⫾ 1.2 mV, k ⫽ 7.4 ⫾ 0.7, and V0.5act-300 nM TTA-A2 ⫽ ⫺43.9 ⫾ 1.9 mV, k ⫽ 5.2 ⫾ 0.6; n ⫽ 4 – 6), indicating that the drug did not interfere with channel activation kinetics. Changes in Cav3 current inhibition as a function of the holding potential suggested that TTA-A2 block was statedependent and that compound may preferentially interact with inactivated channels. Drugs that bind and stabilize the inactivated state of voltage-gated channels shift the equilibrium from closed/resting to inactivated states and consequently induce a shift of the steady-state inactivation curve to more negative potentials. Channel availability curves in the absence of TTA-A2 and after bath equilibration with compound were established using tail current analysis. Following 10-s prepulses to potentials ranging from ⫺120 to ⫺35 mV and a subsequent 8-ms test pulse to ⫺20 mV, tail current amplitudes were measured at ⫺80 mV and plotted against prepulse potential. The resulting inactivation curves were fitted with the Boltzmann equation. Under the experimental conditions used, the midpoint (V0.5inact) of channel inactivation was shifted in a concentration-dependent manner to hyperpolarized potentials by 7 ⫾ 0.8 and 13.8 ⫾ 1.2 mV at 100 nM and 1 ␮M TTA-A2, respectively (Fig. 3B).

0.6 0.4 0.2

0.6 0.4 0.2 0.0

0.0 -80

-60

-40 -20 0 20 Test Potential (mV)

40

C

-120

D

-100 -80 -60 -40 Prepulse Potential (mV)

Control

0.0 800 600

Tau (ms)

Fraction Recovered

0.2 0.4

200 pA

200 pA

400

TTA-A2

200

0.6

0 Control

5 M TTA-A2

0.8 1.0

100 pA100

0

1000

2000

3000

4000

Time Interval (ms)

pA

5000 100 ms 100 ms

-20

Fig. 3. TTA-A2 effects on Cav3.1 channel gating. A, normalized steady-state Cav3.1 channel activation curves. Holding potential was ⫺80 mV; 8-ms test pulses were given from ⫺80 to ⫹50 mV. Tail currents were obtained at ⫺80 mV. Sweep interval was 3 s. Normalized peak tail current amplitudes were plotted against test potential. Activation curves in the absence and presence of 300 nM TTA-A2 were fitted to a single Boltzman function with V0.5act-control ⫽ ⫺40.5 ⫾ 1.2 mV, k ⫽ 7.4 ⫾ 0.7 and V0.5act-300 nM TTA-A2 ⫽ ⫺43.9 ⫾ 1.9 mV, k ⫽ 5.2 ⫾ 0.6; n ⫽ 4 – 6. B, effect of TTA-A2 on Cav3.1 channel availability. Cells were held at ⫺100 mV, and 10-s conditional prepulses were given from ⫺120 to ⫺35 mV followed by an 8-ms test pulse to ⫺20 mV. Tail currents were obtained at ⫺80 mV. Sweep interval was 10 s. Normalized tail current amplitudes are plotted against prepulse potential. TTA-A2 induced a concentration-dependent shift of the availability curve to hyperpolarized potentials (V0.5inact-control ⫽ ⫺75.9 ⫾ 0.5 mV, k ⫽ 4.7 ⫾ 0.2; V0.5inact-100 nM TTA-A2 ⫽ ⫺81.6 ⫾ 0.8 mV, k ⫽ 4.5 ⫾ 0.2; V0.5inact-1000 nM TTA-A2 ⫽ ⫺90.3 ⫾ 1.2 mV, k ⫽ 4.7 ⫾ 0.1; n ⫽ 6 –18). C, Cav3.1 channel recovery was measured using a standard paired-pulse protocol. Current amplitude during the second test pulse was normalized to the first test pulse and plotted as a function of the recovery interval. Membrane potential during the recovery period was ⫺100 mV. Time constants (␶) in the absence (closed circles) and presence of 5 ␮M TTA-A2 (open circles) were deduced from the curves shown. D, representative traces of current recovery experiments under control conditions and in the presence of 5 ␮M TTA-A2.

Downloaded from jpet.aspetjournals.org at Merck & Co., Inc. on July 1, 2011

ity curves, at negative membrane potentials, such as ⫺100 mV, most of the channels are in the closed/resting state, whereas a larger fraction of channels resides in the inactivated state at a more depolarized potential of ⫺80 mV. TTA-A2 showed increased affinity for Cav3.1 channels at this more depolarized potential (IC50 ⫽ 89 nM), indicating that block was state-dependent. Time course of block was similar at the two holding potentials tested. Recovery from block at ⫺80 mV was slower upon washout. TTA-A2 did not change the time course of macroscopic Cav3.1 current activation and inactivation kinetics during applied test pulses (Fig. 2A). Scaled currents remaining in the presence of drug showed an essentially identical time course compared with untreated current traces (not shown). To assess selectivity among Cav3 subtypes, TTA-A2 block of Cav3.2 and Cav3.3 channels was studied. Potency on Cav3.2 was comparable and also voltage-dependent. IC50 values of 5.6 ␮M and 92 nM were measured at ⫺100 and ⫺80 mV, respectively (Fig. 2, C and D). Cav3.3 channels show slower activation and inactivation kinetics than Cav3.1 and Cav3.2 channels. To capture evoked peak current amplitudes and to allow for complete recovery from inactivation between depolarizations, 70-ms test pulses to ⫺20 mV were given every 20 s. TTA-A2 also inhibited Cav3.3 currents voltagedependently, yielding similar IC50 values of 3.7 ␮M and 98 nM at ⫺100 and ⫺80 mV, respectively (Fig. 2, E and F). The Cav3.1 channel cell line was selected for robust current recordings with minimal current rundown during longer experiments to study the mechanism of T-channel block by TTA-A2 in more detail. First we analyzed compound effects on Cav3.1 channel gating. T-type calcium currents deactivate slowly in a volt-

413

414

Kraus et al.

% Current Inhibition

**

100

**

75 50

inhibition during the first pulse after 2-min equilibration in drug, was observed. At 0.1 Hz, negligible current decline due to accumulation of inactivation was observed under control conditions, and 1 ␮M TTA-A2 had only minor inhibitory effects (3 ⫾ 1 and 14.6 ⫾ 0.6%, respectively). At 3 Hz, Cav3.1 channels accumulate greater levels of inactivation because of the apparent slow rate of recovery from inactivation, leading to an overall current reduction of 32.2 ⫾ 6% in the absence of drug. Addition of 1 ␮M TTA-A2 significantly enhanced usedependent inhibition to 69.4 ⫾ 5% at steady state, presumably by stabilizing inactivated channels in between test pulses. Figure 4 summarizes current inhibition under both baseline control conditions and in the presence of TTA-A2 at 0.1 and 3 Hz. Although low-voltage T-type and high-voltage activated calcium channels share structural similarity, TTA-A2 exhibits remarkable selectivity over Cav1.2 (L-type), Cav2.1 (P/Qtype), Cav2.2 (N-type), and Cav2.3 (R-type) channels in patch-clamp studies. Potential state dependence of Cav1.2 inhibition by 30 ␮M TTA-A2 was investigated by clamping cells at ⫺90 mV (the majority of channels at rest) and ⫺40 mV (a larger fraction of channels inactivated). Cells were depolarized to ⫹ 10 mV every 10 s. In contrast to T-type calcium channels, Cav1.2 block by TTA-A2 did not show strong voltage dependence; 40 ⫾ 4.2 and 51.8 ⫾ 7.2% Cav1.2 current inhibition was measured at ⫺40 and ⫺90 mV, respectively (n ⫽ 3–5). Hence, recordings from Cav2.1, Cav2.2, and Cav2.3 channels were performed at ⫺90 mV only. At steady state, 30 ␮M TTA-A2 reduced Cav2.1 whole-cell current amplitude by 35.3 ⫾ 2.8% (n ⫽ 3), Cav2.2 amplitude by 56 ⫾ 3.6% (n ⫽ 4), and Cav2.3 amplitude by 51.2 ⫾ 4% (n ⫽ 4) (Fig. 5). Because of limited solubility of the compound, we were unable to test concentrations above 30 ␮M. Estimated IC50 values of ⬎30 ␮M on the tested high-voltage activated calcium channels suggested a ⬃300-fold selectivity for T-type channels (IC50 of ⬃0.1 ␮M) at ⫺80 mV. Next, we evaluated the effect of TTA-A2 on native T-currents recorded from thalamic relay neurons of the dLGN and STN neurons in rat brain slice preparations (Fig. 6), as these nuclei have been reported to express high levels of Cav3.1 or Cav3.3 channels, respectively (Talley et al., 1999). T-type currents were isolated from high-voltage-activated calcium currents and sodium currents by applying a 400-ms hyperpolarizing prepulse to ⫺100 mV followed by a 200-ms test

25 0

C

TTA-A2

0.1 Hz

C

TTA-A2

A

B

Cav1.2

Cav2.1

3 Hz

Fig. 4. Use-dependent Cav3.1 current inhibition by TTA-A2 at ⫺100 mV. Current inhibition (%) achieved at steady state at 0.1 and 3 Hz in control and in the presence of 1 ␮M TTA-A2 is plotted (n ⫽ 3–5). 3 ⫾ 1% current decline was observed at 0.1 Hz in control (hashed bar). TTA-A2 caused 14.6 ⫾ 0.6% inhibition at 0.1 Hz measured as amplitude reduction at steady state in TTA-A2 relative to the current amplitude at steady state in the absence of drug. At 3 Hz Cav3.1 channels accumulate inactivation in between test pulses, causing a current decrease by 32.2 ⫾ 6% measured as amplitude reduction at steady state during a 3-Hz pulse train relative to the current amplitude elicited by the first depolarization of the pulse train (hashed bar). TTA-A2 enhanced use-dependent inhibition at 3 Hz to 69 ⫾ 5% measured as amplitude reduction at steady state during a 3-Hz pulse train relative to the current amplitude elicited by the first depolarization of the pulse train after incubation in drug. Values are expressed as mean ⫾ S.E.M.; ⴱⴱ, p ⬍ 0.001. Representative current traces are shown on the top. Horizontal scale bar ⫽ 10 ms; vertical scale bar ⫽ 0.5 nA.

0.5 nA

30 M

30 10 M C

C

C 30 10 M C

D

Cav2.2

0.4 nA

Cav2.3

30 0.5 nA

10 M

0.5 nA 20 ms

C

Fig. 5. TTA-A2 activity on high-voltage-activated Cav1.2 (A), Cav2.1 (B), Cav2.2 (C), and Cav2.3 (D) channels. Currents were elicited by 100-ms test pulses to ⫹10 mV from a holding potential of ⫺90 mV. Traces under control conditions and at steady state in the presence of TTA-A2 are shown.

Downloaded from jpet.aspetjournals.org at Merck & Co., Inc. on July 1, 2011

Next, we investigated the effect of TTA-A2 on the time course of Cav3.1 channel recovery from fast inactivation, defined as calcium current recovery within 5 s. Using a standard double-pulse protocol, small current amplitudes and slow channel recovery between double pulses did not allow accurate measurement of compound effects on recovery from inactivation at ⫺80 mV. Consequently, cells were held at ⫺100 mV to obtain robust currents and enable maximal channel recovery within 5 s. Following a 100-ms conditional prepulse to ⫺20 mV, the time course of current recovery at ⫺100 mV was determined by applying 100-ms test pulses to ⫺20 mV at various time intervals after the prepulse. Cav3.1 channels were allowed to recover at ⫺100 mV for 15 s in between sweeps. Peak calcium currents evoked by the test pulse were normalized to the peak current amplitudes measured during the prepulse and plotted against time. As shown in Fig. 3C, Cav3.1 current recovery at ⫺100 mV was well described by a single exponential. A concentration of 5 ␮M TTA-A2, which caused approximately 50% current inhibition at ⫺100 mV, significantly delayed recovery from inactivation, evident as a 3.5-fold increase of the recovery time constant (␶co ⫽ 213 ⫾ 14 ms, ␶5␮M TTA-A2 ⫽ 764 ⫾ 44 ms; n ⫽ 10 –21). More frequent stimulations favor the availability of open and/or inactivated channels, thereby increasing the affinity of state-dependent blockers to their target. We tested to determine whether TTA-A2 exerts a use-dependent inhibitory component by increasing the stimulation frequency from 0.1 to 3 Hz (Fig. 4). To obtain robust Cav3.1 current amplitudes and fast channel recovery from inactivation between pulses, cells were held at ⫺100 mV, and a 3-Hz train of 20-ms test pulses to ⫺20 mV was applied under control conditions and after 2-min incubation in compound. Only a small tonic block component (usually ⬍5%), defined as peak current

T-Type Calcium Channel Inhibition by TTA-A2

A

Vehicle 10 mg/kg TTA-A2

dLGN

WT Mice 10

Cav3.1; 3.3 dual KO Active Wake

1 M

50 ms

C

B

STN

1 M 50 pA

Delta Sleep

REM Sleep

Mean Time in Sleep Stage (min)

50 pA

Mean Time in Sleep Stage (min)

Active Wake

3

415

Delta Sleep

REM Sleep

Fig. 6. TTA-A2 inhibits low-voltage-activated calcium currents recorded from the dorsal lateral geniculate nucleus (A) and the subthalamic nucleus (B) in a patch-clamp rat brain slice assay. Original T-type current traces recorded from neurons in the regions of interest in the absence and presence of TTA-A2 are shown.

pulse from a holding potential of ⫺60 mV in the presence of 250 nM tetrodotoxin. A test potential evoking maximal Ttype current amplitudes (typically ⫺35 mV) was determined for each cell using a standard current-voltage relationship protocol. Cells were stimulated every 15 s. Recorded currents were nickel-sensitive and blocked by the nonspecific T-type channel antagonist mibefradil (Kraus et al., 2007). TTA-A2 inhibited T-type currents recorded from both the dLGN (31.3 ⫾ 0.3% at 1 ␮M, 71 ⫾ 9% at 3 ␮M; 84 ⫾ 9.1% at 10 ␮M; n ⫽ 3) and the STN (44.5% ⫾ 14% at 1 ␮M; n ⫽ 2). Effects slowly reversed upon washout. To demonstrate the potential for TTA-A2 to effect neuronal excitability in vivo, the vigilance state of both wildtype mice and those lacking both Cav3.1 and Cav3.3 channels were monitored by ECoG/EMG polysomnography during administration of 10 mg/kg TTA-A2. This treatment has been demonstrated to elicit CNS exposures sufficient to affect locomotor activity, feeding, and weight gain in mice with a Tmax of 1 h (Uebele et al., 2009a). When administered 1 h before the inactive phase, TTA-A2 acutely suppressed the mean time spent in active wake for up to 2 h following dosing, while promoting delta sleep (Fig. 7). REM also appeared to be reduced, albeit with a delayed time course. As expected, none of these effects was observed in double knockout animals lacking both Cav3.1 and Cav3.3 channels.

Discussion Over the past few years, increasing attention has been focused on the family of low-voltage-activated T-type calcium channels. Functional characterization of T-type channels, including gene knockout and knockdown approaches in rodent models, have furthered our understanding of the physiological role of these channels and have indicated a role for these channels in several neurological disorders. The lack of potent and selective pharmacological tools, however, has so

T i me (h )

Time (h)

Fig. 7. Sleep-promoting effects of TTA-A2 are mediated by Cav3.1 and Cav3.3. ECoG/EMG sleep polysomnography was assessed in free-moving radio telemetry-implanted mice treated with either vehicle (0.5% methylcellulose/10% Tween 80 p.o.; closed circles) or 10 mg/kg TTA-A2 (open circles) in a 5-day crossover paradigm, wild-type mice (n ⫽ 6), Cav3.1, and Cav3.3 double mutant mice (n ⫽ 6). Once-daily treatments occurred at 1:00 PM (arrows), 1 h before lights on in a crossover design, such that each animal received both vehicle and compound for 5 consecutive days. Mean time spent in each sleep stage accumulated over each 30-min interval are plotted. Values represent the average of 5 days of vehicle or compound treatment. Intervals at which significant differences exist between control and experimental values are indicated by gray vertical lines with short, medium, and long tick marks representing p values ⬍0.05, 0.01, or 0.001, respectively (linear mixed effects model for repeated measures).

far hampered research efforts to validate T-type channels as targets for drug discovery. Using the recently synthesized antagonist, TTA-A2 (Barrow et al., 2007; Uebele et al., 2009a), we demonstrate the effects of T-type channel inhibition in a cellular context, in vivo on the vigilance state of mice and the therapeutic potential of T-type channel inhibition. Pharmacological Characterization of TTA-A2. TTA-A2 was found to inhibit all three subtypes of low-voltage-gated T-type channels (Cav3.1, 3.2, and 3.3) with comparable potencies. Changing membrane resting potentials from ⫺100 to ⫺80 mV enhanced compound potency by ⬃40-fold, from an IC50 of ⬃4 ␮M to 0.1 ␮M, respectively, indicating state dependence of inhibition. We could further demonstrate a concentration-dependent shift of the Cav3.1 channel availability curve to hyperpolarized potentials and a slowing of channel recovery from the inactivated state. According to the modulated receptor hypotheses, enhanced T-type channel block by TTA-A2 at depolarized voltages in combination with a shift of the channel availability curve to hyperpolarized potentials can be interpreted as high affinity block of inactivated channels. Furthermore, an increase in channel activation frequency from 0.1 to 3 Hz, to favor the availability of open and/or inactivated channels, revealed an additional use-dependent component of inhibition. Experiments to measure channel recovery from inactivation and use dependence of inhibition had to be performed at ⫺100 mV to obtain robust Cav3.1 current amplitudes, enable fast channel recovery, and

Downloaded from jpet.aspetjournals.org at Merck & Co., Inc. on July 1, 2011

100 ms

C

416

Kraus et al.

ably, none of these effects was observed in the double-knockout animals. In addition to demonstrating the in vivo efficacy of TTA-A2 and its specificity for T-type channels, these results demonstrate that the slow wave sleep-promoting effects of this compound are mediated specifically through Cav3.1 and/or Cav3.3 channels. Furthermore, because Cav3.1 and Cav3.3 are expressed in cortex, as well as subthalamic and thalamic relay nuclei, whereas Cav3.2 is present in sensory regions, amygdala and hippocampus (Talley et al., 1999), these findings implicate thalamocortical circuits in the action of TTA-A2. Interestingly, Lee et al. (2004) demonstrated a reduction of delta (1– 4 Hz) and spindle (7–14 Hz) oscillations measured by non-REM electrocorticogram frequency analysis in Cav3.1 global knockout mice, and Anderson et al. (2005) reported a moderate increase in EEG spectral power within the delta frequency range, when a focal deletion of Cav3.1 was confined to the rostral midline thalamus. These apparently conflicting findings suggest that the complex neuronal network interactions of multiple nuclei in the thalamocortical system, expressing predominantly Cav3.1 and Cav3.3 channels, are not easily understood in genetic models and that the net behavioral effects of an acute pharmacological blockade are not equivalent to constitutive or focal channel ablation. Furthermore, in contrast to an absolute ablation of T-channels in knockout animals, T-channel inhibition by TTA-A2 is dependent on channel state and activation frequency and is therefore modulated by animal state-dependent neuronal network activity in vivo. A potential mechanistic explanation underlying the suppressed active wake and increased delta sleep following TTA-A2 administration may be an increased interburst interval because of slowing of the T-type-mediated initiation of bursting activity of the thalamocortical neuronal network. As previously demonstrated, the inhibition of Ttype currents in thalamic reticular nucleus neurons, which establish the pace of recurrent thalamocortical bursts by providing inhibitory input onto the thalamus, can delay the onset of bursting (Huguenard and Prince, 1994) and thus may promote slower oscillatory activity in the network associated with increased non-REM sleep. T-type channels mediate rebound bursts in cortical and thalamic reticular neurons that are the physiological substrates for spike-wave discharges, the electrographic hallmark of absence seizures. Increased T-type currents have been suggested to elevate the propensity for seizure activity (Khosravani and Zamponi, 2006; Ernst et al., 2009). We have previously shown that selective T-type channel antagonists, including TTA-A2, suppress seizure time in the WAG/Rij rat absence epilepsy model (Shipe et al., 2008; Yang et al., 2008; Uebele et al., 2009b), providing further evidence of the ability of selective T-channel antagonists to modulate thalamocortical network activity. Clinical Potential of T-Type Channel Inhibition. Besides insomnia and epilepsy, pathological changes of membrane potential regulation can lead to thalamocortical dysrhythmia as it was observed in awake patients suffering from neurogenic pain, tinnitus, Parkinson’s disease, schizophrenia, and depression (Llina´s et al., 1999). Compared with healthy controls, these patients showed increased thalamocortical oscillations in the ␪-frequency band. Cortical ␪-waves (4 – 8 Hz) usually appear during drowsy or sleeping states. ␪-Activity is mediated by resonant interaction between the

Downloaded from jpet.aspetjournals.org at Merck & Co., Inc. on July 1, 2011

minimize accumulation of inactivated channels during 3-Hz use-dependence protocols. These experimental conditions provide insight into the mechanism of action of TTA-A2 but may underestimate the compound effects at more depolarized physiological membrane potentials. At such potentials, TTA-A2 most probably causes an even more pronounced delay of channel recovery from inactivation, because of slower dissociation from the receptor, and consequently stronger use-dependent inhibition, consistent with the nanomolar potencies measured at ⫺80 mV and compound-induced effects observed in vivo. Similar state and frequency dependence has been described for a variety of high-voltage-activated calcium and sodium channel blockers in the past (Hondeghem and Katzung, 1984). The exact molecular determinants of TTA-A2 interactions with T-type channels are currently unknown, but previous competition studies in our lab suggested that the binding sites of TTA-A2 and other structurally similar antagonists overlap or interact allosterically with the binding site of mibefradil (Uebele et al., 2009b). Similar to TTAA2, mibefradil has been reported to preferentially interact with T-type channels in the inactivated state (Martin et al., 2000). Selectivity of TTA-A2 for T-type calcium channels was previously demonstrated in 170 binding and fluorescent assays, including numerous voltage-activated sodium, calcium, and potassium channels. No activities with IC50 ⬍10 ␮M were reported (Uebele et al., 2009a). In the present study we have used electrophysiological patch-clamp experiments at holding potentials similar to neuronal resting membrane potentials (⫺90 or ⫺80 mV) to demonstrate a ⬃300-fold selectivity for Cav3 channels over structurally related high-voltage-activated Cav1.2, Cav2.1, Cav2.2, and Cav2.3 channels. Frequency dependence of T-channel inhibition by TTA-A2 also suggests potential functional selectivity over other ion channels. This improvement on selectivity clearly distinguishes TTA-A2 from existing blockers such as mibefradil and ethosuximide (Martin et al., 2000; Gomora et al., 2001). Before testing TTA-A2 in vivo, compound activity was also confirmed on native T-type calcium channel-mediated currents recorded from neurons in the rat subthalamic nucleus and the rat dorsal lateral geniculate nucleus of the thalamus in a brain slice preparation. In Vivo Effects of T-Type Calcium Channel Inhibition. It is well known that T-type calcium channels mediate low-threshold spikes in the thalamus that are essential to sleep-related network oscillations involved in sleep maintenance and sleep/wake transitions (McCormick and Bal, 1997; Cueni et al., 2008). Inhibition of T-currents by depolarization of thalamic neurons or ethosuximide has been shown to affect thalamic oscillations (Huguenard and Prince, 1994; McCormick and Bal, 1997), and mice lacking Cav3.1 channels display altered sleep/wake patterns because of perturbed thalamocortical activity (Kim et al., 2001; Lee et al., 2004; Anderson et al., 2005). Consequently, it had been proposed that potent and selective T-type channel modulators could serve as pharmacological tools to modulate sleep/wake cycles. To test this hypothesis, the effect of TTA-A2 on the vigilance state of both wild-type mice and Cav3.1/Cav3.3 double knockouts was monitored in ECoG/EMG polysomnographic studies. TTA-A2 acutely suppressed mean time spent in active wake and promoted delta sleep in wild-type mice. Remark-

T-Type Calcium Channel Inhibition by TTA-A2

thalamus and the cortex, due to the generation of low-threshold spike bursts by thalamic cells (Crunelli et al., 2006). The abnormal oscillations mediated by T-type calcium channeltriggered low-threshold spikes may be an underlying mechanism for thalamocortical dysrhythmias (Jeanmonod et al., 1996; Llina´s et al., 1999). We have shown that thalamocortical network activity can be modulated by TTA-A2, and that these effects depend on the presence of T-type channels. Hence, the termination of abnormal thalamocortical oscillations with selective T-type calcium channel antagonists could offer an interesting novel therapeutic approach to treat a diversity of diseases associated with thalamocortical dysrhythmias.

Conclusions

References Anderson MP, Mochizuki T, Xie J, Fischler W, Manger JP, Talley EM, Scammell TE, and Tonegawa S (2005) Thalamic Cav3.1 T-type Ca2⫹ channel plays a crucial role in stabilizing sleep. Proc Natl Acad Sci USA 102:1743–1748. Barrow JC, Bieber KAS, Cube RV, Mattern MC, Reger TS, Shu Y, and Yang Z-Q (2007) inventor; Merck & Co., Inc., assignee. Pyridyl amide T-type calcium channel antagonists. World Patent WO 2007120729. 2007 Oct 25. Beenhakker MP and Huguenard JR (2009) Neurons that fire together also conspire together: is normal sleep circuitry hijacked to generate epilepsy? Neuron 62:612– 632. Bourinet E, Alloui A, Monteil A, Barre`re C, Couette B, Poirot O, Pages A, McRory J, Snutch TP, Eschalier A, et al. (2005) Silencing of the Cav3.2 T-type calcium channel gene in sensory neurons demonstrates its major role in nociception. EMBO J 24:315–324. Catterall WA, Perez-Reyes E, Snutch TP, and Striessnig J (2005) International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharmacol Rev 57:411– 425. Choi S, Na HS, Kim J, Lee J, Lee S, Kim D, Park J, Chen CC, Campbell KP, and Shin HS (2007) Attenuated pain responses in mice lacking Ca(V)3.2 T-type channels. Genes Brain Behav 6:425– 431. Crunelli V, Cope DW, and Hughes SW (2006) Thalamic T-type Ca2⫹ channels and NREM sleep. Cell Calcium 40:175–190. Cueni L, Canepari M, Luja´n R, Emmenegger Y, Watanabe M, Bond CT, Franken P, Adelman JP, and Lu¨thi A (2008) T-type Ca2⫹ channels, SK2 channels and SERCAs gate sleep-related oscillations in thalamic dendrites. Nat Neurosci 11:683– 692. Dai G, Haedo RJ, Warren VA, Ratliff KS, Bugianesi RM, Rush A, Williams ME, Herrington J, Smith MM, McManus OB, et al. (2008) A high-throughput assay for evaluating state dependence and subtype selectivity of Cav2 calcium channel inhibitors. Assay Drug Dev Technol 6:195–212. Ernst WL, Zhang Y, Yoo JW, Ernst SJ, and Noebels JL (2009) Genetic enhancement of thalamocortical network activity by elevating alpha 1g-mediated low-voltageactivated calcium current induces pure absence epilepsy. J Neurosci 29:1615– 1625. Gomora JC, Daud AN, Weiergra¨ber M, and Perez-Reyes E (2001) Block of cloned human T-type calcium channels by succinimide antiepileptic drugs. Mol Pharmacol 60:1121–1132. Heady TN, Gomora JC, Macdonald TL, and Perez-Reyes E (2001) Molecular pharmacology of T-type Ca2⫹ channels. Jpn J Pharmacol 85:339 –350. Hille B (1977) Local anesthetics: hydrophilic and hydrophobic pathways for the drug-receptor reaction. J Gen Physiol 69:497–515. Hondeghem LM and Katzung BG (1984) Antiarrhythmic agents: the modulated

receptor mechanism of action of sodium and calcium channel-blocking drugs. Annu Rev Pharmacol Toxicol 24:387– 423. Huguenard JR (1999) Neuronal circuitry of thalamocortical epilepsy and mechanisms of antiabsence drug action. Adv Neurol 79:991–999. Huguenard JR and Prince DA (1994) Intrathalamic rhythmicity studied in vitro: nominal T-current modulation causes robust antioscillatory effects. J Neurosci 14:5485–5502. Jeanmonod D, Magnin M, and Morel A (1996) Low-threshold calcium spike bursts in the human thalamus. Common physiopathology for sensory, motor and limbic positive symptoms. Brain 119:363–375. Khosravani H and Zamponi GW (2006) Voltage-gated calcium channels and idiopathic generalized epilepsies. Physiol Rev 86:941–966. Kim D, Park D, Choi S, Lee S, Sun M, Kim C, and Shin HS (2003) Thalamic control of visceral nociception mediated by T-type Ca2⫹ channels. Science 302:117–119. Kim D, Song I, Keum S, Lee T, Jeong MJ, Kim SS, McEnery MW, and Shin HS (2001) Lack of the burst firing of thalamocortical relay neurons and resistance to absence seizures in mice lacking alpha(1G) T-type Ca(2⫹) channels. Neuron 31:3– 4. Kraus RL, Li Y, Jovanovska A, and Renger JJ (2007) Trazodone inhibits T-type calcium channels. Neuropharmacology 53:308 –317. Lee J, Kim D, and Shin HS (2004) Lack of delta waves and sleep disturbances during non-rapid eye movement sleep in mice lacking alpha1G-subunit of T-type calcium channels. Proc Natl Acad Sci U S A 101:18195–18199. Leuranguer V, Mangoni ME, Nargeot J, and Richard S (2001) Inhibition of T-type and L-type calcium channels by mibefradil: physiologic and pharmacologic bases of cardiovascular effects. J Cardiovasc Pharmacol 37:649 – 661. Llina´s RR, Ribary U, Jeanmonod D, Kronberg E, and Mitra PP (1999) Thalamocortical dysrhythmia: A neurological and neuropsychiatric syndrome characterized by magnetoencephalography. Proc Natl Acad Sci U S A 96:15222–15227. Martin RL, Lee JH, Cribbs LL, Perez-Reyes E, and Hanck DA (2000) Mibefradil block of cloned T-type calcium channels. J Pharmacol Exp Ther 295:302–308. McCormick DA and Bal T (1997) Sleep and arousal: thalamocortical mechanisms. Annu Rev Neurosci 20:185–215. Mehrke G, Zong XG, Flockerzi V, and Hofmann F (1994) The Ca(⫹⫹)-channel blocker Ro 40 –5967 blocks differently T-type and L-type Ca⫹⫹ channels. J Pharmacol Exp Ther 271:1483–1488. Perez-Reyes E (2003) Molecular physiology of low-voltage-activated T-type calcium channels. Physiol Rev 83:117–161. Santi CM, Cayabyab FS, Sutton KG, McRory JE, Mezeyova J, Hamming KS, Parker D, Stea A, and Snutch TP (2002) Differential inhibition of T-type calcium channels by neuroleptics. J Neurosci 22:396 – 403. Shin HS, Cheong EJ, Choi S, Lee J, and Na HS (2008) T-type Ca2⫹ channels as therapeutic targets in the nervous system. Curr Opin Pharmacol 8:33– 41. Shipe WD, Barrow JC, Yang ZQ, Lindsley CW, Yang FV, Schlegel KA, Shu Y, Rittle KE, Bock MG, Hartman GD, et al. (2008) Design, synthesis, and evaluation of a novel 4-aminomethyl-4-fluoropiperidine as a T-type Ca2⫹ channel antagonist. J Med Chem 51:3692–3695. Steriade M, McCormick DA, and Sejnowski TJ (1993) Thalamocortical oscillations in the sleeping and aroused brain. Science 262:679 – 685. Talley EM, Cribbs LL, Lee JH, Daud A, Perez-Reyes E, and Bayliss DA (1999) Differential distribution of three members of a gene family encoding low voltageactivated (T-type) calcium channels. J Neurosci 19:1895–1911. Traboulsie A, Chemin J, Kupfer E, Nargeot J, and Lory P (2006) T-type calcium channels are inhibited by fluoxetine and its metabolite norfluoxetine. Mol Pharmacol 69:1963–1968. Uebele VN, Gotter AL, Nuss CE, Kraus RL, Doran SM, Garson SL, Reiss DR, Li Y, Barrow JC, Reger TS, et al. (2009a) Antagonism of T-type calcium channels inhibits high-fat diet-induced weight gain in mice. J Clin Invest 119:1659 –1667. Uebele VN, Nuss CE, Fox SV, Garson SL, Cristescu R, Doran SM, Kraus RL, Santarelli VP, Li Y, Barrow JC, et al. (2009b) Positive allosteric interaction of structurally diverse T-type calcium channel antagonists. Cell Biochem Biophys 55:81–93. Van der Vring JA, Cleophas TJ, Van der Wall EE, and Niemeyer MG (1999) T-channel-selective calcium channel blockade: a promising therapeutic possibility, only preliminarily tested so far: a review of published data. T-Channel Calcium Channel Blocker Study Group. Am J Ther 6:229 –233. Xia M, Imredy JP, Koblan KS, Bennett P, and Connolly TM (2004) State-dependent inhibition of L-type calcium channels: cell-based assay in high-throughput format. Anal Biochem 327:74 – 81. Yang ZQ, Barrow JC, Shipe WD, Schlegel KA, Shu Y, Yang FV, Lindsley CW, Rittle KE, Bock MG, Hartman GD, et al. (2008) Discovery of 1,4-substituted piperidines as potent and selective inhibitors of T-type calcium channels. J Med Chem 51: 6471– 6477. Zamponi GW, Lewis RJ, Todorovic SM, Arneric SP, and Snutch TP (2009) Role of voltage-gated calcium channels in ascending pain pathways. Brain Res Rev 60: 84 – 89.

Address correspondence to: Richard L. Kraus, Department of Neuroscience, Merck Research Laboratories, 770 Sumneytown Pike, West Point, PA 19486. E-mail: [email protected]

Downloaded from jpet.aspetjournals.org at Merck & Co., Inc. on July 1, 2011

TTA-A2 is a potent, state-dependent, and highly selective T-type calcium channel antagonist. Using TTA-A2 we were able to demonstrate the effects of pharmacological T-type calcium channel inhibition on thalamocortical activity in vivo, as assessed in mouse polysomnographic experiments and previously reported rat absence epilepsy studies. TTA-A2 represents an invaluable tool for elucidating the true physiological roles of T-type calcium channels and for exploring these channels as drug targets for a variety of human neurological diseases, including sleep disorders and epilepsy.

417