Skip to main content
Annals of Clinical and Translational Neurology logoLink to Annals of Clinical and Translational Neurology
. 2023 Feb 4;10(4):462–483. doi: 10.1002/acn3.51735

T‐type calcium channels as therapeutic targets in essential tremor and Parkinson's disease

Lillian G Matthews 1,, Corey B Puryear 1, Susana S Correia 1, Sharan Srinivasan 1,2, Gabriel M Belfort 1, Ming‐Kai Pan 3,4,5,6, # , Sheng‐Han Kuo 7,8, #
PMCID: PMC10109288  PMID: 36738196

Abstract

Neuronal action potential firing patterns are key components of healthy brain function. Importantly, restoring dysregulated neuronal firing patterns has the potential to be a promising strategy in the development of novel therapeutics for disorders of the central nervous system. Here, we review the pathophysiology of essential tremor and Parkinson's disease, the two most common movement disorders, with a focus on mechanisms underlying the genesis of abnormal firing patterns in the implicated neural circuits. Aberrant burst firing of neurons in the cerebello‐thalamo‐cortical and basal ganglia‐thalamo‐cortical circuits contribute to the clinical symptoms of essential tremor and Parkinson's disease, respectively, and T‐type calcium channels play a key role in regulating this activity in both the disorders. Accordingly, modulating T‐type calcium channel activity has received attention as a potentially promising therapeutic approach to normalize abnormal burst firing in these diseases. In this review, we explore the evidence supporting the theory that T‐type calcium channel blockers can ameliorate the pathophysiologic mechanisms underlying essential tremor and Parkinson's disease, furthering the case for clinical investigation of these compounds. We conclude with key considerations for future investigational efforts, providing a critical framework for the development of much needed agents capable of targeting the dysfunctional circuitry underlying movement disorders such as essential tremor, Parkinson's disease, and beyond.

Introduction

Essential tremor (ET) and Parkinson's disease (PD) are the two most common movement disorders worldwide. In the United States, an estimated 7 million individuals have ET, representing approximately 2% of the total population, 1 which is slightly higher than reported global prevalence estimates of 0.32–1.33%. 2 , 3 Prevalence of ET increases with age, with rates reaching 5.8% in patients aged ≥65 years and estimated to increase ~1.7‐fold with every decade of life; indeed, rates exceeding 20% have been observed in the 9th and 10th decades of life. 2 Prevalence estimates for PD indicate there are approximately 1 million affected individuals aged ≥45 years in the United States, with prevalence increasing from 0.4% in patients aged 65–74 years, to nearly 2% in those >80 years. 4 , 5 Importantly, with an increasingly aging population, novel approaches for treating neurologic diseases are paramount. A large focus of ET and PD research has centered around the degeneration of specific cell populations, 6 , 7 , 8 , 9 , 10 with the objective of developing disease‐modifying therapies, although, to date, no such interventions have been approved. 11 , 12 More recently, there has been increasing interest in targeting dysregulated neuronal activity as a novel and promising approach to treating ET and PD.

Aberrant neuronal activity has been implicated in the pathology of several neurologic conditions, including epilepsy, tinnitus, neuropathic pain, dystonia, ET, PD, and major depression. 13 , 14 , 15 , 16 , 17 The role of dysfunctional circuitry is particularly well studied in ET and PD. 18 , 19 , 20 Specifically, the cerebello‐thalamo‐cortical (CTC) and basal ganglia‐thalamo‐cortical (BG) circuits play pivotal roles in motor function, with related CTC and BG disturbances widely implicated in ET and PD, respectively. 21 , 22 , 23 , 24 , 25 , 26 Abnormal burst firing and dysregulated oscillatory neural activity have been identified as important factors mediating tremor and motor symptoms in both conditions. 18 , 27 , 28 , 29 , 30 , 31 , 32 Surgical ablation and deep brain stimulation (DBS) provide supportive evidence for firing pattern modulation in ET and PD. Targeted disruption of the thalamic ventral intermediate nucleus (VIM) via DBS can reduce burst firing in patients with ET and improve tremor symptoms. 33 , 34 , 35 Additionally, surgical thalamic ablation can improve symptoms in both ET and tremor‐dominant PD. 36 , 37 In patients with PD, DBS of the internal globus pallidus (GPi) or subthalamic nucleus (STN) can improve rigidity, whereas bradykinesia can be improved by targeting the STN. 38 Despite success in many patients, DBS is not without risks common to surgical interventions, including perioperative hemorrhage and infection, with reduced efficacy also observed following long‐term use. 39 , 40 , 41 , 42 Consequently, treatment options become limited as the disease advances, 43 implicating an urgent need for targeted, noninvasive pharmacologic approaches that can modulate the aberrant circuit activity in ET and PD and provide symptomatic relief.

The search for pharmacologic options has led to identification of various approaches to modulating neural circuit activity in the CTC and BG. One of the most promising targets are T‐type calcium channels (TTCC), known to be involved in the generation of burst activity in several brain areas. 44 , 45 , 46 Notably, TTCCs are highly expressed in specific structures of CTC and BG circuits and play an important role in modulating neuronal action potential firing patterns. 47 , 48 , 49 , 50 , 51 , 52 , 53 Furthermore, increasing evidence highlights their role in the aberrant bursting phenomenon observed in ET and PD, 18 , 29 , 54 with TTCC blockade shown to be capable of inhibiting this bursting phenotype and ameliorating symptoms in animal models of both ET and PD. 29 , 54 , 55 , 56 Here, we discuss in depth the mechanistic role of TTCCs in the pathophysiology of ET and PD. More specifically, we review the evidence supporting the potential for TTCC blockade as a pharmacologic approach to symptomatic therapy, with the goal of informing future efforts in developing TTCC‐targeted therapies for patients with ET and PD.

TTCC Overview and CNS Expression Patterns

TTCCs—named “T‐type” due to their transient calcium conductance—are low voltage‐activated calcium channels with key roles in membrane excitability. 57 , 58 These channels allow calcium entry into cells at relatively hyperpolarized membrane potentials (CaV3; −70 to −55 mV) 58 , 59 compared with high voltage‐activated calcium channels, which activate and allow calcium entry at more depolarized membrane potentials (CaV1 and CaV2; > −40 mV). 58 A brief hyperpolarizing event within a single cell is sufficient to activate TTCCs, facilitating depolarization of the membrane and activation of low‐threshold calcium current leading to generation of low voltage‐activated calcium spikes, otherwise known as rebound bursting. 59 This voltage sensitivity allows TTCCs to support a window current that is near the resting membrane potential, in turn regulating neuronal excitability. 60 , 61

CaV3.1, 62 CaV3.2, 63 and CaV3.3 64 constitute the three known TTCC isoforms, each with unique biophysical features and distribution patterns. 62 , 63 , 65 , 66 , 67 , 68 In humans, CaV3.1, CaV3.2, and CaV3.3 are encoded by the genes CACNA1G (chromosome 17), CACNA1H (chromosome 16), and CACNA1I (chromosome 22), respectively. 69 TTCCs are expressed widely throughout the brain and in peripheral tissues, with particularly high expression within cerebellar motor circuits and tremor‐related structures (Figures 1 and 2, Figure S1). 66 , 70 , 71 , 72 , 73 , 74 , 75 CaV3.1 channels are expressed in the soma and dendrites of neurons in several brain regions including the cerebellum, thalamus, hypothalamus, olfactory bulb, basal ganglia, cortex, and hippocampus, as shown by rodent and human data. 62 , 65 , 66 , 67 , 71 , 72 , 73 , 74 , 75 One of the key functions of CaV3.1 is to regulate the oscillatory activity of thalamic neurons, as first described in cat dorsal lateral geniculate nucleus. 76 While neural oscillation can refer to rhythmic electrical activity of clusters of activated neurons, individual neurons demonstrate rhythmic oscillation, as shown in the work of McCormick and Pape via intracellular recordings within individual cat dorsal lateral geniculate relay neurons. 76 Notable high expression levels of CaV3.1 in mouse and human cerebellum 71 , 72 , 73 , 74 , 75 suggest its potential further role in cerebellar circuit function. Although more widely expressed than CaV3.1, 77 CaV3.2 channels in mouse are predominantly expressed in the hippocampus (dentate gyrus), 72 , 73 , 78 while in human brain, expression appears greatest in the thalamus and basal ganglia. 71 , 74 , 75 CaV3.2 channels have been shown to control NMDA receptor‐mediated transmission at central synapses in cultured rat hippocampal cells expressing the human variant. 79 CaV3.2 are also required for acute pain perception 80 and excitability of spinal cord lamina II neurons, as demonstrated in mice. 81 CaV3.3 channels are similarly expressed in various brain regions in both mouse 72 , 73 and human brain, 74 , 75 and have been shown to facilitate neuronal oscillation in mice, particularly in the thalamus, where they are key for generating sleep spindles during non‐rapid eye movement (NREM) sleep. 82 Published data to date suggest that CaV3.3, and to a lesser extent CaV3.2, plays the dominant role in sleep spindle rhythmogenesis, whereas CaV3.1 appears to play a minimal role in sleep spindle activity. 82 , 83 , 84 , 85 It should be noted, however, that these reports rely on indirect measurement of sleep spindles (EEG power in the sigma frequency range, 10–15 Hz). As such, a deeper analysis of specific sleep spindle features (e.g., spindle amplitude, frequency, duration, etc.) may highlight more nuanced roles each CaV isoform plays in the generation and/or regulation of sleep spindles.

Figure 1.

Figure 1

Gene expression patterns of TTCCs in mouse brain. Three‐dimensional gene expression patterns of CACNA1G (CaV3.1), CACNA1H (CaV3.2), and CACNA1I (CaV3.3) in mouse whole brain. Gene expression patterns for each isoform are shown for all brain regions on the left (cortical distribution visible), and for regions of greatest subcortical expression on the right—cerebellum, thalamus, and hippocampus. Red and green denote regions of highest and lowest gene expression levels, respectively. The greatest expression of CaV3.1 is in the cerebellum and the greatest expression of CaV3.2 is in the hippocampus. Images created from the Allen Mouse Brain Atlas using Brain Explorer 2.0.

Figure 2.

Figure 2

Gene expression patterns of TTCCs in human brain. (A) Microarray‐based gene expression of CACNA1G (CaV3.1), CACNA1H (CaV3.2), and CACNA1I (CaV3.3) in human brain. Three‐dimensional expression patterns from a representative donor brain are shown for each isoform, overlaid on basal ganglia (purple), thalamus (green), and cerebellum (cyan). Data are normalized across the entire set of microarray samples and shown as z‐scores. Images created from the Allen Human Brain Atlas using Brain Explorer 2.0. (B) Violin plots depicting in situ hybridization‐based gene expression of CACNA1G (CaV3.1), CACNA1H (CaV3.2), and CACNA1I (CaV3.3) in human brain. Gene expression values are shown in transcripts per million, calculated from a gene model with isoforms collapsed to a single gene. No other normalization steps have been applied. Box plots are shown as median and 25th and 75th percentiles. Plots generated and adapted from The Genotype‐Tissue Expression (GTEx) Project portal.

A key function of TTCCs is to regulate neuronal firing patterns (Figure 3). 86 , 87 , 88 , 89 Burst firing and associated subthreshold membrane potential oscillations, mediated by TTCC activation, occur in various brain areas and networks, including populations of thalamic, cerebellar, and cortical neurons, and modulate thalamocortical oscillations, spike–wave discharges, and neuropathic pain. 16 , 44 , 90 , 91 , 92 , 93 , 94 , 95 , 96 , 97 , 98 Of note, TTCC genetic mutations have been identified in three families diagnosed with ET and have been implicated in spinocerebellar ataxia 42, epilepsy, autism, schizophrenia, developmental disorders, and other neurologic conditions (Table 1). 87 , 99 , 100 , 101 , 102 , 103 , 104 , 105 , 106 , 107 , 108 , 109 , 110 Given their key roles in regulating neuronal firing, pharmacologically targeting TTCCs thus has the potential to yield a wealth of therapeutic benefits in various neurologic disorders. Importantly, while various Food and Drug Administration (FDA)‐approved agents, including anticonvulsants, calcium channel blockers, and psychotropics, exhibit some degree of TTCC blockade (Table 2), 111 , 112 , 113 , 114 , 115 , 116 , 117 , 118 , 119 , 120 no currently approved agent selectively targets TTCCs as a primary mechanism of action for its indicated use.

Figure 3.

Figure 3

TTCC neuronal firing patterns. (A) Illustrations representing the firing patterns that are dependent on the contribution of TTCC within various brain areas and networks, including populations of thalamic, cerebellar, and cortical neurons. Tonic firing (left) is more likely to occur in cells with either low TTCC expression or in neurons with deactivated TTCC due to a depolarized membrane potential. Low threshold burst firing (right) is likely to occur in cells that have a higher density of TTCC channels, such as those in the VIM or STN, or in neurons with hyperpolarized membranes that activate TTCC. (B) Membrane hyperpolarization activates TTCC, leading to production of low threshold spikes and subsequent rebound burst firing. Adapted from Park et al. 86 as licensed under CC BY 4.0. (C) CaV3.1, CaV3.2, and CaV3.3 current traces elicited by action potentials in cultured HEK‐293 cells. (D) Illustration of the window current (orange shading) of TTCCs which arises due to overlap of steady‐state inactivation and activation properties. The window current of TTCCs occurs within the range of the resting membrane potential (blue shading) and has been shown to be enlarged in patients with CACNA1G mutations and inhibited by TTCC blockade. 87 (C and D) Adapted from Lory et al. 88 as licensed under CC BY 4.0. I h, HCN channel‐mediated currents; I T, T‐type calcium current; NaV, voltage‐gated sodium channel; Vm, membrane potential.

Table 1.

TTCC mutations in neurologic disorders.

Isoform Disease Protein‐level mutation(s) Effect on TTCC electrophysiological properties
CaV3.1/CACNA1G 99 Essential tremor
  • Gly627Arg

  • Arh456Glu

  • Arg1235Gln

  • Little to no difference versus wild type

CaV3.1/CACNA1G 100 Spinocerebellar ataxia 42
  • Ala961Thr

  • Unknown

CaV3.1/CACNA1G 101 , 102 , 103

Spinocerebellar ataxia 42
  • Arg1715His

  • Shift toward more positive membrane potentials (inactivation and activation)

CaV3.1/CACNA1G 104

Spinocerebellar ataxia 42
  • Met1574Lys

  • Unknown

CaV3.1/CACNA1G 105

Autosomal‐dominant cerebellar syndrome
  • Arg1715His

  • Shift toward more positive membrane potentials (inactivation and activation)

CaV3.1/CACNA1G 106

Infantile‐onset syndromic cerebellar atrophy
  • Ala961Thr

  • Met1531Val

  • Gain of function

  • Gain of function

CaV3.1/CACNA1G 87

Childhood‐onset cerebellar atrophy
  • Ala961Thr

  • Met1531Val

  • Gain of function

  • Gain of function

CaV3.2/CACNA1H 107

Autism
  • Arg212Cys

  • Arg902Trp

  • GLy2886Cys

  • Ala1874Val/Arg1871Gln

  • Loss of function

  • Loss of function

  • Loss of function

  • Loss of function

CaV3.3/CACNA1I 108

Schizophrenia
  • Thr797Met

  • Arg1311His

  • Unknown

  • Unknown

CaV3.3/CACNA1I 109 , 110

Schizophrenia
  • Thr797Met

  • Arg1346His

  • Unknown

Table 2.

FDA‐approved agents with TTCC blocking properties.

Approved clinical uses
Anticonvulsants
Ethosuximide 111 Absence seizures
Zonisamide 112 Partial seizures, absence seizures, infantile spasms, essential tremor (second line)
Trimethadione 113 Absence seizures
Valproic acid 114 Partial seizures, absence seizures, bipolar disorder, migraine
Calcium channel blockers
Verapamil 115 Hypertension, angina
Amlodipine 116 Hypertension, coronary artery disease, angina
Nicardipine 116 Hypertension, angina
Nimodipine 116 Subarachnoid hemorrhage
Psychotropics
Pimozide 117 Tourette's syndrome
Fluoxetine 118 , 119 Major depressive disorder, anxiety disorders, obsessive compulsive disorder, eating disorders, premenstrual dysphoric disorder
Trazodone 120 Major depressive disorder, insomnia

TTCCs in Essential Tremor

Essential tremor (ET) is characterized by a 4–12 Hz postural and action tremor, primarily affecting the upper limbs but often accompanied by head and voice tremors. 121 , 122 , 123 Management of tremor in affected patients is complicated by multiple comorbidities and high treatment discontinuations, all of which contribute to high disease burden. 124 Several genetic studies of families with ET suggest that an autosomal‐dominant inheritance pattern is common. 125 , 126 , 127 , 128 Variants in the CACNA1G gene encoding the CaV3.1 isoform have been reported in three families based on whole exome (of 37 families) and whole genome (of eight families) sequencing studies of familial ET; however, corresponding functional studies showed little to no change from wild type. 99 , 129 Clinical and preclinical data suggest the dysregulation of neuronal activity in the CTC network in ET. 18 Within the CTC, TTCCs are highly expressed in deep cerebellar nuclei (DCN) and Purkinje cells (PC), 6 , 130 , 131 ventral thalamic nuclei (particularly the VIM), 132 and motor cortex (Figure 4). 66 Neuronal burst firing has been recorded in the VIM in ET patients undergoing DBS, and the firing pattern correlates with tremor. 18 Furthermore, bilateral or unilateral VIM‐DBS has both been shown to improve tremor in patients with ET. 33 , 34 , 35 The inferior olive (IO) is a related TTCC‐rich 66 component that modulates CTC activity and may also be implicated in tremor generation. 133 , 134 , 135 , 136 While there are suggestions that broader olivo‐cerebello‐thalamo‐cortical network dysfunction exists in ET, 70 the traditional olivary model focusing on a central IO disturbance in ET has been increasingly in question, with limited imaging or postmortem evidence to support its primary role in tremor. 137 On the other hand, the CTC has been at the center of ET pathophysiology, and neurons at each node (cerebellar cortex, DCN, VIM, and motor cortex) are enriched for TTCC expression to modulate the firing patterns in normal and diseased states.

Figure 4.

Figure 4

CTC circuit dysfunction in ET. (A) Key nodes of the CTC circuit (cerebellar cortex, DCN including the dentate nucleus, thalamic VIM, and motor cortex) and the intersecting white matter pathways that connect them. Orange = Guillain Mollaret triangle connecting red nucleus, IO, and dentate nucleus; red = corticospinal tract, blue = dentatothalamic tract, gray = corticopontine tract, black (infratentorial) = pontocerebellar tract, and black (supratentorial) = thalamocortical tract. (B) Tremor generation is mediated by TTCCs in IO neurons which project climbing fibers to PC dendrites within cerebellar cortex, and form synapses on glutamatergic neurons of the DCN, leading to inhibitor responses followed by rebound excitation. TTCCs play a key role in generating subsequent burst firing in the thalamus, with thalamocortical (dashed line) pathways transmitting tremor‐related signals to brainstem and spinal cord via corticopontine and corticospinal tracts. DCN, deep cerebellar nuclei; IO, inferior olive; VIM, ventral intermediate nucleus.

Animal models have provided insights into the mechanisms underlying tremor generation, the most common of which is the harmaline rodent tremor model. 138 Harmaline, a central nervous system (CNS) stimulant, induces acute tremor in a variety of animals, 139 , 140 , 141 with peak tremor frequency varying by species and ranging from 8–10 Hz in monkeys to 10–16 Hz in pigs. 141 , 142 , 143 , 144 Importantly, harmaline induces tremor by augmenting TTCC function. 145 Both knockout and selective knockdown of CACNA1G have been shown to eliminate tremor in a harmaline mouse model, 146 and TTCC blockers have shown efficacy in harmaline‐induced tremor in mice and rats in a dose‐dependent manner. 55 , 56 , 147 Importantly, harmaline studies have contributed critical insights into the role of TTCC‐rich CTC nodes in tremor genesis. Subthreshold oscillations (i.e., neuronal oscillations that occur as a result of intrinsic membrane potential fluctuations but which are below the voltage threshold required to cause a neuron to fire) are dependent on TTCC conductance 148 and closely associated with generation of tremor rhythm, and have been shown to be amplified in IO neurons following harmaline treatment. 146 , 149 , 150 In harmaline‐induced tremor, hyperactive glutamatergic IO neurons recruit medial regions of cerebellar cortex via climbing fibers to PC dendrites, inducing PC burst firing rhythmically and synchronously. 151 , 152 Additionally, harmaline administration results in the expression of c‐fos, a marker for neuronal activation, in PC nuclei in rats, further supporting PC hyperexcitability. 153 Climbing fiber and PC interactions appear to be particularly important in harmaline‐induced tremor. Indeed, pcd mice with some PC degeneration but intact climbing fibers have an attenuated response to harmaline administration, with reduced tremor frequency and amplitude relative to wild‐type mice. 154 In contrast, Lurcher mice which lack PC neurons and climbing fibers do not develop tremor upon harmaline administration. 154 Interestingly, repeated administration of harmaline to rats results in an attenuated tremor response, which may be due to a reduction in PC neuron burst firing and/or degeneration of PC neurons themselves. 155 , 156 , 157 Within the DCN, dysfunction of parvalbumin‐expressing (PV+) excitatory neurons results in action tremor in mice, and DCN synaptic transmission block reverses the tremor. 158 Burst firing of DCN neurons travels to the cerebral cortex via the VIM of the thalamus, 46 with tremor manifesting as a result of rhythmic modulation of primary motor cortex signals that are conveyed to muscles. In line with this, both VIM neuronal bursting 159 , 160 and enhanced rhythmic neuronal activity in the primary motor cortex measured with magnetoencephalography have demonstrated coherence with tremor. 21

The Grid2 dupE3 mouse model is a recently developed tremor model for ET, in which mice have a genetic mutation of the Grid2 gene (exon 3 duplication) that leads to increased protein degradation and reduced GluRδ2 cerebellar expression levels, resembling the GluRδ2 deficiency observed in patients with ET. 22 GluRδ2 is a synaptic organizer for PC, and the reduction of GluRδ2 levels in Grid2 dupE3 mice results in aberrant climbing fibers extending into parallel fiber synaptic territory, 22 mimicking observations in the human postmortem ET cerebellum. 161 Grid2 dupE3 mice develop 20 Hz cerebellar oscillatory activity (coherent with action tremor), which can be dampened by ethanol and first‐line agents used to treat patients with ET, including propranolol and primidone. 22 Notably, IO neurons have different firing patterns, including simple spikes and bursting activity that heavily depends on TTCC, and the burst firing of IO neurons has been shown to be phase‐coupled with tremor in Grid2 dupE3 mice, 22 supporting that TTCC in IO neurons may contribute to tremor. Interestingly, this mouse model may have some translational value since cerebellar oscillatory activity measured by electroencephalography (EEG) has also been described in patients with ET, 22 with evidence suggesting that cerebellar oscillations may be a relevant biomarker of dysregulated cerebellar synaptic organization in some patient subpopulations. 162

Despite the growing body of preclinical evidence, there has been a paucity of studies on the effects of TTCC blockers on human tremor. Moreover, few selective TTCC blockers exist. For example, 1‐octanol and its primary metabolite octanoic acid have shown evidence of tremor reduction in adults with ET. 163 , 164 Though 1‐octanol has been shown to block TTCC currents in vitro, 165 as a long‐chain alcohol, the biologic mechanism of action is likely complex and similar to ethanol. The antiepileptic drug, zonisamide, was an early agent considered for ET due to its TTCC blocking properties; 166 however, it has also been shown to modulate voltage‐gated sodium channels and GABA receptors, 167 , 168 and its clinical efficacy has not been consistently demonstrated. 169 , 170 , 171 , 172 Ethosuximide is another antiepileptic agent with nonselective TTCC blocking properties 173 that has previously shown promise in mouse models of ET, 56 but not in patients. 174 The clinical benefit of these compounds thus remains unclear. In recent years, renewed interest in the development of more selective TTCC blockers as a therapeutic approach in ET has led to various clinical trials investigating the safety and efficacy of this mechanism. 55 , 175 , 176 , 177 , 178 Suvecaltamide (JZP385/CX‐8998) has demonstrated some improvement in tremor and activities of daily living in a Phase 2 proof‐of‐concept study in ET patients, although the primary efficacy endpoint was not met. 175 , 176 Suvecaltamide is currently under evaluation in Phase 2b trials (NCT05122650). Recent reports from a Phase 2 clinical trial of NBI‐827104 indicate lack of tremor severity reduction in adult participants with ET, although it will be important to understand the full pharmacokinetic, efficacy, safety, and tolerability profile of this compound and its impact on activities of daily living. Notably, none of these trials included biomarkers of TTCC blockade; as such, it is difficult to determine whether functional TTCC blockade was achieved. PRAX‐944 is a selective TTCC blocker previously shown on cryo‐electron microscopy to have high binding affinity to the CaV3.1 pore domain. 179 A recent translational study of PRAX‐944 included measurement of sleep spindle‐related activity (sigma band EEG power) during NREM sleep as a biomarker of functional TTCC blockade, and demonstrated robust sigma band reduction (relative to other spectral frequencies) in both rodents and healthy human participants. 55 Furthermore, recent PRAX‐944 Phase 2 results showed statistically significant and clinically meaningful improvements in both tremor severity and activities of daily living in patients with ET. 177 PRAX‐944 is currently under evaluation in a larger Phase 2b trial. 178 These data with more selective TTCC blockers are promising and highlight the importance of biomarkers of TTCC activity in informing selection of doses expected to produce a functional response in relevant circuitry implicated in ET.

TTCCs in Parkinson's Disease

PD is clinically manifested by cardinal symptoms of bradykinesia (slowness of movement), rigidity, or stiffness that impedes normal motor activities, and/or rest tremor. 180 The motor symptoms of PD are frequently disabled and significantly affect daily function. 181 Although the primary pathologic mechanisms driving the neurodegenerative processes in PD have not yet been fully identified, the hypokinetic symptoms, bradykinesia, and rigidity have been linked to the loss of dopamine (DA) neurons within the substantia nigra pars compacta (SNc) and the resulting reduction of striatal DA tone. 182 , 183 , 184 Importantly, DA is a major regulator of the BG motor circuit (Figure 5), and degeneration of the SNc is correlated with changes in neuronal activity in BG circuits. 186 , 187 , 188 , 189 Thus, the mainstay of PD treatment over the last five decades has focused on levodopa as a DA replacement agent to normalize the proposed underlying BG circuit dysfunction. 190

Figure 5.

Figure 5

BG motor circuit dysfunction in PD. (A) Cross‐sectional illustration of BG motor circuit components. (B) In the BG motor circuit, sensorimotor cortex neurons project to the putamen. Dopamine (blue) from the SNc activates the direct (green) pathway which promotes movement in the healthy state via excitatory D1 receptors. The direct pathway projects from the putamen to the GPi, continuing to the thalamus before terminating in motor cortex via thalamocortical projections. (C) Dopamine (blue) also activates the indirect (red) pathway which suppresses movement in the healthy state via inhibitory D2 receptors. The indirect pathway projects from the putamen to the GPe, continuing to the STN, GPi, and thalamus before terminating in motor cortex. In the parkinsonian state, the loss of dopamine from the SNc leads to hypoactivity of the direct pathway and hyperactivity of the indirect pathway, with the imbalance resulting in PD hypokinetic symptoms. (D) The hyperdirect (purple) pathway projects directly from the cortex to the STN. GPe, external globus pallidus; GPi, internal globus pallidus; SNc, substantia nigra compacta; SNr, substantia nigra reticulata; STN, subthalamic nucleus. Adapted from Hashemiyoon, et al. 185 as licensed under CC BY 4.0.

Although effective, long‐term levodopa therapy can eventually lose efficacy or lead to dyskinesia and hallucinations 191 , 192 and contribute to impulsivity, 193 especially when combined with other dopaminergic agents. 194 For PD with tremor or treatment‐refractory dyskinesias, DBS is an effective therapy for minimizing motor fluctuations, 195 , 196 , 197 where it can be placed in one of the three locations depending on indication and surgical preference. As in ET, DBS of the thalamic VIM is commonly used for tremor suppression in PD. However, VIM‐DBS does not address other symptoms of PD, and DBS of the GPi and STN is frequently used with added benefit for bradykinesia and rigidity. 195 , 198 , 199 , 200 , 201 Indeed, the hypothesis that targeting BG circuit dysfunction may improve motor symptoms in patients with PD is supported by the use of GPi‐ and STN‐DBS, with recent studies demonstrating reduction in PD‐associated action and rest tremor with both GPi‐ and STN‐DBS, although the former may require more time than STN‐DBS before observation of maximum tremor suppression. 38 , 199 , 200 , 201 In patients with advanced disease undergoing BG recordings during DBS placement, hyperactivity and increased bursting are consistently observed in the STN and GPi. 202 , 203 , 204 This is in line with observations in nonhuman primates treated with 1,2,3,6‐tetrahydropyridine (MPTP), which produces a DA‐deficient state and parkinsonian‐like symptoms. 27 , 205 Moreover, lesioning or DBS of the STN reduces hypokinetic symptoms in MPTP‐treated nonhuman primates. 20 , 206 Despite DBS being effective in many patients, the high risk of associated perioperative complications 39 provides a strong impetus for developing novel agents capable of pharmacologically recapitulating the effects of DBS in PD.

While much of our understanding of the dysfunctional circuitry in PD to date has come from observations of abnormal bursting activity in primate models or in patients with PD, data comparing the brain activity of patients with PD to healthy participants, or to patients with PD before disease onset, are lacking. In one study comparing advanced and early PD, patients with more advanced PD were shown to have neurons that fire at a higher rate while demonstrating higher overall level of STN background activity. 207 In the absence of neuronal recordings from healthy brain, studies across different disease populations can provide added insight. A study comparing STN activity in patients with PD versus ET (a disease without known involvement of the BG) showed that the mean firing rate in STN neurons was significantly increased in PD compared to ET, with a larger percentage of neurons in PD exhibiting the burst firing phenotype. 208 Additionally, in a comparison between patients with PD and patients with dystonia (a disease also known to involve the BG), mean GPi discharge rate was also higher in patients with PD. 204 Thus, despite the lack of available control data, the combined preclinical and clinical evidence to date support the hypothesis of BG circuit hyperactivity in the pathophysiology of PD.

A specific role for TTCCs in modulating the underlying circuit dysfunction in PD can be inferred from preclinical studies, which allude to TTCC involvement in burst firing within BG circuit nodes, notably the STN. The abnormal bursting activity in the BG circuit associated with PD has been causally linked to TTCC activity in preclinical models. A TTCC blocker, ML218, inhibits TTCC current and consequently reduces low threshold spike and rebound burst activity in STN neurons in ex vivo thalamic slices, and further reverses haloperidol‐induced catalepsy in a rat model of PD. 53 However, ML218 has demonstrated no effect on parkinsonian symptoms in MPTP‐treated monkeys; 52 , 209 in part, due to its sedative effects, which highlights the need for well‐tolerated agents. Further evidence for TTCC involvement in parkinsonian symptoms is seen in the 6‐hydroxydopamine (6‐OHDA) unilateral lesion model, which results in unilateral depletion of striatal DA innervation. In rats and mice, such DA depletion leads to STN neuronal hyperactivity and rebound bursts thought to be mediated by TTCCs. 29 , 210 , 211 , 212 To this end, local infusion of TTCC blockers into the STN of 6‐OHDA lesioned rats, but not L‐type calcium channel blockers, has demonstrated reduced STN bursting activity and improved motor performance. 29 Another study found that STN‐targeted DBS has differential effects on open field locomotor activity (distance moved, rearing, and movement duration) in 6‐OHDA and healthy rats, whereby delivery of constant negative current reduced STN burst activity and eliminated parkinsonian symptoms in 6‐OHDA rats, while delivery of constant positive current increased STN burst activity and induced a parkinsonian state in healthy rats. 54 Altogether, these studies suggest a strong relationship between STN burst firing and parkinsonian symptoms that can be modulated by both TTCC blockade and DBS.

While the electrical underpinnings of bradykinesia and rigidity seem to relate, at least in part, to disrupted TTCC modulation of BG motor circuits, the mechanisms underlying PD tremor are less clear. Preclinical PD tremor research is particularly scarce due to the lack of a robust rest tremor model. 213 Nonetheless, some evidence suggests that aberrant motor circuitry is similarly implicated in tremor genesis in PD. For example, oscillatory activity in the BG has been observed in patients with PD with limb tremor; specifically, neurons with high‐frequency oscillatory activity have been found in the STN and denoted “tremor cells” due to their spike rate correlating with limb tremor frequency. 214 Notably, recordings derived through thalamic DBS exploration in patients with PD have also identified neuronal subpopulations within the ventral nuclear thalamus, 215 VIM, and ventralis oralis posterior in which neuronal activity correlates with tremor as assessed via electromyography sensors. 216 These findings allude to a potential alternative mechanism for PD tremor implicating TTCC‐rich nodes within the CTC circuit. Indeed, overlapping CTC involvement and oscillatory flow between the cerebellum and sensorimotor cortex have been reported in patients with PD and ET. 217 Furthermore, longitudinal functional magnetic resonance imaging in patients with PD has shown increased CTC activity with disease progression, namely greater recruitment of cortical motor‐associated and cerebellar areas. 218 In an attempt to reconcile a BG versus CTC hypothesis in PD tremor, a “dimmer‐switch” model 219 has been proposed, which postulates that the BG circuit initiates the tremor episodes, whereas the CTC circuit propagates and produces the tremor. This hypothesis is consistent with findings from a study comparing the functional connectivity of the BG (GPi, external globus pallidus [GPe], putamen, and caudate) and CTC circuits in patients with PD with tremor and those without; 220 wherein the GPi, GPe, and putamen were activated at the start of tremor episodes, while CTC activity was aligned with tremor amplitude.

As with ET, few studies have focused on the effectiveness of TTCC blockers in PD tremor. In the rat tacrine‐induced tremulous jaw movement model (used to mimic parkinsonian tremor), 221 mibefradil, a non‐CNS‐penetrant TTCC blocker, has demonstrated largely no effect on tremor suppression, while NNC 55‐0396, a CNS‐penetrant TTCC blocker analog, has been shown to reduce tremor. 222 In MPTP monkeys with levodopa‐responsive tremor, ethosuximide, an anticonvulsant with TTCC blocking properties, 111 , 223 has demonstrated tremor suppression. 224 Together, these studies lend support to the hypothesis that motor deficits in PD, including both hypokinetic and tremor symptoms, may be reduced via central TTCC blockade. Ongoing and future clinical efforts investigating the safety and efficacy of novel TTCC blockers in patients with PD will be critical to better establish the plausibility of these mechanistic relationships.

Conclusions and Future Directions

The concept of treating ET and PD as diseases of dysfunctional circuitry represents an important paradigm shift in the management of movement disorders. The development of agents with novel mechanisms of action that can pharmacologically mimic the effects of DBS has the potential to provide much needed alternatives to invasive surgical interventions. While identifying suitable targets for pharmacologic intervention has been challenging, the current mechanistic exploration of ET and PD points to consistent patterns of disrupted circuitry, including cerebellar and thalamic rhythmic oscillations in ET and STN burst firing in PD. Importantly, TTCC activity appears to be implicated in both ET and PD circuit pathophysiology, with TTCC blockade emerging as a promising therapeutic strategy for alleviating the motor symptoms of these movement disorders (Figure 6). Multiple trials have been completed or are now underway to investigate the safety and efficacy of TTCC blockers in ET and PD. Suvecaltamide (JZP385/CX‐8998, NCT03101241), 176 PRAX‐944 (NCT05021978), 177 and NBI‐827104 (NCT04880616) have been assessed in patients with moderate to severe ET. Additional Phase 2 trials in ET are currently underway for suvecaltamide (NCT05122650) and PRAX‐944 (NCT05021991). 178 PRAX‐944 is also being evaluated in patients with PD with motor fluctuations. Future investigations including those aimed at improved selectivity for certain CaV isoforms may lead to the development of treatments targeted to critical neuronal circuit nodes in ET and PD.

Figure 6.

Figure 6

Targets for pharmacologic intervention in ET and PD. (Middle, top to bottom) The VIM has a high concentration of TTCCs and is a common target for DBS placement in patients with ET. In patients with PD, targets for DBS placement include the VIM, STN, or GPi, all areas with high expression of TTCCs. VIM‐DBS is commonly used for tremor suppression in PD, while the GPi and STN are the preferred targets for managing tremor‐dominant PD, with added benefit for bradykinesia and rigidity. Pharmacologic TTCC blockers therefore have the potential to mimic the effects of DBS as an alternative to surgical intervention in ET and PD. (Left) TTCC isoforms (CaV3.1, CaV3.2, and CaV3.3) expressed in olivocerebellar (blue shading) and thalamocortical (pink shading) nodes, with key CTC circuit targets for potential pharmacologic intervention in ET highlighted. Inhibitory and excitatory neurons are denoted by red and blue lines, respectively. (Right) TTCC isoforms (CaV3.1, CaV3.2, and CaV3.3) expressed in BG (green shading) and thalamocortical nodes (pink shading), with key BG circuit targets for potential pharmacologic intervention in PD highlighted. Inhibitory and excitatory neurons are denoted by red and blue lines. DCN, deep cerebellar nuclei; GPe, external globus pallidus; GPi, internal globus pallidus; nRT, nucleus reticularis of the thalamus; IO, inferior olive; SNc, substantia nigra compacta; SNr, substantia nigra reticulata; STN, subthalamic nucleus. Adapted from Park et al. 86 as licensed under CC BY 4.0.

We conclude with some considerations for future investigational efforts. To successfully measure the effect of TTCC blockers on motor symptoms, studies need to be designed carefully to (i) optimize CNS target engagement and drug titration, and (ii) monitor/minimize potential adverse events. Firstly, differences in efficacy among TTCC blockers may reflect differences in CNS penetrance, with some agents more able to cross the blood–brain barrier. Assessing target engagement in response to TTCC blockade via CNS pharmacodynamic biomarkers 55 may therefore be useful for informing dose range finding in clinical trials. Secondly, TTCCs have been linked to hormone secretion, neurotransmitter release, sleep–wake cycles, and feeding behavior, with many other physiologic roles outside the CNS, including vascular and cardiomyocyte functions. 89 , 225 , 226 , 227 , 228 , 229 Thus, attenuation of aberrant neural activity with minimal disruption to normal TTCC activity should be a therapeutic objective when developing novel agents for ET and PD. Relatedly, drug‐target affinity determination will be an important consideration for future drug development efforts. Taken together, these recommendations provide a critical framework for the development of much needed agents capable of targeting the dysfunctional circuitry underlying pathologic conditions such as ET, PD, and beyond.

Author Contributions

Conceptualization and literature review: Lillian G. Matthews, Corey B. Puryear, Susana S. Correia, Sharan Srinivasan, Gabriel M. Belfort, Ming‐Kai Pan, and Sheng‐Han Kuo. Analysis and interpretation: Lillian G. Matthews, Corey B. Puryear, Susana S. Correia, Sharan Srinivasan, Gabriel M. Belfort, Ming‐Kai Pan, and Sheng‐Han Kuo. Visualization: Lillian G. Matthews, Corey B. Puryear, Sharan Srinivasan, Gabriel M. Belfort, and Sheng‐Han Kuo. Writing—original draft: Lillian G. Matthews, Corey B. Puryear, Susana S. Correia, and Sharan Srinivasan. Writing—review and editing: Lillian G. Matthews, Corey B. Puryear, Susana S. Correia, Sharan Srinivasan, Gabriel M. Belfort, Ming‐Kai Pan, and Sheng‐Han Kuo. All authors read and approved the final article.

Funding information

Funding for editorial support was provided by Praxis Precision Medicines. SHK is supported by NINDS #R01 NS104423, NINDS #R01 NS118179, and NINDS #R01 NS124854.

Conflict of Interest Statement

LGM serves as a paid consultant to Praxis Precision Medicines and is a Praxis shareholder. CBP, GMB, and SC were employees of Praxis Precision Medicines during article preparation and are Praxis shareholders. MKP and SHK serve as scientific advisors for Praxis Precision Medicines and Sage Therapeutics.

Supporting information

Figure S1

Acknowledgments

We thank Sheryl Torr‐Brown, Alyssa Theodore, and Lorena Puto of Simpson Healthcare for assistance with preparing the article. The Genotype‐Tissue Expression (GTEx) Project was supported by the Common Fund of the Office of the Director of the National Institutes of Health, and by NCI, NHGRI, NHLBI, NIDA, NIMH, and NINDS. Figure 2 and Figure S1 in this article were obtained from the GTEx Portal on September 9, 2022.

Funding Statement

This work was funded by Praxis Precision Medicines.

References

  • 1. Louis ED, Ottman R. How many people in the USA have essential tremor? Deriving a population estimate based on epidemiological data. Tremor Other Hyperkinet Mov (NY). 2014;4:259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Louis ED, McCreary M. How common is essential tremor? Update on the worldwide prevalence of essential tremor. Tremor Other Hyperkinet Mov (NY). 2021;11:28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Song P, Zhang Y, Zha M, et al. The global prevalence of essential tremor, with emphasis on age and sex: a meta‐analysis. J Glob Health. 2021;11:04028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Marras C, Beck JC, Bower JH, et al. Prevalence of Parkinson's disease across North America. NPJ Parkinsons Dis. 2018;4:21. doi: 10.1038/s41531-018-0058-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Pringsheim T, Jette N, Frolkis A, Steeves TD. The prevalence of Parkinson's disease: a systematic review and meta‐analysis. Mov Disord. 2014;29(13):1583‐1590. doi: 10.1002/mds.25945 [DOI] [PubMed] [Google Scholar]
  • 6. Louis ED, Faust PL, Vonsattel JPG, et al. Neuropathological changes in essential tremor: 33 cases compared with 21 controls. Brain. 2007;130(12):3297‐3307. [DOI] [PubMed] [Google Scholar]
  • 7. Axelrad JE, Louis ED, Honig LS, et al. Reduced Purkinje cell number in essential tremor: a postmortem study. Arch Neurol. 2008;65(1):101‐107. doi: 10.1001/archneurol.2007.8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Greffard S, Verny M, Bonnet AM, et al. Motor score of the unified Parkinson disease rating scale as a good predictor of Lewy body‐associated neuronal loss in the substantia nigra. Arch Neurol. 2006;63(4):584‐588. doi: 10.1001/archneur.63.4.584 [DOI] [PubMed] [Google Scholar]
  • 9. Louis ED, Lee M, Babij R, et al. Reduced Purkinje cell dendritic arborization and loss of dendritic spines in essential tremor. Brain. 2014;137(Pt 12):3142‐3148. doi: 10.1093/brain/awu314 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Choe M, Cortés E, Vonsattel JP, Kuo SH, Faust PL, Louis ED. Purkinje cell loss in essential tremor: random sampling quantification and nearest neighbor analysis. Mov Disord. 2016;31(3):393‐401. doi: 10.1002/mds.26490 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. Lang AE, Siderowf AD, Macklin EA, et al. Trial of cinpanemab in early Parkinson's disease. N Engl J Med. 2022;387(5):408‐420. doi: 10.1056/NEJMoa2203395 [DOI] [PubMed] [Google Scholar]
  • 12. Pagano G, Taylor KI, Anzures‐Cabrera J, et al. Trial of prasinezumab in early‐stage Parkinson's disease. N Engl J Med. 2022;387(5):421‐432. doi: 10.1056/NEJMoa2202867 [DOI] [PubMed] [Google Scholar]
  • 13. Jeanmonod D, Magnin M, Morel A, et al. Thalamocortical dysrhythmia II: clinical and surgical aspects. Thalamus Related Syst. 2001;1(3):245‐254. [Google Scholar]
  • 14. Llinás R, Ribary U, Jeanmonod D, et al. Thalamocortical dysrhythmia I: functional and imaging aspects. Thalamus Related Syst. 2001;1(3):237‐244. [Google Scholar]
  • 15. Llinás RR, Ribary U, Jeanmonod D, Kronberg E, Mitra PP. Thalamocortical dysrhythmia: a neurological and neuropsychiatric syndrome characterized by magnetoencephalography. Proc Natl Acad Sci U S A. 1999;96(26):15222‐15227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Lenz FA, Kwan HC, Dostrovsky JO, Tasker RR. Characteristics of the bursting pattern of action potentials that occurs in the thalamus of patients with central pain. Brain Res. 1989;496(1–2):357‐360. [DOI] [PubMed] [Google Scholar]
  • 17. Fremont R, Tewari A, Khodakhah K. Aberrant Purkinje cell activity is the cause of dystonia in a shRNA‐based mouse model of rapid onset dystonia‐parkinsonism. Neurobiol Dis. 2015;82:200‐212. doi: 10.1016/j.nbd.2015.06.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Milosevic L, Kalia SK, Hodaie M, Lozano AM, Popovic MR, Hutchison WD. Physiological mechanisms of thalamic ventral intermediate nucleus stimulation for tremor suppression [published correction appears in Brain 2018;141(9):e72]. Brain. 2018;141(7):2142‐2155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Mitchell IJ, Clarke CE, Boyce S, et al. Neural mechanisms underlying parkinsonian symptoms based upon regional uptake of 2‐deoxyglucose in monkeys exposed to 1‐methyl‐4‐phenyl‐1,2,3,6‐tetrahydropyridine. Neuroscience. 1989;32(1):213‐226. [DOI] [PubMed] [Google Scholar]
  • 20. Bergman H, Wichmann T, DeLong MR. Reversal of experimental parkinsonism by lesions of the subthalamic nucleus. Science. 1990;249(4975):1436‐1438. [DOI] [PubMed] [Google Scholar]
  • 21. Schnitzler A, Münks C, Butz M, Timmermann L, Gross J. Synchronized brain network associated with essential tremor as revealed by magnetoencephalography. Mov Disord. 2009;24(11):1629‐1635. [DOI] [PubMed] [Google Scholar]
  • 22. Pan MK, Li YS, Wong SB, et al. Cerebellar oscillations driven by synaptic pruning deficits of cerebellar climbing fibers contribute to tremor pathophysiology. Sci Transl Med. 2020;12(526):eaay1769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23. Connolly AT, Bajwa JA, Johnson MD. Cortical magnetoencephalography of deep brain stimulation for the treatment of postural tremor. Brain Stimul. 2012;5(4):616‐624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Louis ED, Vonsattel JP, Honig LS, et al. Essential tremor associated with pathologic changes in the cerebellum. Arch Neurol. 2006;63(8):1189‐1193. [DOI] [PubMed] [Google Scholar]
  • 25. Reck C, Florin E, Wojtecki L, et al. Characterisation of tremor‐associated local field potentials in the subthalamic nucleus in Parkinson's disease. Eur J Neurosci. 2009;29(3):599‐612. [DOI] [PubMed] [Google Scholar]
  • 26. Hirschmann J, Hartmann CJ, Butz M, et al. direct relationship between oscillatory subthalamic nucleus‐cortex coupling and rest tremor in Parkinson's disease. Brain. 2013;136(Pt 12):3659‐3670. [DOI] [PubMed] [Google Scholar]
  • 27. Bergman H, Wichmann T, Karmon B, DeLong MR. The primate subthalamic nucleus II: neuronal activity in the MPTP model of parkinsonism. J Neurophysiol. 1994;72(2):507‐520. [DOI] [PubMed] [Google Scholar]
  • 28. Leblois A, Meissner W, Bioulac B, Gross CE, Hansel D, Boraud T. Late emergence of synchronized oscillatory activity in the pallidum during progressive parkinsonism. Eur J Neurosci. 2007;26(6):1701‐1713. [DOI] [PubMed] [Google Scholar]
  • 29. Tai CH, Yang YC, Pan MK, Huang CS, Kuo CC. Modulation of subthalamic T‐type Ca(2+) channels remedies locomotor deficits in a rat model of Parkinson disease. J Clin Invest. 2011;121(8):3289‐3305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Pan MK, Tai CH, Liu WC, Pei JC, Lai WS, Kuo CC. Deranged NMDAergic cortico‐subthalamic transmission underlies parkinsonian motor deficits. J Clin Invest. 2014;124(10):4629‐4641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Sharott A, Gulberti A, Zittel S, et al. Activity parameters of subthalamic nucleus neurons selectively predict motor symptom severity in Parkinson's disease. J Neurosci. 2014;34(18):6273‐6285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Pan MK, Kuo SH, Tai CH, et al. Neuronal firing patterns outweigh circuitry oscillations in parkinsonian motor control. J Clin Invest. 2016;126(12):4516‐4526. doi: 10.1172/JCI88170 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Benabid AL, Pollak P, Gervason C, et al. Long‐term suppression of tremor by chronic stimulation of the ventral intermediate thalamic nucleus. Lancet. 1991;337(8738):403‐406. [DOI] [PubMed] [Google Scholar]
  • 34. Huss DS, Dallapiazza RF, Shah BB, Harrison MB, Diamond J, Elias WJ. Functional assessment and quality of life in essential tremor with bilateral or unilateral DBS and focused ultrasound thalamotomy. Mov Disord. 2015;30(14):1937‐1943. [DOI] [PubMed] [Google Scholar]
  • 35. Pahwa R, Lyons KE, Wilkinson SB, et al. Long‐term evaluation of deep brain stimulation of the thalamus. J Neurosurg. 2006;104(4):506‐512. [DOI] [PubMed] [Google Scholar]
  • 36. Jankovic J, Cardoso F, Grossman RG, Hamilton WJ. Outcome after stereotactic thalamotomy for parkinsonian, essential, and other types of tremor. Neurosurgery. 1995;37(4):680‐686. [DOI] [PubMed] [Google Scholar]
  • 37. Goldman MS, Ahlskog JE, Kelly PJ. The symptomatic and functional outcome of stereotactic thalamotomy for medically intractable essential tremor. J Neurosurg. 1992;76(6):924‐928. [DOI] [PubMed] [Google Scholar]
  • 38. Krack P, Benazzouz A, Pollak P, et al. Treatment of tremor in Parkinson's disease by subthalamic nucleus stimulation. Mov Disord. 1998;13(6):907‐914. [DOI] [PubMed] [Google Scholar]
  • 39. Zarzycki MZ, Domitrz I. Stimulation‐induced side effects after deep brain stimulation ‐ a systematic review. Acta Neuropsychiatr. 2020;32(2):57‐64. [DOI] [PubMed] [Google Scholar]
  • 40. Lu G, Luo L, Liu M, et al. Outcomes and adverse effects of deep brain stimulation on the ventral intermediate nucleus in patients with essential tremor. Neural Plast. 2020;2020:2486065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Paschen S, Forstenpointner J, Becktepe J, et al. Long‐term efficacy of deep brain stimulation for essential tremor: an observer‐blinded study. Neurology. 2019;92(12):e1378‐e1386. [DOI] [PubMed] [Google Scholar]
  • 42. Enatsu R, Kitagawa M, Morishita T, et al. Effect of cycling thalamosubthalamic stimulation on tremor habituation and rebound in Parkinson disease. World Neurosurg. 2020;12(144):64‐67. [DOI] [PubMed] [Google Scholar]
  • 43. Louis ED, Rohl B, Rice C. Defining the treatment gap: what essential tremor patients want that they are not getting. Tremor Other Hyperkinet Mov (NY). 2015;5:331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Vickstrom CR, Liu X, Zhang Y, et al. T‐type calcium channels contribute to burst firing in a subpopulation of medial habenula neurons. eNeuro. 2020;7(4):ENEURO.0201‐20.2020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Deschênes M, Paradis M, Roy JP, Steriade M. Electrophysiology of neurons of lateral thalamic nuclei in cat: resting properties and burst discharges. J Neurophysiol. 1984;51(6):1196‐1219. doi: 10.1152/jn.1984.51.6.1196 [DOI] [PubMed] [Google Scholar]
  • 46. Llinás R, Jahnsen H. Electrophysiology of mammalian thalamic neurones in vitro. Nature. 1982;297(5865):406‐408. doi: 10.1038/297406a0 [DOI] [PubMed] [Google Scholar]
  • 47. Molineux ML, McRory JE, McKay BE, et al. Specific T‐type calcium channel isoforms are associated with distinct burst phenotypes in deep cerebellar nuclear neurons. Proc Natl Acad Sci U S A. 2006;103(14):5555‐5560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Perez‐Reyes E. Molecular physiology of low‐voltage‐activated t‐type calcium channels. Physiol Rev. 2003;83(1):117‐161. [DOI] [PubMed] [Google Scholar]
  • 49. Lauro PM, Lee S, Akbar U, Asaad WF. Subthalamic‐cortical network reorganization during Parkinson's tremor. J Neurosci. 2021;41(47):9844‐9858. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Beurrier C, Congar P, Bioulac B, Hammond C. Subthalamic nucleus neurons switch from single‐spike activity to burst‐firing mode. J Neurosci. 1999;19(2):599‐609. doi: 10.1523/JNEUROSCI.19-02-00599.1999 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Song WJ, Baba Y, Otsuka T, Murakami F. Characterization of Ca(2+) channels in rat subthalamic nucleus neurons. J Neurophysiol. 2000;84(5):2630‐2637. doi: 10.1152/jn.2000.84.5.263 [DOI] [PubMed] [Google Scholar]
  • 52. Devergnas A, Chen E, Ma Y, et al. Anatomical localization of Cav3.1 calcium channels and electrophysiological effects of T‐type calcium channel blockade in the motor thalamus of MPTP‐treated monkeys. J Neurophysiol. 2016;115(1):470‐485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53. Xiang Z, Thompson AD, Brogan JT, et al. The discovery and characterization of ML218: a novel, centrally active T‐type calcium channel inhibitor with robust effects in STN neurons and in a rodent model of Parkinson's disease. ACS Chem Nerosci. 2011;2(12):730‐742. doi: 10.1021/cn200090z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Tai CH, Pan MK, Lin JJ, Huang CS, Yang YC, Kuo CC. Subthalamic discharges as a causal determinant of parkinsonian motor deficits. Ann Neurol. 2012;72(3):464‐476. [DOI] [PubMed] [Google Scholar]
  • 55. Scott L, Puryear CB, Belfort GM, et al. Translational pharmacology of PRAX‐944, a novel T‐type calcium channel blocker in development for the treatment of essential tremor. Mov Disord. 2022;37(6):1193‐1201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Handforth A, Homanics GE, Covey DF, et al. T‐type calcium channel antagonists suppress tremor in two mouse models of essential tremor. Neuropharmacology. 2010;59(6):380‐387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Catterall WA, Perez‐Reyes E, Snutch TP, Striessnig J. International Union of Pharmacology. XLVIII. Nomenclature and structure‐function relationships of voltage‐gated calcium channels. Pharmacol Rev. 2005;57(4):411‐425. [DOI] [PubMed] [Google Scholar]
  • 58. Powell KL, Cain SM, Snutch TP, O'Brien TJ. Low threshold T‐type calcium channels as targets for novel epilepsy treatments. Br J Clin Pharmacol. 2014;77(5):729‐739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. McCormick DA, Huguenard JR. A model of the electrophysiological properties of thalamocortical relay neurons. J Neurophysiol. 1992;68:1384‐1400. [DOI] [PubMed] [Google Scholar]
  • 60. Chevalier M, Lory P, Mironneau C, Macrez N, Quignard JF. T‐type CaV3.3 calcium channels produce spontaneous low‐threshold action potentials and intracellular calcium oscillations. Eur J Neurosci. 2006;23:2321‐2329. [DOI] [PubMed] [Google Scholar]
  • 61. Dreyfus FM, Tscherter A, Errington AC, et al. Selective T‐type calcium channel block in thalamic neurons reveals channel redundancy and physiological impact of I(T)window. J Neurosci. 2010;30(1):99‐109. doi: 10.1523/JNEUROSCI.4305-09.2010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Perez‐Reyes E, Cribbs LL, Daud A, et al. Molecular characterization of a neuronal low‐voltage‐activated T‐type calcium channel. Nature. 1998;391(6670):896‐900. [DOI] [PubMed] [Google Scholar]
  • 63. Cribbs LL, Lee JH, Yang J, et al. Cloning and characterization of alpha1H from human heart, a member of the T‐type Ca2+ channel gene family. Circ Res. 1998;83(1):103‐109. [DOI] [PubMed] [Google Scholar]
  • 64. Lee JH, Daud AN, Cribbs LL, et al. Cloning and expression of a novel member of the low voltage‐activated T‐type calcium channel family. J Neurosci. 1999;19(6):1912‐1921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Monteil A, Chemin J, Bourinet E, Mennessier G, Lory P, Nargeot J. Molecular and functional properties of the human alpha(1G) subunit that forms T‐type calcium channels. J Biol Chem. 2000;275(9):6090‐6100. [DOI] [PubMed] [Google Scholar]
  • 66. Talley EM, Cribbs LL, Lee JH, Daud A, Perez‐Reyes E, Bayliss DA. Differential distribution of three members of a gene family encoding low voltage‐activated (T‐type) calcium channels. J Neurosci. 1999;19(6):1895‐1911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67. Craig PJ, Beattie RE, Folly EA, et al. Distribution of the voltage‐dependent calcium channel alpha1G subunit mRNA and protein throughout the mature rat brain. Eur J Neurosci. 1999;11(8):2949‐2964. [DOI] [PubMed] [Google Scholar]
  • 68. Klöckner U, Lee JH, Cribbs LL, et al. Comparison of the Ca2 + currents induced by expression of three cloned alpha1 subunits, alpha1G, alpha1H and alpha1I, of low‐voltage‐activated T‐type Ca2 + channels. Eur J Neurosci. 1999;11(12):4171‐4178. [DOI] [PubMed] [Google Scholar]
  • 69. Weiss N, Zamponi GW. Genetic T‐type calcium channelopathies. J Med Genet. 2020;57(1):1‐10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70. Shukla AW. Reduction of neuronal hyperexcitability with modulation of T‐type calcium channel or SK channel in essential tremor. Int Rev Neurobiol. 2022;163:335‐355. doi: 10.1016/bs.irn.2022.02.008 [DOI] [PubMed] [Google Scholar]
  • 71. The Genotype‐Tissue Expression (GTEx) Project . GTEx Portal. V8. Accessed September 9, 2022. https://gtexportal.org/home/.
  • 72. Dataset: Allen Institute for Brain Science (2004) . Allen Mouse Brain Atlas [dataset]. mouse.brain‐map.org. RRID:SCR_002978; Primary publication: Lein ES, Hawrylycz MJ, Ao N, et al. Genome‐wide atlas of gene expression in the adult mouse brain. Nature. 2007;445(7124):168–176. doi: 10.1038/nature05453. [DOI] [PubMed]
  • 73. Allen Reference Atlas – Mouse Brain [brain atlas] . atlas.brain‐map.org.
  • 74. Dataset: Allen Institute for Brain Science (2010) . Allen Human Brain Atlas: Microarray [dataset]. human.brain‐map.org. RRID:SCR_007416; Primary publication: Hawrylycz MJ, Lein ES, Guillozet‐Bongaarts AL, et al. An anatomically comprehensive atlas of the adult human transcriptome. Nature, 2012;489(7416), 391–399. doi: 10.1038/nature11405 [DOI] [PMC free article] [PubMed]
  • 75. Allen Reference Atlas – Human Brain [brain atlas] . atlas.brain‐map.org.
  • 76. McCormick DA, Pape HC. Properties of a hyperpolarization‐activated cation current and its role in rhythmic oscillation in thalamic relay neurones. J Physiol. 1990;431:291‐318. doi: 10.1113/jphysiol.1990.sp018331 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Aguado C, García‐Madrona S, Gil‐Minguez M, Luján R. Ontogenic changes and differential localization of T‐type Ca(2+) channel subunits Cav3.1 and Cav3.2 in mouse hippocampus and cerebellum. Front Neuroanat. 2016;10:83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78. Bernal Sierra YA, Haseleu J, Kozlenkov A, Bégay V, Lewin GR. Genetic tracing of Cav3.2 T‐type calcium channel expression in the peripheral nervous system. Front Mol Neurosci. 2017;10:70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79. Wang G, Bochorishvili G, Chen Y, et al. CaV3.2 calcium channels control NMDA receptor‐mediated transmission: a new mechanism for absence epilepsy. Genes Dev. 2015;29(14):1535‐1551. doi: 10.1101/gad.260869.115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80. Choi S, Na HS, Kim J, et al. Attenuated pain responses in mice lacking Ca(V)3.2 T‐type channels. Genes Brain Behav. 2007;6(5):425‐431. doi: 10.1111/j.1601-183X.2006.00268.x [DOI] [PubMed] [Google Scholar]
  • 81. Candelas M, Reynders A, Arango‐Lievano M, et al. Cav3.2 T‐type calcium channels shape electrical firing in mouse lamina II neurons. Sci Rep. 2019;9(1):3112. doi: 10.1038/s41598-019-39703-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82. Astori S, Wimmer RD, Prosser HM, et al. The Ca(V)3.3 calcium channel is the major sleep spindle pacemaker in thalamus. Proc Natl Acad Sci U S A. 2011;108(33):13823‐13828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83. Pellegrini C, Lecci S, Lüthi A, Astori S. Suppression of sleep spindle rhythmogenesis in mice with deletion of CaV3.2 and CaV3.3 T‐type Ca(2+) channels. Sleep. 2016;39(4):875‐885. doi: 10.5665/sleep.5646 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84. Anderson MP, Mochizuki T, Xie J, et al. Thalamic Cav3.1 T‐type Ca2+ channel plays a crucial role in stabilizing sleep. Proc Natl Acad Sci U S A. 2005;102(5):1743‐1748. doi: 10.1073/pnas.0409644102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85. Lee J, Kim D, Shin HS. 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. 2004;101(52):18195‐18199. doi: 10.1073/pnas.0408089101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Park YG, Kim J, Kim D. The potential roles of T‐type Ca2+ channels in motor coordination. Front Neural Circuits. 2013;7:172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87. Chemin J, Siquier‐Pernet K, Nicouleau M, et al. De novo mutation screening in childhood‐onset cerebellar atrophy identifies gain‐of‐function mutations in the CACNA1G calcium channel gene. Brain. 2018;141:1998‐2013. [DOI] [PubMed] [Google Scholar]
  • 88. Lory P, Nicole S, Monteil A. Neuronal Cav3 channelopathies: recent progress and perspectives. Pflugers Arch. 2020;472(7):831‐844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Zamponi GW, Striessnig J, Koschak A, Dolphin AC. The physiology, pathology, and pharmacology of voltage‐gated calcium channels and their future therapeutic potential. Pharmacol Rev. 2015;67(4):821‐870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90. Gerke MB, Duggan AW, Xu L, Siddall PJ. Thalamic neuronal activity in rats with mechanical allodynia following contusive spinal cord injury. Neuroscience. 2003;117(3):715‐722. [DOI] [PubMed] [Google Scholar]
  • 91. Iwata M, LeBlanc BW, Kadasi LM, et al. High‐frequency stimulation in the ventral posterolateral thalamus reverses electrophysiologic changes and hyperalgesia in a rat model of peripheral neuropathic pain. Pain. 2011;152(11):2505‐2513. [DOI] [PubMed] [Google Scholar]
  • 92. Walton KD, Llinás RR. Central pain as a thalamocortical dysrhythmia: a thalamic efference disconnection? In: Kruger L, Light AR, eds. Translational Pain Research: from Mouse to Man. CRC Press/Taylor & Francis; 2010. [PubMed] [Google Scholar]
  • 93. Sarnthein J, Jeanmonod D. High thalamocortical theta coherence in patients with neurogenic pain. Neuroimage. 2008;39(4):1910‐1917. doi: 10.1016/j.neuroimage.2007.10.019 [DOI] [PubMed] [Google Scholar]
  • 94. Sarnthein J, Stern J, Aufenberg C, Rousson V, Jeanmonod D. Increased EEG power and slowed dominant frequency in patients with neurogenic pain. Brain. 2006;129(Pt 1):55‐64. doi: 10.1093/brain/awh631 [DOI] [PubMed] [Google Scholar]
  • 95. Amir R, Michaelis M, Devor M. Burst discharge in primary sensory neurons: triggered by subthreshold oscillations, maintained by depolarizing afterpotentials. J Neurosci. 2002;22(3):1187‐1198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Pouille F, Cavelier P, Desplantez T, et al. Dendro‐somatic distribution of calcium‐mediated electrogenesis in purkinje cells from rat cerebellar slice cultures. J Physiol. 2000;527(Pt 2):265‐282. doi: 10.1111/j.1469-7793.2000.00265.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97. Swensen AM, Bean BP. Ionic mechanisms of burst firing in dissociated Purkinje neurons. J Neurosci. 2003;23(29):9650‐9663. doi: 10.1523/JNEUROSCI.23-29-09650.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Womack MD, Khodakhah K. Dendritic control of spontaneous bursting in cerebellar Purkinje cells. J Neurosci. 2004;24(14):3511‐3521. doi: 10.1523/JNEUROSCI.0290-04.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99. Odgerel Z, Sonti S, Hernandez N, et al. Whole genome sequencing and rare variant analysis in essential tremor families. PLoS One. 2019;14(8):e0220512. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Casas‐Alba D, López‐Sala L, Pérez‐Ordóñez M, et al. Early‐onset severe spinocerebellar ataxia 42 with neurodevelopmental deficits (SCA42ND): case report, pharmacological trial, and literature review. Am J Med Genet A. 2021;185(1):256‐260. doi: 10.1002/ajmg.a.61939 [DOI] [PubMed] [Google Scholar]
  • 101. Kimura M, Yabe I, Hama Y, et al. SCA42 mutation analysis in a case series of Japanese patients with spinocerebellar ataxia. J Hum Genet. 2017;62(9):857‐859. [DOI] [PubMed] [Google Scholar]
  • 102. Hashiguchi S, Doi H, Kunii M, et al. Ataxic phenotype with altered CaV3.1 channel property in a mouse model for spinocerebellar ataxia 42. Neurobiol Dis. 2019;130:104516. doi: 10.1016/j.nbd.2019.104516 [DOI] [PubMed] [Google Scholar]
  • 103. Morino H, Matsuda Y, Muguruma K, et al. A mutation in the low voltage‐gated calcium channel CACNA1G alters the physiological properties of the channel, causing spinocerebellar ataxia. Mol Brain. 2015;8:89. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104. Li X, Zhou C, Cui L, et al. A case of a novel CACNA1G mutation from a Chinese family with SCA42: a case report and literature review. Medicine (Baltimore). 2018;97(36):e12148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105. Coutelier M, Blesneac I, Monteil A, et al. A recurrent mutation in CACNA1G alters Cav3.1 T‐type calcium‐channel conduction and causes autosomal‐dominant cerebellar ataxia. Am J Hum Genet. 2015;97(5):726‐737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106. Barresi S, Dentici ML, Manzoni F, et al. Infantile‐onset syndromic cerebellar ataxia and CACNA1G mutations. Pediatr Neurol. 2020;104:40‐45. [DOI] [PubMed] [Google Scholar]
  • 107. Splawski I, Yoo DS, Stotz SC, Cherry A, Clapham DE, Keating MT. CACNA1H mutations in autism spectrum disorders. J Biol Chem. 2006;281(31):22085‐22091. [DOI] [PubMed] [Google Scholar]
  • 108. Gulsuner S, Walsh T, Watts AC, et al. Consortium on the genetics of schizophrenia (COGS), PAARTNERS study group. Spatial and temporal mapping of de novo mutations in schizophrenia to a fetal prefrontal cortical network. Cell. 2013;154:518‐529. doi: 10.1016/j.cell.2013.06.049 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109. Schizophrenia Working Group of the Psychiatric Genomics Consortium . Biological insights from 108 schizophrenia‐associated genetic loci. Nature. 2014;511(7510):421‐427. doi: 10.1038/nature13595 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. Andrade A, Hope J, Allen A, Yorgan V, Lipscombe D, Pan JQ. A rare schizophrenia risk variant of CACNA1I disrupts CaV3.3 channel activity. Sci Rep. 2016;6:6. doi: 10.1038/srep34233 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111. Gomora JC, Daud AN, Weiergraber M, Perez‐Reyes E. Block of cloned human T‐type calcium channels by succinimide antiepileptic drugs. Mol Pharmacol. 2001;60(5):1121‐1132. [PubMed] [Google Scholar]
  • 112. Matar N, Jin W, Wrubel H, Hescheler J, Schneider T, Weiergräber M. Zonisamide block of cloned human T‐type voltage‐gated calcium channels. Epilepsy Res. 2009;83(2–3):224‐234. doi: 10.1016/j.eplepsyres.2008.11.010 [DOI] [PubMed] [Google Scholar]
  • 113. Zhang YF, Gibbs JW 3rd, Coulter DA. Anticonvulsant drug effects on spontaneous thalamocortical rhythms in vitro: ethosuximide, trimethadione, and dimethadione. Epilepsy Res. 1996;23(1):15‐36. doi: 10.1016/0920-1211(95)00079-8 [DOI] [PubMed] [Google Scholar]
  • 114. Mattson RH, Cramer JA, Williamson PD, Novelly RA. Valproic acid in epilepsy: clinical and pharmacological effects. Ann Neurol. 1978;3:20‐25. doi: 10.1002/ana.410030105 [DOI] [PubMed] [Google Scholar]
  • 115. Bergson P, Lipkind G, Lee SP, Duban ME, Hanck DA. Verapamil block of T‐type calcium channels. Mol Pharmacol. 2011;79(3):411‐419. doi: 10.1124/mol.110.069492 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116. Kopecky BJ, Liang R, Bao J. T‐type calcium channel blockers as neuroprotective agents. Pflugers Arch. 2014;466(4):757‐765. doi: 10.1007/s00424-014-1454-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117. Bancila M, Copin JC, Daali Y, Schatlo B, Gasche Y, Bijlenga P. Two structurally different T‐type Ca2+ channel inhibitors, mibefradil and pimozide, protect CA1 neurons from delayed death after global ischemia in rats. Fundam Clin Pharmacol. 2011;25(4):469‐478. doi: 10.1111/j.1472-8206.2010.00879.x [DOI] [PubMed] [Google Scholar]
  • 118. Traboulsie A, Chemin J, Kupfer E, Nargeot J, Lory P. T‐type calcium channels are inhibited by fluoxetine and its metabolite norfluoxetine. Mol Pharmacol. 2006;69(6):1963‐1968. doi: 10.1124/mol.105.020842 [DOI] [PubMed] [Google Scholar]
  • 119. Deák F, Lasztóczi B, Pacher P, Petheö GL, Kecskeméti V, Spät A. Inhibition of voltage‐gated calcium channels by fluoxetine in rat hippocampal pyramidal cells. Neuropharmacology. 2000;39(6):1029‐1036. doi: 10.1016/s0028-3908(99)00206-3 [DOI] [PubMed] [Google Scholar]
  • 120. Kraus RL, Li Y, Jovanovska A, Renger JJ. Trazodone inhibits T‐type calcium channels. Neuropharmacology. 2007;53(2):308‐317. doi: 10.1016/j.neuropharm.2007.05.011 [DOI] [PubMed] [Google Scholar]
  • 121. Bhatia KP, Bain P, Bajaj N, et al. Consensus statement on the classification of tremors. From the task force on tremor of the International Parkinson and Movement Disorder Society. Mov Disord. 2018;33(1):75‐87. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122. Brennan KC, Jurewicz EC, Ford B, Pullman SL, Louis ED. Is essential tremor predominantly a kinetic or a postural tremor? A clinical and electrophysiological study. Mov Disord. 2002;17(2):313‐316. [DOI] [PubMed] [Google Scholar]
  • 123. Louis ED. Clinical practice. Essential tremor. N Engl J Med. 2001;345(12):887‐891. doi: 10.1056/NEJMcp010928 [DOI] [PubMed] [Google Scholar]
  • 124. Vetterick C, Lyons KE, Matthews LG, Pendal R, Ravina B. The hidden burden of disease and treatment experiences of patients with essential tremor: A retrospective claims data analysis. Adv Ther. 2022;39(12):5546‐5567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125. Gulcher JR, Jónsson P, Kong A, et al. Mapping of a familial essential tremor gene, FET1, to chromosome 3q13. Nat Genet. 1997;17(1):84‐87. [DOI] [PubMed] [Google Scholar]
  • 126. Higgins JJ, Pho LT, Nee LE. A gene (ETM) for essential tremor maps to chromosome 2p22‐p25. Mov Disord. 1997;12(6):859‐864. [DOI] [PubMed] [Google Scholar]
  • 127. Kovach MJ, Ruiz J, Kimonis K, et al. Genetic heterogeneity in autosomal dominant essential tremor. Genet Med. 2001;3(3):197‐199. [DOI] [PubMed] [Google Scholar]
  • 128. Shatunov A, Sambuughin N, Jankovic J, et al. Genomewide scans in north American families reveal genetic linkage of essential tremor to a region on chromosome 6p23. Brain. 2006;129(Pt 9):2318‐2331. [DOI] [PubMed] [Google Scholar]
  • 129. Liu X, Hernandez N, Kisselev S, et al. Identification of candidate genes for familial early‐onset essential tremor. Eur J Hum Genet. 2016;24(7):1009‐1015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130. Ly R, Bouvier G, Schonewille M, et al. T‐type channel blockade impairs long‐term potentiation at the parallel fiber‐Purkinje cell synapse and cerebellar learning. Proc Natl Acad Sci U S A. 2013;110(50):20302‐20307. doi: 10.1073/pnas.1311686110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131. Cavelier P, Lohof AMB, Lonchamp E, et al. Participation of low‐threshold Ca2+ spike in the Purkinje cells complex spike. Neuroreport. 2008;19(3):299‐303. doi: 10.1097/WNR.0b013e3282f4cb96 [DOI] [PubMed] [Google Scholar]
  • 132. Pedrosa DJ, Reck C, Florin E, et al. Essential tremor and tremor in Parkinson's disease are associated with distinct ‘tremor clusters’ in the ventral thalamus. Exp Neurol. 2012;237(2):435‐443. [DOI] [PubMed] [Google Scholar]
  • 133. Beitz AJ, Saxon D. Harmaline‐induced climbing fiber activation causes amino acid and peptide release in the rodent cerebellar cortex and a unique temporal pattern of Fos expression in the olivo‐cerebellar pathway. J Neurocytol. 2004;33:49‐74. [DOI] [PubMed] [Google Scholar]
  • 134. Hallett M, Dubinsky RM. Glucose metabolism in the brain of patients with essential tremor. J Neurol Sci. 1993;114(1):45‐48. [DOI] [PubMed] [Google Scholar]
  • 135. Boecker H, Wills AJ, Ceballos‐Baumann A, et al. The effect of ethanol on alcohol‐responsive essential tremor: a positron emission tomography study. Ann Neurol. 1996;39(5):650‐658. [DOI] [PubMed] [Google Scholar]
  • 136. Simantov R, Snyder SH, Oster‐Granite ML. Harmaline‐induced tremor in the rat: abolition by 3‐acetylpyridine destruction of cerebellar climbing fibers. Brain Res. 1976;114(1):144‐151. [DOI] [PubMed] [Google Scholar]
  • 137. Louis ED, Lenka A. The olivary hypothesis of essential tremor: time to lay this model to rest? Tremor Other Hyperkinet Mov (N Y). 2017;7:473. doi: 10.7916/D8FF40RX [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138. Martin FC, Le AT, Handforth A. Harmaline‐induced tremor as a potential preclinical screening method for essential tremor medications. Mov Disord. 2005;20:298‐305. [DOI] [PubMed] [Google Scholar]
  • 139. De Montigny C, Lamarre Y. Effects produced by local applications of harmaline in the inferior olive. Can J Physiol Pharmacol. 1975;53(5):845‐849. [DOI] [PubMed] [Google Scholar]
  • 140. Wilms H, Sievers J, Deuschl G. Animal models of tremor. Mov Disord. 1999;14(4):557‐571. [DOI] [PubMed] [Google Scholar]
  • 141. Lee J, Kim I, Lee J, et al. Development of harmaline‐induced tremor in a swine model. Tremor Other Hyperkinet Mov (N Y). 2018;8:532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142. Miwa H. Rodent models of tremor. Cerebellum. 2007;6(1):66‐72. doi: 10.1080/14734220601016080 [DOI] [PubMed] [Google Scholar]
  • 143. Lamarre Y, Mercier LA. Neurophysiological studies of harmaline‐induced tremor in the cat. Can J Physiol Pharmacol. 1971;49(12):1049‐1058. [DOI] [PubMed] [Google Scholar]
  • 144. Kosmowska B, Wardas J. The pathophysiology and treatment of essential tremor: the role of adenosine and dopamine receptors in animal models. Biomolecules. 2021;11(12):1813. doi: 10.3390/biom11121813 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145. Handforth A. Harmaline tremor: underlying mechanisms in a potential animal model of essential tremor. Tremor Other Hyperkinet Mov (NY). 2012;2:02‐92‐769‐1. doi: 10.7916/D8TD9W2P [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146. Park YG, Park HY, Lee CJ, et al. Ca(V)3.1 is a tremor rhythm pacemaker in the inferior olive. Proc Natl Acad Sci U S A. 2010;107(23):10731‐10736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147. Sinton CM, Krosser BI, Walton KD, Llinás RR. The effectiveness of different isomers of octanol as blockers of harmaline‐induced tremor. Pflugers Arch. 1989;414(1):31‐36. [DOI] [PubMed] [Google Scholar]
  • 148. Choi S, Yu E, Kim D, et al. Subthreshold membrane potential oscillations in inferior olive neurons are dynamically regulated by P/Q‐ and T‐type calcium channels: a study in mutant mice. J Physiol. 2010;588(16):3031‐3043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149. Llinás R, Yarom Y. Oscillatory properties of Guinea‐pig inferior olivary neurones and their pharmacological modulation: an in vitro study. J Physiol. 1986;376:163‐182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150. Crunelli V, Lightowler S, Pollard CE. A T‐type Ca2+ current underlies low‐threshold Ca2+ potentials in cells of the cat and rat lateral geniculate nucleus. J Physiol. 1989;413:543‐561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151. Lang EJ. Organization of olivocerebellar activity in the absence of excitatory glutamatergic input. J Neurosci. 2001;21(5):1663‐1675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152. Brown AM, White JJ, van der Heijden ME, Zhou J, Lin T, Sillitoe RV. Purkinje cell misfiring generates high‐amplitude action tremors that are corrected by cerebellar deep brain stimulation. Elife. 2020;9:e51928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153. Tian JB, Bishop GA. Stimulus‐dependent activation of c‐Fos in neurons and glia in the rat cerebellum. J Chem Neuroanat. 2002;23(3):157‐170. [DOI] [PubMed] [Google Scholar]
  • 154. Milner TE, Cadoret G, Lessard L, Smith AM. EMG analysis of harmaline‐induced tremor in normal and three strains of mutant mice with Purkinje cell degeneration and the role of the inferior olive. J Neurophysiol. 1995;73(6):2568‐2577. [DOI] [PubMed] [Google Scholar]
  • 155. Lorden JF, Stratton SE, Mays LE, Oltmans GA. Purkinje cell activity in rats following chronic treatment with harmaline. Neuroscience. 1988;27:465‐472. [DOI] [PubMed] [Google Scholar]
  • 156. Bernard JF, Buisseret‐Delmas C, Compoint C, Laplante S. Harmaline induced tremor III a combined simple units, horseradish peroxidase, and 2‐deoxyglucose study of the olivocerebellar system in the rat. Exp Brain Res. 1984;57:128‐137. [DOI] [PubMed] [Google Scholar]
  • 157. O'Hearn E, Molliver ME. Degeneration of Purkinje cells in parasagittal zones of the cerebellar vermis after treatment with ibogaine or harmaline. Neuroscience. 1993;55:303‐310. doi: 10.1016/0306-4522(93)90500-F [DOI] [PubMed] [Google Scholar]
  • 158. Zhou M, Melin MD, Xu W, Südhof TC. Dysfunction of parvalbumin neurons in the cerebellar nuclei produces an action tremor. J Clin Invest. 2020;130(10):5142‐5156. doi: 10.1172/JCI135802 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159. Hua SE, Lenz FA, Zirh TA, Reich SG, Dougherty PM. Thalamic neuronal activity correlated with essential tremor. J Neurol Neurosurg Psychiatry. 1998;64(2):273‐276. doi: 10.1136/jnnp.64.2.273 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160. Hua SE, Lenz FA. Posture‐related oscillations in human cerebellar thalamus in essential tremor are enabled by voluntary motor circuits. J Neurophysiol. 2005;93(1):117‐127. doi: 10.1152/jn.00527.2004 [DOI] [PubMed] [Google Scholar]
  • 161. Lin H, Magrane J, Clark EM, et al. Early VGLUT1‐specific parallel fiber synaptic deficits and dysregulated cerebellar circuit in the KIKO mouse model of Friedreich ataxia. Dis Model Mech. 2017;10(12):1529‐1538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162. Wong SB, Wang YM, Lin CC, et al. Cerebellar oscillations in familial and sporadic essential tremor. Cerebellum. 2022;21(3):425‐431. doi: 10.1007/s12311-021-01309-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163. Voller B, Lines E, McCrossin G, et al. Dose‐escalation study of octanoic acid in patients with essential tremor. J Clin Invest. 2016;126(4):1451‐1457. doi: 10.1172/JCI83621 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164. Haubenberger D, McCrossin G, Lungu C, et al. Octanoic acid in alcohol‐responsive essential tremor: a randomized controlled study. Neurology. 2013;80(10):933‐940. doi: 10.1212/WNL.0b013e3182840c4f [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165. Joksovic PM, Choe WJ, Nelson MT, Orestes P, Brimelow BC, Todorovic SM. Mechanisms of inhibition of T‐type calcium current in the reticular thalamic neurons by 1‐octanol: implication of the protein kinase C pathway. Mol Pharmacol. 2010;77(1):87‐94. doi: 10.1124/mol.109.059931 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166. Kito M, Maehara M, Watanabe K. Mechanisms of T‐type calcium channel blockade by zonisamide. Seizure. 1996;5(2):115‐119. [DOI] [PubMed] [Google Scholar]
  • 167. Leppik IE. Zonisamide: chemistry, mechanism of action, and pharmacokinetics. Seizure. 2004;13:S5‐S9. [DOI] [PubMed] [Google Scholar]
  • 168. Mimaki T, Suzuki Y, Tagawa T, Karasawa T, Yabuuchi H. Interaction of zonisamide with benzodiazepine and GABA receptors in rat brain. Med J Osaka Univ. 1990;39(1–4):13‐17. [PubMed] [Google Scholar]
  • 169. Ondo WG. Zonisamide for essential tremor. Clin Neuropharmacol. 2007;30(6):345‐349. [DOI] [PubMed] [Google Scholar]
  • 170. Zesiewicz TA, Ward CL, Hauser RA, Sanchez‐Ramos J, Staffetti JF, Sullivan KL. A double‐blind placebo‐controlled trial of zonisamide (zonegran) in the treatment of essential tremor. Mov Disord. 2007;22(2):279‐282. [DOI] [PubMed] [Google Scholar]
  • 171. Morita S, Miwa H, Kondo T. Effect of zonisamide on essential tremor: a pilot crossover study in comparison with arotinolol. Parkinsonism Relat Disord. 2005;11(2):101‐103. [DOI] [PubMed] [Google Scholar]
  • 172. Handforth A, Martin FC, Kang GA, Vanek Z. Zonisamide for essential tremor: an evaluator‐blinded study. Mov Disord. 2009;24(3):437‐440. [DOI] [PubMed] [Google Scholar]
  • 173. Gören MZ, Onat F. Ethosuximide: from bench to bedside. CNS Drug Rev. 2007;13:224‐239. doi: 10.1111/j.1527-3458.2007.00009.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174. Gironell A, Marin‐Lahoz J. Ethosuximide for essential tremor: an open‐label trial. Tremor Other Hyperkinet Mov (NY). 2016;6:378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175. Papapetropoulos S, Lee MS, Boyer S, Newbold EJ. A phase 2, randomized, double‐blind, placebo‐controlled trial of CX‐8998, a selective modulator of the T‐type calcium channel in inadequately treated moderate to severe essential tremor: T‐CALM study design and methodology for efficacy endpoint and digital biomarker selection. Front Neurol. 2019;10:597. doi: 10.3389/fneur.2019.00597 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176. Papapetropoulos S, Lee MS, Versavel S, et al. A phase 2 proof‐of‐concept, randomized, placebo‐controlled trial of CX‐8998 in essential tremor. Mov Disord. 2021;36(8):1944‐1949. doi: 10.1002/mds.28584 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177. Olhaye O, Belfort G, Raines S, et al. A phase 2 clinical trial evaluating the efficacy, safety, tolerability, and pharmacokinetics of PRAX‐944 in adults with essential tremor [abstract]. Mov Disord. 2022;37(suppl 1). Accessed September 21, 2022. https://www.mdsabstracts.org/abstract/a‐phase‐2‐clinical‐trial‐evaluating‐the‐efficacy‐safety‐tolerability‐and‐pharmacokinetics‐of‐prax‐944‐in‐adults‐with‐essential‐tremor/ [Google Scholar]
  • 178. Olhaye O, Belfort G, Raines S, et al. A phase 2 randomized, double‐blind, placebo‐controlled trial of PRAX‐944 for the treatment of essential tremor [abstract]. Mov Disord. 2022;37(suppl 1). Accessed September 21, 2022. https://www.mdsabstracts.org/abstract/a‐phase‐2‐randomized‐double‐blind‐placebo‐controlled‐trial‐of‐prax‐944‐for‐the‐treatment‐of‐essential‐tremor/ [Google Scholar]
  • 179. Zhao Y, Huang G, Wu Q, et al. Cryo‐EM structures of apo and antagonist‐bound human Cav3.1. Nature. 2019;576:492‐497. doi: 10.1038/s41586-019-1801-3 [DOI] [PubMed] [Google Scholar]
  • 180. Váradi C. Clinical features of Parkinson's disease: the evolution of critical symptoms. Biology (Basel). 2020;9(5):103. doi: 10.3390/biology9050103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181. Bock MA, Brown EG, Zhang L, Tanner C. Association of motor and nonmotor symptoms with health‐related quality of life in a large online cohort of people with Parkinson disease. Neurology. 2022;98(22):e2194‐e2203. doi: 10.1212/WNL.0000000000200113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182. Braak H, Del Tredici K, Rüb U, et al. Staging of brain pathology related to sporadic Parkinson's disease. Neurobiol Aging. 2003;24(2):197‐211. [DOI] [PubMed] [Google Scholar]
  • 183. Gröger A, Kolb R, Schäfer R, Klose U. Dopamine reduction in the substantia nigra of Parkinson's disease patients confirmed by in vivo magnetic resonance spectroscopic imaging. PLoS One. 2014;9(1):e84081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184. Gerlach M, Gsell W, Kornhuber J, et al. A post mortem study on neurochemical markers of dopaminergic, GABA‐ergic and glutamatergic neurons in basal ganglia‐thalamocortical circuits in Parkinson syndrome. Brain Res. 1996;741(1–2):142‐152. [DOI] [PubMed] [Google Scholar]
  • 185. Hashemiyoon R, Kuhn J, Visser‐Vandewalle V. Putting the pieces together in Gilles de la Tourette syndrome: exploring the link between clinical observations and the biological basis of dysfunction. Brain Topogr. 2017;30(1):3‐29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186. Wang L, Zhou C, Cheng W, et al. Dopamine depletion and subcortical dysfunction disrupt cortical synchronization and metastability affecting cognitive function in Parkinson's disease. Hum Brain Mapp. 2022;43(5):1598‐1610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187. Albin RL, Young AB, Penney JB. The functional anatomy of basal ganglia disorders. Trends Neurosci. 1989;12(10):366‐375. doi: 10.1016/0166-2236(89)90074-x [DOI] [PubMed] [Google Scholar]
  • 188. Alexander GE, Crutcher MD. Functional architecture of basal ganglia circuits: neural substrates of parallel processing. Trends Neurosci. 1990;13(7):266‐271. doi: 10.1016/0166-2236(90)90107-l [DOI] [PubMed] [Google Scholar]
  • 189. DeLong MR. Primate models of movement disorders of basal ganglia origin. Trends Neurosci. 1990;13(7):281‐285. doi: 10.1016/0166-2236(90)90110-v [DOI] [PubMed] [Google Scholar]
  • 190. Gao LL, Zhang JR, Chan P, Wu T. Levodopa effect on basal ganglia motor circuit in Parkinson's disease. CNS Neurosci Ther. 2017;23(1):76‐86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191. Thanvi BR, Lo TC. Long term motor complications of levodopa: clinical features, mechanisms, and management strategies. Postgrad Med J. 2004;80(946):452‐458. doi: 10.1136/pgmj.2003.013912 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192. Banerjee AK, Falkai PG, Savidge M. Visual hallucinations in the elderly associated with the use of levodopa. Postgrad Med J. 1989;65(764):358‐361. doi: 10.1136/pgmj.65.764.358 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193. Molina JA, Sáinz‐Artiga MJ, Fraile A, et al. Pathologic gambling in Parkinson's disease: a behavioral manifestation of pharmacologic treatment? Mov Disord. 2000;15(5):869‐872. doi: [DOI] [PubMed] [Google Scholar]
  • 194. Weiss HD, Marsh L. Impulse control disorders and compulsive behaviors associated with dopaminergic therapies in Parkinson disease. Neurol Clin Pract. 2012;2(4):267‐274. doi: 10.1212/CPJ.0b013e318278be9b [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195. Wong JK, Viswanathan VT, Nozile‐Firth KS, et al. STN versus GPi deep brain stimulation for action and rest tremor in Parkinson's disease. Front Hum Neurosci. 2020;14:578615. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196. Parihar R, Alterman R, Papavassiliou E, Tarsy D, Shih LC. Comparison of VIM and STN DBS for parkinsonian resting and postural/action tremor. Tremor Other Hyperkinet Mov (NY). 2015;5:321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197. Katlowitz K, Ko M, Mogilner AY, Pourfar M. Effect of deep brain simulation on arm, leg, and chin tremor in Parkinson disease. J Neurosurg. 2018;131:1‐6. doi: 10.3171/2018.7.JNS18784 [DOI] [PubMed] [Google Scholar]
  • 198. Hacker ML, DeLong MR, Turchan M, et al. Effects of deep brain stimulation on rest tremor progression in early stage Parkinson disease. Neurology. 2018;91(5):e463‐e471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199. Lee DJ, Dallapiazza RF, De Vloo P, Lozano AM. Current surgical treatments for Parkinson's disease and potential therapeutic targets. Neural Regen Res. 2018;13(8):1342‐1345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200. Kumar R, Lozano AM, Kim YJ, et al. Double‐blind evaluation of subthalamic nucleus deep brain stimulation in advanced Parkinson's disease. Neurology. 1998;51(3):850‐855. [DOI] [PubMed] [Google Scholar]
  • 201. Dallapiazza RF, De Vloo P, Fomenko A, et al. Considerations for patient and target selection in deep brain stimulation surgery for Parkinson's disease. In: Stoker TB, Greenland JC, eds. Parkinson's Disease: Pathogenesis and Clinical Aspects. Codon Publications; 2018:8. [PubMed] [Google Scholar]
  • 202. Hutchison WD, Allan RJ, Opitz H, et al. Neurophysiological identification of the subthalamic nucleus in surgery for Parkinson's disease. Ann Neurol. 1998;44(4):622‐628. [DOI] [PubMed] [Google Scholar]
  • 203. Benazzouz A, Breit S, Koudsie A, Pollak P, Krack P, Benabid AL. Intraoperative microrecordings of the subthalamic nucleus in Parkinson's disease. Mov Disord. 2002;17(Suppl 3):S145‐S149. [DOI] [PubMed] [Google Scholar]
  • 204. Starr PA, Rau GM, Davis V, et al. Spontaneous pallidal neuronal activity in human dystonia: comparison with Parkinson's disease and normal macaque. J Neurophysiol. 2005;93:3165‐3176. [DOI] [PubMed] [Google Scholar]
  • 205. Wichmann T, Bergman H, DeLong MR. The primate subthalamic nucleus. III. Changes in motor behavior and neuronal activity in the internal pallidum induced by subthalamic inactivation in the MPTP model of parkinsonism. J Neurophysiol. 1994;72:521‐530. [DOI] [PubMed] [Google Scholar]
  • 206. Benazzouz A, Gross C, Féger J, Boraud T, Bioulac B. Reversal of rigidity and improvement in motor performance by subthalamic high‐frequency stimulation in MPTP‐treated monkeys. Eur J Neurosci. 1993;5(4):382‐389. [DOI] [PubMed] [Google Scholar]
  • 207. Remple MS, Bradenham CH, Kao CC, Charles PD, Neimat JS, Konrad PE. Subthalamic nucleus neuronal firing rate increases with Parkinson's disease progression. Mov Disord. 2011;26(9):1657‐1662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208. Steigerwald F, Pötter M, Herzog J, et al. Neuronal activity of the human subthalamic nucleus in the parkinsonian and nonparkinsonian state. J Neurophysiol. 2008;100(5):2515‐2524. doi: 10.1152/jn.00858.2015 [DOI] [PubMed] [Google Scholar]
  • 209. Galvan A, Devergnas A, Pittard D, et al. Lack of antiparkinsonian effects of systemic injections of the specific T‐type calcium channel blocker ML218 in MPTP‐treated monkeys. ACS Chem Nerosci. 2016;7(11):1543‐1551. doi: 10.1021/acschemneuro.6b00186 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210. Bichler EK, Cavarretta F, Jaeger D. Changes in excitability properties of ventromedial motor thalamic neurons in 6‐OHDA lesioned mice. eNeuro. 2021;8(1):ENEURO.0436‐20.2021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211. Walters JR, Hu D, Itoga CA, Parr‐Brownlie LC, Bergstrom DA. Phase relationships support a role for coordinated activity in the indirect pathway in organizing slow oscillations in basal ganglia output after loss of dopamine. Neuroscience. 2007;144(2):762‐776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212. Magill PJ, Bolam JP, Bevan MD. Dopamine regulates the impact of the cerebral cortex on the subthalamic nucleus‐globus pallidus network. Neuroscience. 2001;106(2):313‐330. [DOI] [PubMed] [Google Scholar]
  • 213. Potashkin JA, Blume SR, Runkle NK. Limitations of animal models of Parkinson's disease. Parkinsons Dis. 2010;2011:658083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214. Rodriguez‐Oroz MC, Rodriguez M, Guridi J, et al. The subthalamic nucleus in Parkinson's disease: somatotopic organization and physiological characteristics. Brain. 2001;124(9):1777‐1790. doi: 10.1093/brain/124.9.1777 [DOI] [PubMed] [Google Scholar]
  • 215. Lenz FA, Tasker RR, Kwan HC, et al. Single unit analysis of the human ventral thalamic nuclear group: correlation of thalamic “tremor cells” with the 3‐6 Hz component of parkinsonian tremor. J Neurosci. 1988;8(3):754‐764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216. Zirh TA, Lenz FA, Reich SG, Dougherty PM. Patterns of bursting occurring in thalamic cells during parkinsonian tremor. Neuroscience. 1998;83(1):107‐121. [DOI] [PubMed] [Google Scholar]
  • 217. Muthuraman M, Raethjen J, Koirala N, et al. Cerebello‐cortical network fingerprints differ between essential, Parkinson's and mimicked tremors. Brain. 2018;141(6):1770‐1781. [DOI] [PubMed] [Google Scholar]
  • 218. Sen S, Kawaguchi A, Truong Y, Lewis MM, Huang X. Dynamic changes in cerebello‐thalamo‐cortical motor circuitry during progression of Parkinson's disease. Neuroscience. 2010;166(2):712‐719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219. Helmich RC, Hallett M, Deuschl G, Toni I, Bloem BR. Cerebral causes and consequences of parkinsonian resting tremor: a tale of two circuits? Brain. 2012;135(Pt 11):3206‐3226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220. Helmich RC, Janssen MJ, Oyen WJ, et al. Pallidal dysfunction drives a cerebellothalamic circuit into Parkinson tremor. Ann Neurol. 2011;69(2):269‐281. [DOI] [PubMed] [Google Scholar]
  • 221. Salamone JD, Mayorga AJ, Trevitt JT, Cousins MS, Conlan A, Nawab A. Tremulous jaw movements in rats: a model of parkinsonian tremor. Prog Neurobiol. 1998;56(6):591‐611. doi: 10.1016/s0301-0082(98)00053-7 [DOI] [PubMed] [Google Scholar]
  • 222. Miwa H, Koh J, Kajimoto Y, Kondo T. Effects of T‐type calcium channel blockers on a parkinsonian tremor model in rats. Pharmacol Biochem Behav. 2011;97(4):656‐659. [DOI] [PubMed] [Google Scholar]
  • 223. Coulter DA, Huguenard JR, Prince DA. Characterization of ethosuximide reduction of low‐threshold calcium current in thalamic neurons. Ann Neurol. 1989;25(6):582‐593. [DOI] [PubMed] [Google Scholar]
  • 224. Gomez‐Mancilla B, Latulippe JF, Boucher R, Bédard PJ. Effect of ethosuximide on rest tremor in the MPTP monkey model. Mov Disord. 1992;7(2):137‐141. [DOI] [PubMed] [Google Scholar]
  • 225. Ge W, Ren J. Combined L−/T‐type calcium channel blockers: ready for prime time. Hypertension. 2009;53(4):592‐594. doi: 10.1161/HYPERTENSIONAHA.108.127548 [DOI] [PubMed] [Google Scholar]
  • 226. Hagiwara N, Irisawa H, Kameyama M. Contribution of two types of calcium currents to the pacemaker potentials of rabbit sino‐atrial node cells. J Physiol. 1988;395:233‐253. doi: 10.1113/jphysiol.1988.sp016916 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227. Hirano Y, Fozzard HA, January CT. Characteristics of L‐ and T‐type Ca2+ currents in canine cardiac Purkinje cells. Am J Physiol. 1989;256(5 Pt 2):H1478‐H1492. doi: 10.1152/ajpheart.1989.256.5.H1478 [DOI] [PubMed] [Google Scholar]
  • 228. Tseng GN, Boyden PA. Multiple types of Ca2+ currents in single canine Purkinje cells. Circ Res. 1989;65(6):1735‐1750. doi: 10.1161/01.res.65.6.1735 [DOI] [PubMed] [Google Scholar]
  • 229. Bohn G, Moosmang S, Conrad H, Ludwig A, Hofmann F, Klugbauer N. Expression of T‐ and L‐type calcium channel mRNA in murine sinoatrial node. FEBS Lett. 2000;481(1):73‐76. doi: 10.1016/s0014-5793(00)01979-7 [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Figure S1


Articles from Annals of Clinical and Translational Neurology are provided here courtesy of Wiley

RESOURCES