This is an open access publication, available online and distributed under the terms of a Creative Commons Attribution-Non Commercial-No Derivatives 4.0 International licence (CC BY-NC-ND 4.0), a copy of which is available at https://creativecommons.org/licenses/by-nc-nd/4.0/. Subject to this license, all rights are reserved.
NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.
Noebels JL, Avoli M, Rogawski MA, et al., editors. Jasper's Basic Mechanisms of the Epilepsies. 5th edition. New York: Oxford University Press; 2024. doi: 10.1093/med/9780197549469.003.0022
Abstract
The thalamus, a deep brain structure with broad connectivity, plays key roles in local and global rhythmic activity in sleep, arousal, and cognition. Disruptions in thalamic and thalamocortical circuits—and the ensuing hypersynchrony and hyperexcitability—have been widely studied in the context of genetic epilepsies and have garnered increasing interest in the context of acquired epilepsies. In this chapter, key structural, synaptic, cellular, and biophysical elements underlying the rhythmogenic properties of the thalamus are described, and then how diverse perturbations of these properties in rodent genetic models ultimately generate or facilitate seizures. The chapter briefly highlights recent studies which have identified mechanisms of thalamic involvement in the development or modulation of epilepsies acquired following incidents such as traumatic brain injury and ischemic stroke. Understanding how diverse etiologies converge upon thalamic hyperexcitability can pinpoint elements of vulnerability, resilience, redundancy, necessity, and sufficiency in the thalamocortical circuit. Understanding how the thalamus generates and modulates aberrant activity—even when it is not the primary or sole site of genetic perturbation—will be key to identifying therapeutic targets and paradigms to treat epilepsies. Such efforts will benefit from continued advancements in our knowledge of cell-type heterogeneity, meso- and macro-scale connectivity, and interspecies differences in the thalamus.
Introduction
The thalamus, a subcortical diencephalic structure involved in sleep, arousal, consciousness, and cognition, has been a site of investigation in seizures since the 1940s, starting with experimental studies in cats and monkeys by Dempsey, Morison, Jasper, and Droogleever-Fortuyn. These studies showed that intralaminar stimulation evoked bilateral 3-Hz spike-and-wave discharges and behavioral arrest (reviewed by Jasper, 1948, 1991). In 1950, the first depth recordings in the thalamus during petit mal seizures (now known as “absence seizures”) were obtained from patients with absence type epilepsy (Spiegel and Wycis, 1950), and subsequent studies of thalamic activity during seizures supported the hypothesis that “the clinical case of petit mal epilepsy is due to a disturbance in the thalamus which causes a rhythmic discharge throughout the cortex” (Williams, 1953). While there has been a debate on whether the seizures initiate in cortex or thalamus, thalamic firing has indeed been found necessary to maintain absence seizures in rodents (Sorokin et al., 2017). The study of epileptic discharges, in both patients and animal models, has yielded basic insights about thalamic organization and function. Reciprocally, advances in our understanding of thalamic structure and function have elucidated mechanisms of the generation and propagation of epileptic activity such as spike-wave discharges underlying absence seizures, and it has opened avenues for the therapeutic modulation of seizures.
Endowed with extensive, reciprocal connections with the entire cerebral cortex and other subcortical structures (Jones, 1985), and possessing rhythmogenic properties (reviewed by Fogerson and Huguenard, 2016), the thalamus plays a critical role in the generation of normal and pathological synchronized oscillatory activity (Steriade, McCormick, and Sejnowski, 1993). Decades of research have revealed disruptions in thalamocortical excitability and synchrony as a convergence point for epileptic activity related to multiple genetic etiologies and pharmacological manipulations. More recently, cell-type-specific optogenetics have enabled causal manipulations of rodent thalamocortical synapses and circuits, demonstrating a role for the thalamus as a “choke-point” in certain forms of seizures and epilepsies (Paz et al., 2013; reviewed by Paz and Huguenard, 2015). Indeed, in 2018 the U.S. Food and Drug Administration (FDA) approved deep brain stimulation of the anterior thalamus as a treatment for refractory focal epilepsy (reviewed by Fisher and Velasco, 2014).
By understanding exactly which aspects of thalamic function are vulnerable—or resilient—to disruptions, we can clarify strategies to effectively interfere with seizure expression while minimizing side effects. In this chapter, we describe the key anatomical, structural, cellular, and biophysical elements of the thalamus which confer its rhythmogenic properties. We then provide a brief overview of the extensive literature on genetic rodent models of seizures, with particular emphasis on the diverse role of calcium channels in absence seizures and compensatory changes in the thalamocortical network. Finally, we briefly discuss recent work highlighting thalamic involvement in the development or modulation of epilepsies acquired after brain injuries.
Thalamic Organization and Rhythmogenesis
The thalamus is composed of various nuclei embedded in diverse, multiregional neural circuits, whose complexity and detailed connectivity are not yet fully understood (reviewed by Halassa and Sherman, 2019; Shepherd and Yamawaki, 2021). Given the widespread connectivity between thalamus and the entire cerebral cortex, the thalamus is well-positioned to control the generation and maintenance of cortical rhythms. Furthermore, the thalamus is notable for its generation and maintenance of synchronous oscillatory activity (reviewed by Steriade, McCormick, and Sejnowski, 1993; Huguenard and McCormick, 2007; Beenhakker and Huguenard, 2009; Crunelli and Hughes, 2010; Fogerson and Huguenard, 2016). Such oscillations are considered critical substrates of states of arousal, sleep, and consciousness (reviewed by Steriade and Llinas, 1988; Steriade, McCormick, and Sejnowski, 1993; McCormick and Bal, 1997; Crunelli et al., 2018).
Three cardinal rhythms of the thalamus include delta, spindles, and slow-wave oscillations, all of which occur during non–rapid eye movement (NREM) sleep (reviewed by Crunelli and Hughes, 2010; Fogerson and Huguenard, 2016; Fernandez and Lüthi, 2020). We will describe them in the sixth section. Key features of the thalamus which enable the generation of these synchronous rhythms include cell-intrinsic firing properties, microcircuit connectivity, long-range connections with the cerebral cortex and basal ganglia (reviewed by Deschênes, Veinante, and Zhang, 1998; Jones, 2007; Paz et al., 2007; Charpier, Beurrier, and Paz, 2010), and the precise timing of synaptic inputs (reviewed by Huguenard and McCormick, 2007). The cellular and circuit properties underlying the generation and regulation of these oscillations can subserve aberrant oscillations underlying epileptic seizures, and they will therefore be reviewed in the following sections. Thalamic connectivity with basal ganglia will not be reviewed in this chapter.
Structural Elements of Thalamic and Thalamocortical Circuits: From Gross Anatomy to Cell Types
The multiple nuclei of the thalamus are each embedded in reciprocal circuits that can be local (i.e., between thalamic relay nuclei and the reticular thalamic nucleus) or long range (i.e., between thalamic relay nuclei and cortex) (Fig. 22–1A).

Figure 22–1.
Basic structural elements of thalamic and thalamocortical circuits. A. Coronal section of mouse brain labeled with the neuronal nuclear protein NeuN. Colored regions highlight the reticular thalamic nucleus (known as nRT, for nucleus reticularis thalami (more...)
The reticular thalamic nucleus (abbreviated nRT for nucleus reticularis thalami) is a thin shell of GABAergic neurons which surrounds the lateral boundary of dorsal thalamocortical (TC) relay nuclei (Houser et al., 1980; Pinault, 2004). nRT neurons form reciprocal connections with glutamatergic neurons in thalamic relay nuclei (Jones, 1975; Sherman and Guillery, 1996; Guillery and Sherman, 2002) and receive corticothalamic (CT) collateral afferents; thus, the nRT provides both feedback and feedforward inhibition in the thalamus and critically shapes net thalamic output (reviewed by Pinault, 2004; Paz and Huguenard, 2015; Halassa and Acsády, 2016; Crabtree, 2018). Approximately 10%–25% of synapses in the nRT are GABAergic (Williamson, Ohara, and Ralston, 1993; Liu and Jones, 1999). There is histological and ultrastructural evidence that the nRT receives GABAergic input from the external segment of the globus pallidus in squirrel monkeys (Asanuma, 1994), the substantia nigra pars reticulata in cats (Paré et al., 1990), the basal forebrain and mesopontine tegmentum in rats (Jourdain, Semba, and Fibiger, 1989), and notably, from the nRT itself in cats, rats, and mice (Houser et al., 1980; Pinault, Bourassa, and Deschênes, 1995; Liu and Jones, 1999; Clemente-Perez et al., 2017). However, the existence of functional, GABAergic synapses between nRT neurons is debated (discussed below).
The majority of thalamic territory is occupied by various thalamic relay nuclei, which earned the term “relay” due to their well-studied role in transmitting afferent sensory information to other brain regions. Other than the nRT, sources of GABAergic inhibition in thalamic relay nuclei include the zona incerta, anterior pretectal nucleus, basal ganglia (including substantia nigra pars compacta, internal globus pallidus, ventral pallidum), and the pontine reticular formation, which are well described elsewhere (reviewed by Halassa and Acsády, 2016). Based on their primary afferent source, thalamic relay nuclei are broadly organized into first-order and higher-order nuclei. First-order nuclei receive sensor-specific driving input from subcortical sources, while higher-order nuclei receive driving input from the cortex (reviewed by Halassa and Sherman, 2019). Based on their input-output connectivity, thalamic relay nuclei can also be broadly organized into sensory, limbic, motor, and executive/cognitive domains, which are preserved along the thalamo-cortico-reticular pathway (Sherman and Guillery, 1996; Jones, 2007; also see recent studies, e.g. Harris et al., 2019; Phillips et al., 2019). The vast majority of neurons in thalamic relay nuclei are glutamatergic and are called “thalamocortical” (TC) neurons because they form reciprocal, long-range, excitatory, and topographic connections with cortical neurons. For example, key aspects of visual information processing involve TC neurons in the dorsal lateral geniculate nucleus (dLGN), which receive retinal input, project to primary visual cortex (V1), and receive corticothalamic input from V1; in addition, dLGN neurons form reciprocal connections with the GABAergic neurons of the nRT, which also receive corticothalamic input from V1 (Sherman and Guillery, 1996). Similarly, key aspects of somatosensory information processing involve TC neurons from the ventrobasal nucleus (VB), which receive input from the spinothalamic tract and medial lemniscus (Mo et al., 2017), project to primary somatosensory cortex (S1), and receive corticothalamic input from S1; in addition, VB neurons form reciprocal connections with the GABAergic neurons of the nRT, which also receive corticothalamic input from S1 (Deschênes, Veinante, and Zhang, 1998; Jones, 2009). The somatosensory VB-S1-nRT circuit is one of the most well-characterized circuits in the context of basic intrinsic and synaptic properties, and in rodent models of seizures and epilepsy.
To understand the cellular basis of thalamocortical rhythmic activity, it is important to understand the relevant cell types and their connectivity (i.e., microcircuit motifs). Thalamocortical circuits consist of multiple key cell types, including glutamatergic TC relay neurons in thalamic relay nuclei, glutamatergic corticothalamic (CT) neurons, primarily in cortical layers 5 and 6, and GABAergic neurons in the nRT (Fig. 22–1B) (Deschênes, Veinante, and Zhang, 1998; Shepherd and Yamawaki, 2021). Although there is increasing appreciation of cell-type heterogeneity among these populations (Clemente-Perez et al., 2017; Phillips et al., 2019; Li et al., 2020; Martinez-Garcia et al., 2020), the basic organization of these microcircuit motifs in the thalamus is generally accepted (reviewed by Paz and Huguenard, 2015).
Thalamocortical:
- •
Recurrent excitation (formed by reciprocal glutamatergic projections; prominent in cortical networks)
- •
Feedforward inhibition (disynaptic component, whereby excitatory inputs recruit local inhibition en route to its recipient; can effectively modulate the strength of the efferent signal)
Intra-thalamic:
- •
Feedback inhibition (reciprocally connected glutamatergic and GABAergic neurons; an inhibitory mechanism to control local excitatory activity)
- •
Counter-inhibition (reciprocally connected GABAergic neurons)
In addition to cell-type-specific connectivity patterns, these distinct microcircuits include layer-specific connectivity patterns (Halassa and Sherman, 2019; Shepherd and Yamawaki, 2021). For example, in first-order thalamic relay nuclei such as the dLGN and VB, TC neurons send long-range glutamatergic projections to related cortical regions in a topographic manner, primarily in layer 4 (Deschênes, Veinante, and Zhang, 1998; Jones, 2007). Layer 6 CT neurons send long-range glutamatergic projections to TC neurons, and also form glutamatergic collaterals onto nRT, which enable feedforward inhibition of TC neurons (Jones, 2007; Sherman, 2007).
Functional connections between nRT neurons, which form a reciprocal counter-inhibition microcircuit, introduce a distinct set of principles into the framework of thalamic function in health and disease (Huntsman et al., 1999; Sohal and Huguenard, 2003; Paz and Huguenard, 2015; Makinson et al., 2017; Crabtree, 2018). Gap junction-coupled intra-nRT electrical synapses have been consistently demonstrated by multiple groups (Landisman et al., 2002; Long, Landisman, and Connors, 2004; Deleuze and Huguenard, 2006). However, the existence of intra-nRT chemical GABAergic synapses has been a subject of controversy (reviewed by Paz and Huguenard, 2015; Crabtree, 2018), with multiple electrophysiological studies, suggesting their existence and functionality in ferrets, rats, and mice (Shu and McCormick, 2002; Deleuze and Huguenard, 2006; Makinson et al., 2017), and others failing to find them (Parker, Cruikshank, and Connors, 2009; Cruikshank et al., 2010; Hou, Smith, and Zhang, 2016). Nevertheless, in this chapter we refer to several studies which highlight the relevance of intra-nRT connections for our understanding of thalamic circuit synchrony.
Disruption of any microcircuit motif may lead to thalamic hyperexcitability and seizures. For example, in the Gria4–/– mouse model of absence epilepsy, there is a specific and selective deficit of the cortico-nRT projection (Paz et al., 2011). The consequent loss of feedforward inhibition (despite normal feedback inhibition) underlies the generation of pathological oscillation. In mice lacking the GABAA receptor β3 subunit, a selective loss of GABAAR-mediated inhibition in reciprocally connected nRT neurons leads to dramatic thalamic hypersynchrony in slices. The latter finding highlights the desynchronizing role of counter-inhibition within the nRT (Huntsman et al., 1999; Sohal, Huntsman, and Huguenard, 2000; Sohal and Huguenard, 2003). This role is further supported by the observation that selective loss of the voltage-gated sodium channel Scn8a in nRT neurons, which leads to loss of intra-nRT, but not nRT-TC, inhibition, causes mice to develop spontaneous absence seizures (Makinson et al., 2017).
Thalamic Firing
TC and nRT neurons are generally thought to exhibit two distinct firing modes: tonic firing in response to depolarization (Jahnsen and Llinás, 1984), and burst firing (in response to hyperpolarization, or depolarization in nRT (Steriade and Llinas, 1988; Huguenard, 1996; Sherman, 2001; however, see Wolfart et al., 2005). During states of arousal (e.g., wake), thalamic neurons are relatively depolarized and exhibit “tonic” firing of action potentials at low frequencies; such tonic firing is proposed to be involved in sensory information processing (Swadlow, Gusev, and Bezdudnaya, 2002; Sherman, 2007). During states of low arousal (e.g., sleep), thalamic neurons are relatively hyperpolarized and exhibit high-frequency bursts of action potentials. Burst firing can occur periodically through both cell-intrinsic and circuit mechanisms (as described in the sixth section), and it is a critical building block of thalamic oscillations (Llinás and Steriade, 2006; Beenhakker and Huguenard, 2009; Fogerson and Huguenard, 2016). T-type voltage-gated calcium channels (subtypes CaV3.1, CaV3.2, and CaV3.3) are crucial elements of burst firing in thalamic neurons and are described in detail in the fifth section.
Oscillatory burst firing is a critical substrate of sleep. It is abolished in the nRT neurons of CaV3.3–/– mice, whereas tonic firing is not, according to ex vivo whole-cell, patch-clamp recordings (Astori et al., 2011). Lack of CaV3.3-mediated oscillatory bursting results in a selective reduction of the sleep-associated sigma frequency band power in the EEG at transitions from NREM to REM (where sleep spindles occur). Thus, CaV3.3-mediated burst firing has been proposed to be the predominant pacemaker underlying sleep spindles in the thalamus (Astori et al., 2011).
Burst firing is also a critical substrate of thalamic seizures, because of its involvement in spike-and-wave discharges (SWDs). Generalized SWDs are hallmarks of absence seizures (reviewed by Crunelli and Leresche, 2002; Blumenfeld, 2005; Tenney and Glauser, 2013; Fogerson and Huguenard, 2016). The electroencephalographic (EEG) features of SWDs are bilateral, high-amplitude events composed of stereotyped spike-and-wave signatures which typically occur at 3–4.5 Hz in humans and 6–10 Hz in rodents (reviewed by Crunelli and Leresche, 2002; Tenney and Glauser, 2013; Fogerson and Huguenard, 2016). Multi-unit recordings have shown that TC, CT, and cortical neurons exhibit synchronous burst firing, time-locked to the “spike” component of the SWD in vivo in rats and mice (Inoue et al., 1993; Sorokin et al., 2017); at the intracellular level, TC and nRT neurons exhibit burst firing correlated with the “spike” component of the SWD in vivo in rats (Slaght et al., 2002; Paz et al., 2007).
Recent studies have provided further compelling evidence for the role of thalamic burst firing in seizures (reviewed by Huguenard, 2019). Sorokin and colleagues expressed a stable step-function opsin (Yizhar et al., 2011) in VB TC neurons of Wistar Albino Glaxo from Rijswijk (WAG/Rij) rats and stargazer mice—rodent models of absence epilepsy (Sorokin et al., 2017)—which enabled them to switch the firing mode of TC cells between tonic and burst mode. They found that toggling the firing from bursting to tonic mode stopped ongoing absence seizures. Stable step-function opsins have also been used to demonstrate the role of TC bursting in atypical, nonconvulsive seizures observed in a mouse model of Dravet syndrome—seizures that have been found in human patients (Ritter-Makinson et al., 2019). Thus, there is strong evidence for the role of thalamic neuronal population bursting in the maintenance of nonconvulsive seizures (Sorokin et al., 2017). Consistent with these findings, in a different rodent model of absence epilepsy—the Genetic Absence Epilepsy Rat from Strasbourg (GAERS), McCafferty and colleagues have found that the output of TC neurons during absence seizures is strong and rhythmic, even though burst firing in each TC cell is infrequent, suggesting that bursting in each TC cell is not required for maintaining the seizure (McCafferty et al., 2018). In contrast, burst firing in nRT neurons is very robust during absence seizures and likely serves to synchronize the TC output to the cortex (Slaght et al., 2002; McCafferty et al., 2018).
Ultimately, the thalamocortical network dynamics underlying seizures is likely an emergent property of both cell-intrinsic mechanisms (e.g., T-type mediated burst firing) and circuit properties (e.g., cortical feedforward inhibition of the thalamus). Nevertheless, given that disruptions to firing mode can alter thalamic synchrony in ways that can lead to seizures, and that thalamic synchrony depends in part on T-type calcium channels, we next provide an overview of the key enabler of thalamic switching in firing mode: the low-threshold T-type calcium channel.
Calcium Channels and Thalamic Firing
Calcium Channels
Voltage-gated calcium channels (VGCCs) are widely expressed throughout the mammalian body and brain (reviewed by Catterall, 2000), and support numerous critical functions, including cellular excitability, synaptic transmission, and synaptic plasticity (reviewed by Catterall and Few, 2008) (Fig. 22–2). VGCCs are organized into two main classes, distinguished by their biophysical properties: high-voltage activated (HVA) VGCCs, which include L-type, N-type, and P/Q-type channels (Nowycky, Fox, and Tsien, 1985; Rajakulendran, Kaski, and Hanna, 2012), and low-voltage activated (LVA) VGCCs, which include T-type calcium channels (Perez-Reyes, 2003); R-type channels are considered to have intermediate properties (Soong et al., 1993; Randall and Tsien, 1997) and are discussed further in the seventh section. VGCCs are composed of four or five distinct subunits, including a pore-forming α1 subunit (of which there are three families, CaV1, CaV2, and CaV3), a transmembrane complex of α2 and δ subunits, an intracellular β subunit, and in some cases a transmembrane γ subunit (Catterall, 2000). Because the α1 subunit contains the conduction pore, voltage sensors, gating mechanisms, and most of the known sites of channel regulation, its family and subtype typically correlate with whether the calcium channel is of the HVA or LVA type (Catterall and Few, 2008). Here we provide an overview of LVA T-type calcium channels, given their prominent role in thalamic bursting rhythmogenesis (which we will describe in the sixth section). The role of HVA R and P/Q-type calcium channels in seizures will be discussed in the seventh section.

Figure 22–2.
CaV3.1 and CaV3.3 are the dominant T-type calcium channel subunits expressed in TC relay nuclei and in the nRT. A. Sagittal section of mouse brain indicating location of the nRT and TC relay nuclei (labels are not drawn to scale). All images obtained (more...)
Spotlight on T-Type Calcium Channels and Burst Firing
T-type calcium channels are critical determinants of thalamic neuron electrophysiology (Huguenard, 1996). In addition to their activation at low voltages, T-type calcium channels have unique inactivation properties (time-dependent inactivation, steady-state inactivation, and recovery from inactivation), as well as lower conductance (thus resulting in transient and “tiny” currents), compared to HVA VGCCs. Thus, currents generated by T-type calcium channels (referred to as IT, or T-type current) are uniquely positioned to regulate cellular excitability and firing pattern, particularly near the resting membrane potential in neurons with robust burst firing behaviors (discussed further in the sixth section).
In TC neurons at resting membrane potential (between –60 to –65 mV), the majority of T-type calcium channels are in an inactivated state. Hyperpolarization (to at least –70 mV) enables de-inactivation, and subsequent depolarization leads to activation which allows a transient influx of Ca2+ and generates the low-threshold spike (LTS), also known as the rebound plateau potential. The LTS further depolarizes the membrane potential such that voltage-gated sodium channels generate action potentials, culminating in a barrage of rebound high-frequency burst firing. At this point, T-type calcium channels are inactivated (reviewed by Cheong and Shin, 2013; Fogerson and Huguenard, 2016). The sequence of events that underlies oscillatory activity is summarized in Figure 22–3, and discussed later in various pathophysiological contexts.

Figure 22–3.
Cell-intrinsic and synaptic properties enabling thalamic burst firing and oscillations. (A) Schematic of the interplay between nRT and TC neurons, which gives rise to rhythmic activity in the intra-thalamic circuit such as spindle oscillations. Cortical (more...)
Spatial Distribution of T-Type Calcium Channel Subunits
The α1 subunit of T-type calcium channels are encoded by the CaV3 family, which consists of three subtypes which are differentially expressed in the brain. Within the rodent thalamus, channels containing CaV3.1 (encoded by Cacna1g) are the primary source of T-type calcium channels in TC neurons. nRT neurons primarily express the CaV3.3-containing T-type calcium channel (encoded by Cacna1i) and to a lesser extent, CaV3.2-containing T-type channel calcium (encoded by Cacna1h) (Talley et al., 1999; Joksovic et al., 2006) (Fig. 22–2). Differences among the three subtypes of T-type calcium channels include inactivation kinetics and recovery from steady-state inactivation. Because inactivation influences the ability of T-type calcium channels to trigger an LTS, the expression pattern and subcellular localization of subtypes are important considerations for thalamic electrophysiological properties. T-type channels are prominently expressed in the soma and dendrites. In rodent TC neurons, electrophysiological, immunohistochemical, and computational models have found that CaV3.1 is prominently expressed in the soma and dendrites as well. While the distribution along the dendritic axis remains debated (e.g., Destexhe et al., 1998; Parajuli et al., 2010), it is a particularly interesting question, given that the spatial distribution should shape the interaction of synaptic inputs, burst generation, and downstream calcium-mediated signaling pathways (e.g., Cueni et al., 2008; Crandall, Govindaiah, and Cox, 2010).
TC neurons mainly express Cav3.1-containing channels, which recover from inactivation the fastest out of the three subtypes (reviewed by Cheong and Shin, 2013). nRT neurons mainly express Cav3.3-containing channels, which display relatively slow inactivation and are nearly voltage-independent (Huguenard and Prince, 1992). This results in longer calcium-dependent spike bursts generated in nRT, which translates into longer bouts of GABA release onto TC neurons. Thus, the difference in inactivation kinetics critically determines the precise timing and rhythmicity of the intra-thalamic oscillations.
The rich literature in this field is well-described in other reviews (reviewed by Crunelli, Cope, and Hughes, 2006; Cheong and Shin, 2013; Fogerson and Huguenard, 2016). In the future, increased specificity of genetic manipulations and increased sampling capacity of recording techniques may lead to a more definitive understanding of the combinations of calcium channel subtypes in distinct cell types that are necessary and sufficient for the expression of various rhythmic oscillations.
Rhythmogenesis in the Thalamus: Strengths and Weaknesses
Thalamic circuits play key roles in local and global oscillations (reviewed by Huguenard and McCormick, 2007; Fogerson and Huguenard, 2016; Crunelli et al., 2018). Three cardinal rhythms of the thalamus during NREM sleep include delta (1–4 Hz), spindles (7–15 Hz), and slow-wave oscillations (<1 Hz). Recent work has also unveiled the thalamic involvement in alpha waves (Hughes et al., 2011) and gamma oscillations (Hoseini et al., 2021), which will not be addressed in this chapter. We have already highlighted how burst firing is mediated by T-type calcium channels (fifth section). Below we discuss how these channels, along with cellular and circuit properties of the thalamus, enable the generation of delta, spindles, and slow-wave rhythms, and what disruptions in these circuits can hijack these rhythms and contribute to seizures.
Delta Oscillations
TC relay neurons display intrinsic pacemaker activity in the 1–4 Hz frequency range; this frequency band, referred to as delta, is associated with deep sleep (i.e., stage 3 NREM) and anesthesia (Brown, Lydic, and Schiff, 2010).
Cell-intrinsic mechanisms: This pacemaker activity arises from an intrinsic interplay between two currents generated by the low-threshold T-type calcium channel (IT) and the hyperpolarization-activated cyclic nucleotide-gated (HCN) (Ih) (McCormick and Pape, 1990; Soltesz et al., 1991; Nuñez, Amzica, and Steriade, 1992; Destexhe et al., 1996) (Fig. 22–3). Upon sufficient membrane hyperpolarization, IT is de-inactivated and Ih is activated. Ih is a slowly developing, non-inactivating inward current which has a depolarizing effect. Ih depolarization of the membrane activates IT and triggers a LTS, which further depolarizes the membrane and triggers Na+-dependent action potentials. The depolarization associated with the LTS deactivates Ih; in the absence of Ih activation, the membrane hyperpolarizes, which then reactivates Ih and de-inactivates IT, and the cycle continues.
Synaptic mechanisms: Synchronization of this autonomously generated delta-frequency membrane oscillation is thought to be mediated by cortico-TC and cortico-nRT-TC feedforward inhibition during cortical UP states (Steriade, Dossi and Nunez, 1991; Neske, 2016). Evidence suggests that the local thalamic delta-frequency oscillations indeed contribute to global delta waves observed in the EEG during NREM sleep. The extent to which (and the mechanisms whereby) this local thalamic delta-frequency oscillation contributes to global delta waves observed in the EEG during NREM sleep has been an active area of research. Global CaV3.1–/– mice lack delta waves at the EEG level (Lee, Kim, and Shin, 2004), at least in part due to the loss of burst firing in TC neurons (Kim et al., 2001). Recent studies employing optogenetic and chemogenetic techniques have suggested that the nRT is capable of modulating delta waves in spatially restricted regions of the cortex (Lewis et al., 2015; Fernandez et al., 2018).
Pro-epileptic mechanisms: Disruption of calcium channels appears to be a common node of dysfunction across various genetic models of absence seizures. T-type calcium channels have been the most widely studied calcium channels in absence epilepsy, and the diverse ways in which IT has been linked to absence seizures are discussed further in the seventh section. The disruption of Ih in the thalamus can also promote thalamic hyperexcitability and seizures. For instance, deletion of HCN2, an Ih subunit prominently expressed in both the thalamic relay nuclei and the nRT (Notomi and Shigemoto, 2004), leads to spontaneous absence seizures in mice. Although Ih plays a key role in pacemaker bursting activity, the loss of Ih in TC neurons in these models can still lead to seizures because it can lead to hyperpolarization of TC neurons, which in turn facilitates the de-inactivation of T-type calcium channels and burst firing (Ludwig et al., 2003; Chung et al., 2009). Interestingly, deletion of the tetratricopeptide-repeat-containing Rab8b-interaction protein (TRIP8b), an auxiliary subunit of HCN channels expressed in TC and CT neurons but not in nRT neurons, results in a milder absence seizure phenotype (Heuermann et al., 2016). Thus, preservation of Ih in the nRT of TRIP8b–/– mice appears to partially constrain the heightened excitability of the thalamocortical circuit due to the loss of Ih in TC neurons. This property is reminiscent of the apparent ability of nRT Ih to constrain cortico-nRT synaptic integration in normal contexts (Ying et al., 2007), and it suggests that Ih in the nRT confers partial resilience to hyperexcitability in the circuit.
Thalamic Ih can also be disrupted in the context of epileptogenesis. In a rodent model of post-stroke epilepsy, whereby stroke in the S1 cortex leads to secondary injury in the VB thalamus, TC neurons exhibited alterations in the biophysical properties of Ih, such as depolarized half-activation voltage (Paz et al., 2013). Biophysical modeling demonstrated that these changes in Ih were sufficient to enhance thalamic network oscillations, and optogenetic inhibition of these hyperexcitable TC neurons was sufficient to abort epileptic seizures (Paz et al., 2013).
Finally, Ih in TC neurons has also been reported to be increased, rather than decreased, in two different rat models of absence epilepsy (Kanyshkova et al., 2012; Cain et al., 2015). Perhaps a noteworthy commonality across several of these models featuring Ih disruption (in which Ih is abolished in both TC and nRT, Ih is abolished in TC, or Ih exhibits depolarized half-activation) is that they result in seizures which appear to have slower internal frequencies than typical absence seizures observed in rodents. While typical rodent SWDs are 6–10 Hz (Fogerson and Huguenard, 2016) and characteristically faster than typical human SWDs which are 3–4.5 Hz (Tenney and Glauser, 2013), the mice described above exhibited SWDs in the range of 5 Hz (HCN2–/– mice), 5.3 ± 1.7 Hz (Trip8b–/– mice), and 4–5 Hz (post-stroke epilepsy rats) (Ludwig et al., 2003; Chung et al., 2009; Paz et al., 2013; Heuermann et al., 2016).
Spindle Oscillations
Spindle oscillations (7–15 Hz), a global rhythmic signature of stage 2 NREM sleep, emerge in the intrathalamic circuit as a result of recurrently connected TC and nRT neurons, which exhibit alternating bouts of rebound burst firing (Llinás and Steriade, 2006; Fernandez and Lüthi, 2020).
Cell-intrinsic mechanisms: nRT neurons exhibit intrinsic rhythmicity in the spindle frequency range (7–12 Hz), which can be maintained for up to 1 second following release from hyperpolarization (von Krosigk, Bal, and Mccormick, 1993). This intrinsic oscillation in nRT arises from the interplay of IT and Ca2+-dependent small-conductance type 2 potassium channels (SK2 channels) (Köhler et al., 1996; Cueni et al., 2008; Wimmer et al., 2012), similar to the interplay of IT and Ih which gives rise to the intrinsic delta oscillation in TC neurons described above. Ca2+ influx through the Cav3.3 subtype (the main T-type calcium channel in nRT) activates the SK2 channel-mediated K+ current (ISK), which generates a slow afterhyperpolarization (AHP) following the LTS burst. The slow AHP allows the de-inactivation of IT, and thus the cycle of repetitive burst firing continues (Cueni et al., 2008) (Fig. 22–3). Another key contributor is the Cav3.2 R-type calcium channel, which has been proposed to enable prolonged elevation of Ca2+ required for induction of the slow AHP, which in turn is critical for allowing the continued oscillatory bursting in nRT (Zaman et al., 2011; Paz and Huguenard, 2012).
Synaptic mechanisms: During a spindle oscillation, rhythmic firing of nRT neurons triggers a transient burst of IPSPs in reciprocally connected TC neurons. The barrage of IPSPs, mediated by both GABAA and GABAB receptors, transiently hyperpolarizes TC neurons such that IT is de-inactivated. Upon release of inhibition from nRT, TC neurons fire LTS-mediated rebound bursts, which in turn re-excite nRT neurons and initiate another nRT burst (Huguenard and Prince, 1994a; Fogerson and Huguenard, 2016) (Fig. 22–3). The precise timing of this interplay is influenced by the duration of the IPSPs (governed by the relative expression of GABAA versus GABAB receptors in the recipient TC neurons), as well as the kinetics of the LTS (governed by the T-type calcium channel subtype expressed in each cell type). Recent studies have shown that nRT neurons can play a causal role in initiating spindles (e.g., Halassa et al., 2011; Barthó et al., 2014; Fernandez et al., 2018).
Pro-epileptic mechanisms: The relationship between spindle oscillations and spike-and-wave discharges is an unresolved debate (reviewed by Beenhakker and Huguenard, 2009; Leresche et al., 2012). Recent studies have provided support for the hypothesis that spindles can be hijacked into epileptiform activity. Children with childhood epilepsy with centrotemporal spikes (the most common focal epilepsy syndrome) exhibit transient, sleep-activated focal deficits in spindles (Kramer et al., 2021). Using high-density EEGs during sleep and cognitive testing, the authors found an inverse relationship between spike and spindle rate during NREM sleep. In addition, they found that spindle rate predicted performance on cognitive tasks. These findings provide compelling support for the hypothesis that the same underlying thalamocortical circuit can give rise to spindles and epileptic activity (Beenhakker and Huguenard, 2009), as well as underlie seizures and cognitive dysfunction (Kramer et al., 2021).
Further support for the shared circuit hypothesis comes from a recent study of a rodent model of thalamocortical dysfunction following mild traumatic brain injury (TBI) (Holden et al., 2021). Unilateral injury in the primary somatosensory cortex led to chronic secondary damage of the thalamus, including loss of nRT neurons and a selective deficit in nRT synaptic function, but not in CT and TC neuronal functions. These mice develop epileptic spikes and exhibit a loss in sleep spindles; both changes were focal to the S1 cortex ipsilateral to injury (Holden et al., 2021).
Slow Oscillations
Finally, thalamic circuits can also generate slow oscillations (<1 Hz) (reviewed by Crunelli and Hughes, 2010; Neske, 2016; Crunelli et al., 2018), although these were traditionally considered to be exclusively generated in cortical networks (Steriade, McCormick, and Sejnowski, 1993). The slow rhythm, present in almost all NREM stages, has been proposed to group together periods of sleep spindles and delta waves (Crunelli and Hughes, 2010). Both TC and nRT neurons have been shown to play an active role in shaping the slow rhythm during sleep. In TC neurons, the slow oscillation arises from the interplay of IT and the leak K+ current, and is influenced by Ca2+- activated nonselective cation (CAN) current and mGluR1a activation (Hughes et al., 2002). In nRT neurons, the slow oscillation arises from IT and Ih, as well as being influenced by the CAN current, a Ca2+-activated K+ current, and a Na+-activated K+ current (Blethyn et al., 2006).
Pro-epileptic mechanisms: Pathological slow oscillations such as the slow spike-and-wave discharges (1.5–2.5 Hz) are observed in certain epilepsies such as Lennox-Gastaut Syndrome (Steriade and Amzica, 2003), but the mechanisms that lead to these remain largely unknown.
The Thalamus in Absence Epilepsy and Beyond
There is a growing appreciation of the genetic determinants underlying a number of epilepsies linked to thalamic dysfunction. For instance, childhood absence epilepsy (CAE) has been linked with multiple susceptibility loci and polymorphisms in genes encoding voltage-gated calcium channels and GABA receptors, which will be discussed below.
Another epileptic condition also associated with aberrant thalamocortical network activity is Lennox-Gastaut syndrome (LGS), a severe childhood epileptic encephalopathy characterized by multiple seizure types (Scheffer et al., 2017). Although LGS was not traditionally recognized as having a genetic etiology, recent advances in whole-exome sequencing have identified genetic variations in LGS patients, including loci encoding voltage-gated calcium channels (Yang et al., 2021). In LGS, epileptic activity is detected both in the cortex and in the centromedian thalamic nucleus, during generalized paroxysmal fast activity and slow spike-and-wave (1–2 Hz) epileptic discharges (Velasco et al., 1991; Dalic et al., 2020; Warren et al., 2020). Notably, high-frequency deep brain stimulation of the centromedian thalamus has been shown to alleviate seizure frequency in some LGS patients (Velasco et al., 2006; Fisher and Velasco, 2014; Warren et al., 2020), confirming the importance of the thalamus in this disorder.
In the case of Dravet syndrome, a severe myoclonic epilepsy of infancy and epileptic encephalopathy with a known monogenic mutation in SCN1A (Claes et al., 2001; Dravet, 2011) (further described later), a small but notable clinical finding has implicated the potential modulatory role of the thalamus: deep brain stimulation of the centromedian thalamus resulted in seizure reduction in one of two patients with Dravet syndrome (Andrade et al., 2010).
In this section, we will discuss findings from rodent models of genetic epilepsies and highlight the diversity of etiologies which feature some extent of disruption to thalamic firing and connectivity (reviewed by Beenhakker and Huguenard, 2009; Cheong and Shin, 2013; Maheshwari and Noebels, 2014; Paz and Huguenard, 2015; Fogerson and Huguenard, 2016; Gobbo, Scheller, and Kirchhoff, 2021). In particular, we will highlight disruptions of the rhythmogenic machinery in thalamic neurons in two epilepsies: CAE, for which there is an abundance of evidence for thalamic involvement, and Dravet syndrome, for which there exists limited but convincing evidence for thalamic involvement. Understanding how various genetic etiologies converge upon a similar phenotype can pinpoint elements of vulnerability, resilience, redundancy, necessity, and sufficiency in the thalamocortical circuit, and help enumerate strategies to intervene, prevent, or compensate for epileptic activity.
Insights from Human Genetic Studies of Absence Epilepsy
The role of the thalamus in epilepsy has been best characterized in the context of typical absence seizures (formerly known as petit mal seizures), which occur in several idiopathic generalized epilepsies such as CAE and juvenile epilepsies (Scheffer et al., 2017). Absence seizures are characterized by generalized SWDs which occur alongside abrupt behavioral arrest and loss of consciousness. SWDs consist of stereotyped spike-and-wave signatures which occur at 3–10 Hz, depending on the species (generally, 6–10 Hz in rodents, and 3–4.5 Hz in humans; reviewed by Crunelli and Leresche, 2002; Blumenfeld, 2005; Tenney and Glauser, 2013; Fogerson and Huguenard, 2016), typically lasts 2–30 seconds, and occur frequently (few to several hundreds of times a day; Panayiotopoulos, 2008). SWDs are bilateral, high-amplitude waveforms, easily detected by cortical EEG, and are well-understood to reflect an aberrant, hypersynchronous thalamocortical network.
CAE includes a group of genetically determined epilepsies—some of which show strong familial inheritance and incomplete penetrance—with SWDs being inherited as an autosomal-dominant feature (Crunelli and Leresche, 2002). Numerous susceptibility loci and polymorphisms with various extents of linkage to CAE have been identified (reviewed by Crunelli and Leresche, 2002; Maheshwari and Noebels, 2014), and the evidence indicates that absence epilepsies are complex polygenic disorders. CAE mutations have been identified in voltage-gated calcium channel subunits CACNA1A, CACNA1G, CACNA1H, and CACNG3, some of which are described further later (Jouvenceau et al., 2001; Chen et al., 2003; Liang et al., 2006; Everett et al., 2007). Mutations in GABA receptor subunit genes have also been identified in human genetic studies of CAE, including GABRG2 (Baulac et al., 2001; Wallace et al., 2001), GABRA1, and GABRB3 (reviewed by Cheong and Shin, 2013).
While genetic mutations identified in human patients typically inform the design of preclinical disease models, in some cases, the identification of genetic mutations in humans has provided powerful support for mechanisms of seizure generation previously identified from genetic and pharmacological models in rodents in vivo and ex vivo. For example, the pro-epileptic effects of the GABRB2 mutation identified in human patients, which abolished benzodiazepine-induced potentiation of GABA currents when expressed in a heterologous cell system (Wallace et al., 2001) was consistent with the previously identified anti-oscillatory mechanism of clonazepam—a benzodiazepine anti-epileptic drug used to treat absence seizures (Farrello, 1986)—in rodent ex vivo thalamic slices wherein inadequate endogenous GABAAR-mediated inhibition of nRT neurons leads to strengthening of the synchronizing nRT-TC output (Huguenard and Prince, 1994b; Huguenard, 1999).
Absence epilepsies are rarely monogenic disorders, and variants in a given gene may not all lead to aberrant function. Thus, it is critical to understand how single-nucleotide polymorphisms (SNPs) identified within the same locus give rise to epileptic cells and circuits. For example, Vitko and colleagues carried out a functional characterization of 12 SNPs found in the CACANA1 (CaV3.2) locus of CAE patients in a heterologous cell system, and then used computational modeling to understand the impact of various SNPs on T-type calcium channel function; interestingly, some, but not all, SNPs were predicted to increase burst firing propensity (Vitko et al., 2005). Such efforts will only be facilitated by the increased access to whole-exome sequencing and patient-derived induced pluripotent stem cell platforms, and they will aid in the development of precision therapies for patients with epilepsy (Klassen et al., 2011; Noebels, 2017).
T-Type Calcium Channels in Genetic Models of Absence Epilepsy
Basic science and epilepsy research alike have been advanced by rodent models carrying monogenic mutations in genes encoding voltage-gated calcium channels, GABA receptors, and glutamate receptors subunits associated with absence epilepsy (reviewed by Maheshwari and Noebels, 2014). Such rodent models have provided a fertile foundation for the investigation of the thalamus’s contribution to aberrant oscillations, in some cases at the level of specific synapses (e.g., Paz et al., 2011).
For thalamic neurons, IT critically determines the coordinated rhythmic bursting between TC and nRT neurons (described in the sixth section). Unsurprisingly, aberrant calcium channel function—and subsequent changes to cellular and circuit excitability and/or firing—often leads to thalamic hyperexcitability and seizures. Aberrant T-type calcium channel function can arise from a mutation in a channel subunit, a compensatory reaction, and/or the interaction of multiple susceptibility variants. Here, we review diverse ways in which T-type calcium channels have been linked to absence seizures.
Two of the three T-type calcium channel subtypes have been linked to absence epilepsy in human genetic studies: gain-of-function polymorphisms in the CACNA1H gene (encoding CaV3.2) to sporadic CAE, and functional polymorphisms in the CACNA1G gene (encoding CaV3.1) to juvenile absence syndromes (reviewed by Maheshwari and Noebels, 2014). Similarly, enhanced T-type calcium channel function—and subsequently, enhanced burst firing—in thalamic neurons seems to be a common feature across multiple rodent models of absence epilepsies. For instance, nRT neurons in GAERS rats display elevated amplitude of IT (Tsakiridou et al., 1995) and increased CaV3.2 mRNA (Talley et al., 2000), owing to a gain-of-function point mutation in Cacna1h (Powell et al., 2009). And CaV3.1–/– mice—in which burst firing, but not tonic firing, is abolished in TC neurons—are resistant to SWDs induced by baclofen (GABABR agonist) and gamma-hydroxybutyrate (Kim et al., 2001). Accordingly, overexpression of CaV3.1 results in increased T-type calcium currents in TC neurons (although rebound burst firing was not measured) and spontaneous, ethosuximide-sensitive absence seizures (Ernst et al., 2009). The latter two did not investigate the nRT, because it expressed very low levels of endogenous and exogenous CaV3.1. Nevertheless, we cannot rule out the possibility that nRT neurons made post-translational compensatory changes in T-type calcium channels in these conditions.
Even when T-type calcium channels are not directly affected by mutations, they appear to be critical components of spontaneous absence seizures in genetically determined rodent models. Across several inbred mouse strains exhibiting absence epilepsy and ataxia, spontaneous mutations have been identified in various subunits of P/Q-type calcium channels (e.g., CaV2.1/α1A subunit in tottering mice; β4 subunit in lethargic mice; and γ2 subunit in stargazer mice (reviewed by Maheshwari and Noebels, 2014); notably, the Cacna1a mutation was identified in tottering mice prior to the first discovery of a CACNA1A mutation in cases of CAE (Fletcher et al., 1996; Jouvenceau et al., 2001). Interestingly, in tottering, lethargic, and stargazer mice with mutations in P/Q-type calcium channels, T-type calcium channels exhibit compensatory changes in their electrophysiological properties despite no changes in mRNA expression of their α1 subunits (Zhang et al., 2002). In particular, whole-cell recordings of TC neurons in thalamic slice preparations demonstrated increased peak current density and depolarized shifts in the steady-state inactivation curve (i.e., increased availability of T-type calcium channel availability upon activation threshold) (Zhang et al., 2002). Although IT properties of nRT neurons were not assessed, the changes observed in TC neurons were hypothesized to favor the enhanced cellular excitability, rhythmic bursting, and overall propensity to generate SWDs observed in these mice. Indeed, a later study cross-bed these P/Q-channel-deficient mice with T-channel- deficient mice and provided compelling evidence that T-type CaV3.1/α1G is a crucial component of the absence seizure phenotype in these mice: when harboring double null mutations in both CaV2.1/α1A and CaV3.1/α1G subunits, TC neurons are depleted of IT and these mice displayed no SWDs (Song et al., 2004). Yet another example are the HCN2-deficient mice which exhibit spontaneous absence seizures (Ludwig et al., 2003). TC neurons in HCN2–/– mice exhibit a complete loss of Ih, which leads to TC neuron hyperpolarization, which facilitates T-type calcium channel de-inactivation and therefore facilitates burst firing; ex vivo thalamic slices exhibited hypersynchronous oscillatory activity which was blocked by a T-type calcium channel inhibitor (Ludwig et al., 2003). Finally, ethosuximide, a T-type calcium channel antagonist, reduces absence seizures in GAERS rats (Polack and Charpier, 2009). Altogether, these studies provide strong support for the critical role of T-type calcium channels in absence seizures.
Nonetheless, several studies have called into question the extent to which IT-mediated burst firing in various components of single TC versus nRT cells is necessary for the generation of absence seizures (reviewed by Crunelli and Leresche, 2002; Cheong and Shin, 2013; Maheshwari and Noebels, 2014). For example, in GAERS rats and in a pharmacological model of absence seizures, McCafferty and colleagues have reported infrequent burst firing in TC neurons during absence seizures but an increased burst firing in nRT neurons (McCafferty et al., 2018). They suggested that despite the rare burst firing of TC neurons, overall thalamic output (i.e., shaped by both TC and nRT) remained strong and rhythmic during seizures (McCafferty et al., 2018). An alternative interpretation is that it is IT-mediated burst firing in the nRT neurons, and not the TC neurons, that is critical to the expression of absence seizures in these rat models. On the other hand, Lee and colleagues have reported that burst firing in nRT neurons is not essential for pharmacologically induced absence seizures in mice (Lee et al., 2014). Mice harboring double mutations of CaV3.2 and CaV3.3 displayed enhanced pharmacologically induced SWDs despite the complete abolishment of nRT burst firing; the authors attributed the SWD susceptibility to an increase in nRT tonic firing (Lee et al., 2014).
In spite of these controversies, there is abundant evidence for the powerful role of T-type calcium channels—and their disruptions—in mediating thalamic hyperexcitability and SWDs.
R-Type and P/Q-Type Channels in Genetic Models of Absence Epilepsy
Although T-type calcium channels have sustained the most attention in the study of thalamic rhythms and seizures in the past decades, the role of R-type and P/Q-type calcium channels in thalamic hypersynchrony and hyperexcitability cannot be discounted. Loss-of-function mutations in CACNA1A (encoding the CaV2.1 subunit of P/Q-type channels) have been identified in CAE (Jouvenceau et al., 2001). In addition, de novo pathogenic variants in CACNA1E (encoding the CaV2.3 subunit of R-type channels) have recently been identified in patients with developmental and epileptic encephalopathies (Helbig et al., 2018). Although these patients were not reported to exhibit absence seizures, the CaV2.3 subunit is highly expressed in nRT neurons (Soong et al., 1993; Weiergräber et al., 2006), and it has recently been shown to play an important role in oscillatory burst firing of the nRT and in pharmacologically induced SWDs (Zaman et al., 2011; Paz and Huguenard, 2012). R-type calcium channels are structurally similar to HVA calcium channels, yet exhibit electrophysiological properties closer to T-type calcium channels; a critical difference, however, is that R-type channels are activated at higher thresholds than T-type channels (Soong et al., 1993; Randall and Tsien, 1997). In nRT neurons of CaV2.3–/– mice, or nRT neurons of WT mice in the presence of SNX-482 (a specific blocker of R-type channels), reduced AHP rendered them unable to generate oscillatory bursts. Notably, genetic deletion or pharmacological blockade of R-type channels promoted resilience to gamma-hydroxybutyrate-induced SWDs in mice, suggesting that nRT oscillatory bursts, enabled by R-type mediated-AHP, are critical for generation of SWDs (Zaman et al., 2011).
P/Q-type channels are widely expressed throughout the central nervous system but especially abundant in presynaptic terminals in Purkinje and granule cells of the cerebellum. Unlike postsynaptic T-type calcium channels which play a major role in shaping firing properties, HVA P/Q-type channels are primarily involved in coupling calcium influx to neurotransmitter release in presynaptic terminals (reviewed by Rajakulendran, Kaski, and Hanna, 2012). Loss-of-function mutations in various subunits of P/Q-type channels lead to absence epilepsy, in some cases accompanied by ataxia (Noebels, 2012); the enhancement of T-type channel function in mice harboring these mutations was described earlier. Interestingly, investigations of rodent absence models harboring P/Q-type loss-of-function mutations have highlighted another source of vulnerability at the synaptic level: an incomplete overlap of P/Q-type subunit family members across the brain (Noebels, 2012). Different brain regions have distinct patterns of β1–4 subunit expression; thus, P/Q-type calcium channel composition, and thus function, is differentially affected. One of the earliest demonstrations of “subunit reshuffling” (a compensatory process) was in lethargic mice, which exhibit SWDs and ataxia (Burgess et al., 1999). In the cerebellum of lethargic mice, in the absence of the β4 subunit, there is increased steady-state association of the α1A subunit with β2 and β3 subunits instead; these complexes generate functional P/Q-type channel, with normal amplitude and voltage dependence of the P-type current in dissociated cerebellar Purkinje neurons (Burgess et al., 1999). Interestingly, thalamic nuclei appear to lack significant expression of β1 and β3 subunits, thus decreasing the chance that the α1A subunit can form a functional heteromer in the absence of a functional β4 subunit (Noebels, 2012).
Thus, selective deficits or compensations at specific synapses may arise from regulatory subunit “reshuffling.” In addition, cell-type-specific alternative splicing of various subunits may also explain resilience and vulnerability (Noebels, 2012).
SK Channels and Thalamic Bursting in Dravet Syndrome
Dravet syndrome is a childhood epileptic encephalopathy associated with severe nonconvulsive seizures and sudden unexplained death in epilepsy. It is caused by loss-of- function mutations in the SCN1A gene encoding the type I voltage-gated sodium channel (Claes et al., 2001; Dravet, 2011). In mouse models of Dravet syndrome, cortical inhibitory neurons exhibit reduced excitability, consistent with a loss of function in a channel critical for action potential generation and propagation (Han et al., 2020; Rubinstein et al., 2015). Specific inactivation of Scn1a in parvalbumin-positive interneurons in the hippocampus and cortex increases seizure susceptibility (Dutton et al., 2013). Interestingly, hypoexcitability of parvalbumin-positive cortical interneurons have been reported to normalize in adulthood in mouse models, suggesting that there may be other mechanisms which explain the ongoing, chronic epilepsy in Dravet syndrome (Favero et al., 2018). Recent work by our group has uncovered a surprising role for thalamic bursting in Dravet syndrome.
Given that Scn1a is highly expressed in parvalbumin-positive nRT cells, our group investigated thalamic firing in the Scn1a-deficient mouse model of Dravet syndrome (Ritter-Makinson et al., 2019). We found that GABAergic nRT neurons—but not glutamatergic TC neurons—exhibited heightened intrinsic excitability, most notably in the form of augmented rebound burst firing. Interestingly, enhanced nRT rebound burst firing was due to a compensatory reduction in the calcium-activated small potassium current (ISK) density rather than an enhancement of the low-threshold T-type calcium current. While nRT neurons also exhibited changes expected from Scn1a loss-of-function such as depolarized action potential threshold and increased action potential width, nRT neurons fired more action potentials in response to depolarization. This finding was in contrast to the decreased excitability of cortical inhibitory neurons described in previous studies. Moreover, reducing SK conductance, but not the sodium conductance, in a biophysical model was sufficient to generate enhanced rebound bursts as observed in nRT neurons. As a consequence of SK-deficit mediated augmentation of nRT burst firing, mice exhibited thalamic microcircuit hyperexcitability, an ex vivo hallmark of seizures involving the thalamus (reviewed by Cho et al., 2017). Finally, treatment with an SK channel agonist (1-ethyl-2-benzimidazolinone, or EBIO) was sufficient to reduce the frequency of non-convulsive seizures in Dravet syndrome mice. Altogether, we propose that Scn1a loss-of-function results in a compensatory reduction of ISK in nRT neurons, which contributes to enhanced input resistance and intrinsic excitability, and increased rebound burst firing. Augmented burst firing in nRT neurons recruits stronger oscillatory bursting in the intra-thalamic circuit, which maintains nonconvulsive seizures (Ritter-Makinson et al., 2019).
This study demonstrated a novel role for the thalamus in a type of nonconvulsive seizures observed in human patients. It also points to a novel role for SK2 channels (a normal component of nRT rhythmicity; Cueni et al., 2008)) and raises questions about how this compensatory mechanism arises from a Scn1a loss of function mutation. Importantly, by tapping into the rhythmogenic machinery of thalamic neurons, this study identified the nRT as a potential novel therapeutic target to treat nonconvulsive seizures in Dravet syndrome.
The Emerging Role of the Thalamus in Acquired Epilepsies
While the thalamus has been well-studied in the context of genetic epilepsies, there is also an increasing appreciation of thalamic involvement in acquired epilepsies such as those that develop after acute insults and in mesial temporal lobe epilepsies. Following cortical injuries such as traumatic brain injury and stroke, the thalamus sustains robust and chronic secondary damage including neurodegeneration and neuroinflammation in humans (Pappata et al., 2000; Scott et al., 2015; Grossman and Inglese, 2016) and in rodent models (Paz et al., 2010; Hazra et al., 2014; Cao et al., 2020; Holden et al., 2021; Necula et al., 2021). Thalamic damage has gained increasing recognition as a prognostic biomarker of posttraumatic epileptogenesis (PTE) in rodent models (Paz et al., 2010; Immonen et al., 2013; Pitkänen et al., 2020; Manninen et al., 2021), and it has also been suggested as a biomarker of cognitive impairment after TBI in patients (Ramlackhansingh et al., 2011; Grossman et al., 2012). Despite being remote from the initial epileptogenic insult, the thalamus can modulate seizure activity in rodent models of PTE (Paz et al., 2013; Paz and Huguenard, 2015), and it is the only FDA-approved brain region for deep brain stimulation in patients with refractory temporal lobe epilepsy (Fasano et al., 2021; Fisher and Velasco, 2014).
Neuroinflammation, commonly implicated in epileptogenesis, has been notoriously challenging to dissect and target effectively (reviewed by Aronica et al., 2017; Klein et al., 2018; Patel et al., 2019; Vezzani, Balosso, and Ravizza, 2019). The extent to which neuroinflammation—specifically in the thalamus—can drive epileptogenic changes following injuries, and whether it can be a disease-modifying target in the context of PTE, is an active field of investigation.
Our group recently conducted a systematic study of secondary thalamic inflammation across four distinct thalamocortical circuits in two rodent models of cortical lesions and found that secondary thalamic neuroinflammation mirrors functional connectivity of thalamocortical circuits (Necula et al., 2021). Such investigations, coupled with open access to large-scale anatomical and functional connectivity atlases—for example, the Allen Mouse Brain Connectivity Atlas (Harris et al., 2019), the Janelia MouseLight project (Winnubst et al., 2019), and the Allen Brain Observatory (Siegle et al., 2021)—may help start to elucidate potential “rules” by which secondary neuroinflammation evolves after cortical injuries.
Neuroinflammation following stroke or TBI colocalizes with robust changes in cellular and circuit properties of the thalamus. In rodent models, cortical stroke and TBI both lead to thalamic circuit hyperexcitability and thalamocortical epileptic discharges (Paz et al., 2010, 2013; Holden et al., 2021). Our group recently identified the C1q complement pathway as a mediator of the chronic emergence of epileptic activity and sleep disruption after a mild TBI in mice (Holden et al., 2021), highlighting the feasibility and promise of targeting thalamic neuroinflammation to prevent circuit dysfunction after injuries. Altogether, these findings support the role of thalamic neuroinflammation as a meaningful driver of chronic pathological changes implicated in epileptogenesis (Cho et al., 2022) and motivate further dissection of neuronal-glial interactions in the secondarily damaged thalamus.
Conclusions
In summary, we have reviewed evidence of pro-epileptic disruptions in thalamic and thalamocortical circuits, and propose that the thalamus as a key convergence point of dysfunction in genetic epilepsies. Understanding how the thalamus generates and modulates aberrant activity will aid in the identification of therapeutic targets and paradigms to treat epilepsies.
What aspects of thalamic organization and function predispose the thalamus to vulnerability? The thalamus generates vigorous rhythmic burst firing at the cellular and microcircuit levels, and the mechanics underlying such firing appears to be vulnerable to genetic perturbations. However, whether that means they can just as easily be “corrected,” without equally debilitating off-target effects, remains to be seen.
Cell-intrinsic vulnerabilities: We reviewed diverse ways in which T-type calcium channels, as well as R-type and P/Q-type channels, have been linked to absence seizures, including mutations which lead to direct or compensatory increases in the magnitude of IT, or to changes in channel subunit expression and composition. We particularly highlighted the abundant evidence that burst firing properties of thalamic neurons appear to be a common node of dysfunction across genetic models of epilepsies. One novel and notable example is the identification of a compensatory reduction of the SK current and subsequent increase of nRT burst firing in a Scn1a loss-of-function mouse model of Dravet syndrome (Ritter-Makinson et al., 2019). The finding that pharmacological augmentation of the SK current reduced the severity of nonconvulsive seizures suggests that restoring burst firing, even when it may not be the sole or primary electrophysiological disruption, can be potentially beneficial in modulating seizures.
Micro-circuit vulnerabilities: We also reviewed the diverse ways in which disruptions to feedforward inhibition and counter-inhibition increase the propensity of thalamic circuits to become hypersynchronous and hyperexcitable. Further investigations of what mediates selective vulnerability of specific synapses, and whether restoration of unperturbed microcircuit functions can compensate for pro-epileptic changes, will undoubtedly be fruitful in identifying therapeutic strategies to prevent and treat seizures.
Are there similar vulnerabilities in the thalamocortical circuit across genetic and acquired epilepsies and what are they? Thalamic damage after cortical injuries has recently emerged as a biomarker of PTE. Due to its broad connectivity, focal and even remote damage to the thalamus is well-positioned to mediate disruptions in “bystander” circuits which are not immediately involved in seizure dynamics. Thus, recent advances in our understanding of thalamic organization and function in various psychiatric and cognitive conditions (e.g., Krol et al., 2018)—which can often be comorbidities of seizures—may be harnessed to understand and modulate the potentially far-reaching role (beyond seizure activity) of the thalamus in acquired epilepsies. Ultimately, if the thalamus is indeed an Achilles’ heel—a critical node of dysfunction—in the context of epilepsy, then understanding how diverse genetic and acquired etiologies converge upon thalamic hyperexcitability will pinpoint elements that contribute to its vulnerability as well as those that contribute to its resilience.
Acknowledgments
J. T. P. is supported by NIH/NINDS grant R01NS096369, the Department of Defense (EP150038), the Gladstone Institutes, and the Kavli Institute for Fundamental Neuroscience. F. S. C. is supported by NINDS F31 NS111819-01A1. The authors would like to thank Drs. Francoise Chanut and Kathryn Claiborn for editorial assistance. F. S. C. would also like to thank Dr. Tilo Gschwind for constructive feedback and insightful discussions.
References
- 1. Jasper HH. Charting the Sea of Brain Waves. Science. 1948;108 (2805):343–347. PMID: 17810990 [PubMed: 17810990]
- 2. Jasper HH. Current evaluation of the concepts of centrecephalic and cortico-reticular seizures. Electroencephalogr Clin Neurophysiol. 1991;78 (1):2–11. PMID: 1701711 [PubMed: 1701711]
- 3. Spiegel EA, Wycis HT. Thalamic recordings in man with special reference to seizure discharges. Electroencephalogr Clin Neurophysiol. 1950;2 (1–4):23–27. PMID: not available. https://doi
.org/10.1016 /0013-4694(50)90003-4 - 4. Williams D. A Study of Thalamic and Cortical Rhythms in Petit Mal. Brain. 1953;76 (1):50–69. PMID: 13041922 [PubMed: 13041922]
- 5. Sorokin JM, Davidson TJ, Frechette E, Abramian AM, Deisseroth K, Huguenard JR, Paz JT. Bidirectional Control of Generalized Epilepsy Networks via Rapid Real-Time Switching of Firing Mode. Neuron. Elsevier Inc.; 2017;93(1):194–210. PMID: 27989462 [PMC free article: PMC5268077] [PubMed: 27989462]
- 6. Jones EG. The Thalamus. Plenum Press; 1985. PMID: not available. https://doi
.org/10.1007 /978-1-4615-1749-8 - 7. Fogerson PM, Huguenard JR. Tapping the Brakes: Cellular and Synaptic Mechanisms that Regulate Thalamic Oscillations. Neuron. Elsevier Inc.; 2016;92 (4):687–704. PMID: 27883901 [PMC free article: PMC5131525] [PubMed: 27883901]
- 8. Steriade M, McCormick DA, Sejnowski TJ. Thalamocortical oscillations in the sleeping and aroused brain. Science. 1993;262 (5134):679–85. PMID: 8235588 [PubMed: 8235588]
- 9. Paz JT, Davidson TJ, Frechette ES, Delord B, Parada I, Peng K, Deisseroth K, Huguenard JR. Closed-loop optogenetic control of thalamus as a tool for interrupting seizures after cortical injury. Nat Neurosci. 2013;16(1):64–70. PMID: 23143518 [PMC free article: PMC3700812] [PubMed: 23143518]
- 10. Paz JT, Huguenard JR. Microcircuits and their interactions in epilepsy: is the focus out of focus? Nat Neurosci. 2015;18 (3):351–9. PMID: 25710837 [PMC free article: PMC4561622] [PubMed: 25710837]
- 11. Fisher RS, Velasco AL. Electrical brain stimulation for epilepsy. Nat Rev Neurol. Nature Publishing Group; 2014;10(5):261–270. PMID: 24709892 [PubMed: 24709892]
- 12. Halassa MM, Sherman SM. Thalamocortical Circuit Motifs: A General Framework. Neuron. Elsevier Inc.; 2019;103 (5):762–770. PMID: 31487527 [PMC free article: PMC6886702] [PubMed: 31487527]
- 13. Shepherd GMG, Yamawaki N. Untangling the cortico-thalamo-cortical loop: cellular pieces of a knotty circuit puzzle. Nat Rev Neurosci. Springer US; 2021;22(7):389–406. PMID: 33958775 [PMC free article: PMC9006917] [PubMed: 33958775]
- 14. Huguenard JR, McCormick DA. Thalamic synchrony and dynamic regulation of global forebrain oscillations. Trends Neurosci. 2007;30 (7):350–356. PMID: 17544519 [PubMed: 17544519]
- 15. Crunelli V, Hughes SW. The slow (1 Hz) rhythm of non-REM sleep: A dialogue between three cardinal oscillators. Nat Neurosci. Nature Publishing Group; 2010;13(1):9–17. PMID: 19966841 [PMC free article: PMC2980822] [PubMed: 19966841]
- 16. Beenhakker MP, Huguenard JR. Neurons that Fire Together Also Conspire Together: Is Normal Sleep Circuitry Hijacked to Generate Epilepsy? Neuron. Elsevier Inc.; 2009;62 (5):612–632. PMID: 19524522 [PMC free article: PMC2748990] [PubMed: 19524522]
- 17. Steriade M, Llinas R. The functional states of the thalamus and the associated neuronal interplay. Physiol Rev. 1988;68 (3):648–742. PMID: 2839857 [PubMed: 2839857]
- 18. McCormick DA, Bal T. Sleep and arousal: Thalamocortical mechanisms. Annu Rev Neurosci. 1997;20:185–215. PMID: 9056712 [PubMed: 9056712]
- 19. Crunelli V, Larincz ML, Connelly WM, David F, Hughes SW, Lambert RC, Leresche N, Errington AC. Dual function of thalamic low-vigilance state oscillations: Rhythm-regulation and plasticity. Nat Rev Neurosci. 2018; 19 (2):107–118. PMID: 29321683 [PMC free article: PMC6364803] [PubMed: 29321683]
- 20. Fernandez LMJ, Lüthi A. Sleep spindles: Mechanisms and functions. Physiol Rev. 2020;100(2):805–868. PMID: 31804897 [PubMed: 31804897]
- 21. Jones EG. The Thalamus. Cambridge, UK: Cambridge University Press; 2007.
- 22. Deschênes M, Veinante P, Zhang ZW. The organization of corticothalamic projections: Reciprocity versus parity. Brain Res Rev. 1998;28 (3):286–308. PMID: 9858751 [PubMed: 9858751]
- 23. Charpier S, Beurrier C, Paz JT. The Subthalamic Nucleus: From In Vitro to In Vivo Mechanisms. Handbook of Behavioral Neuroscience. Elsevier Inc.; 2010.
- 24. Paz JT, Chavez M, Saillet S, Deniau J, Charpier S. Activity of Ventral Medial Thalamic Neurons during Absence Seizures and Modulation of Cortical Paroxysms by the Nigrothalamic Pathway. J Neurosci. 2007;27 (4):929–941. PMID: 17251435 [PMC free article: PMC6672924] [PubMed: 17251435]
- 25. Houser CR, Vaughn JE, Barber RP, Roberts E. GABA Neurons are the Major Cell Type of the Nucleus Reticularis Thalami. Brain Res. 1980;200:341–354. PMID: 25246403 [PubMed: 7417821]
- 26. Pinault D. The thalamic reticular nucleus: Structure, function and concept. Brain Res Rev. 2004;46 (1):1–31. PMID: 15297152 [PubMed: 15297152]
- 27. Sherman SM, Guillery RW. Functional organization of thalamocortical relays. J Neurophysiol. 1996;76(3):1367–1395. PMID: 8890259 [PubMed: 8890259]
- 28. Guillery RW, Sherman SM. Thalamic relay functions and their role in corticocortical communication: Generalizations from the visual system. Neuron. 2002;33 (2):163–175. PMID: 11804565 [PubMed: 11804565]
- 29. Jones EG. Some aspects of the organization of the thalamic reticular complex. J Comp Neurol. 1975;162 (3):285–308. PMID: 1150923 [PubMed: 1150923]
- 30. Halassa MM, Acsády L. Thalamic Inhibition: Diverse Sources, Diverse Scales. Trends Neurosci. Elsevier Ltd; 2016;39 (10):680–693. PMID: 27589879 [PMC free article: PMC5048590] [PubMed: 27589879]
- 31. Crabtree JW. Functional diversity of thalamic reticular subnetworks. Front Syst Neurosci. 2018;12(October):1–18. PMID: 30405364 [PMC free article: PMC6200870] [PubMed: 30405364]
- 32. Williamson AM, Ohara PT, Ralston HJ. Electron microscopic evidence that cortical terminals make direct contact onto cells of the thalamic reticular nucleus in the monkey. Brain Res. 1993;631 (1):175–179. PMID: 7507789 [PubMed: 7507789]
- 33. Liu XB, Jones EG. Predominance of corticothalamic synaptic inputs to thalamic reticular nucleus neurons in the rat. J Comp Neurol. 1999;414 (1):67–79. PMID: 10494079 [PubMed: 10494079]
- 34. Asanuma C. GABAergic and pallidal terminals in the thalamic reticular nucleus of squirrel monkeys. Exp Brain Res. 1994;101 (3):439–451. PMID: 7531651 [PubMed: 7531651]
- 35. Paré D, Hazrati LN, Parent A, Steriade M. Substantia nigra pars reticulata projects to the reticular thalamic nucleus of the cat: a morphological and electrophysiological study. Brain Res. 1990;535 (1):139–146. PMID: 1705469 [PubMed: 1705469]
- 36. Jourdain A, Semba K, Fibiger HC. Basal forebrain and mesopontine tegmental projections to the reticular thalamic nucleus: an axonal collateralization and immunohistochemical study in the rat. Brain Res. 1989;505 (1):55–65. PMID: 2575437 [PubMed: 2575437]
- 37. Pinault D, Bourassa J, Deschênes M. The Axonal Arborization of Single Thalamic Reticular Neurons in the Somatosensory Thalamus of the Rat. Eur J Neurosci. 1995;7 (1):31–40. PMID: 7711934 [PubMed: 7711934]
- 38. Clemente-Perez A, Makinson SR, Higashikubo B, Brovarney S, Cho FS, Urry A, Holden SS, Wimer M, Dávid C, Fenno LE, Acsády L, Deisseroth K, Paz JT. Distinct thalamic reticular cell types differentially modulate normal and pathological cortical rhythms. Cell Rep. 2017;19 (10):2130–2142. PMID: 28591583 [PMC free article: PMC5557038] [PubMed: 28591583]
- 39. Harris JA, Mihalas S, Hirokawa KE, Whitesell JD, Choi H, Bernard A, Bohn P, Caldejon S, Casal L, Cho A, Feiner A, Feng D, Gaudreault N, Gerfen CR, Graddis N, Groblewski PA, Henry AM, Ho A, Howard R, Knox JE, Kuan L, Kuang X, Lecoq J, Lesnar P, Li Y, Luviano J, McConoughey S, Mortrud MT, Naeemi M, Ng L, Oh SW, Ouellette B, Shen E, Sorensen SA, Wakeman W, Wang Q, Wang Y, Williford A, Phillips JW, Jones AR, Koch C, Zeng H. Hierarchical organization of cortical and thalamic connectivity. Nature. Springer US; 2019;575(7781):195–202. PMID: 31666704 [PMC free article: PMC8433044] [PubMed: 31666704]
- 40. Phillips J W, Schulmann A, Hara E, Winnubst J, Liu C, Valakh V, Wang L, Shields BC, Korff W, Chandrashekar J, Lemire AL, Mensh B, Dudman JT, Nelson SB, Hantman AW. A repeated molecular architecture across thalamic pathways. Nat Neurosci. Springer US; 2019;22(11):1925–1935. PMID: 31527803 [PMC free article: PMC6819258] [PubMed: 31527803]
- 41. Mo C, Petrof I, Viaene AN, Sherman SM. Synaptic properties of the lemniscal and paralemniscal pathways to the mouse somatosensory thalamus. Proc Natl Acad Sci U S A. 2017;114 (30):E6212–E6221. PMID: 28696281 [PMC free article: PMC5544298] [PubMed: 28696281]
- 42. Jones EG. Synchrony in the interconnected circuitry of the thalamus and cerebral cortex. Ann N Y Acad Sci. 2009;1157:10–23. PMID: 19351352 [PubMed: 19351352]
- 43. Martinez-Garcia RI, Voelcker B, Zaltsman JB, Patrick SL, Stevens TR, Connors BW, Cruikshank SJ. Two dynamically distinct circuits drive inhibition in the sensory thalamus. Nature. Springer US; 2020;583 (7818):813–818. PMID: 32699410 [PMC free article: PMC7394732] [PubMed: 32699410]
- 44. Li Y, Lopez-Huerta VG, Adiconis X, Levandowski K, Choi S, Simmons SK, Arias-Garcia MA, Guo B, Yao AY, Blosser TR, Wimmer RD, Aida T, Atamian A, Naik T, Sun X, Bi D, Malhotra D, Hession CC, Shema R, Gomes M, Li T, Hwang E, Krol A, Kowalczyk M, Peça J, Pan G, Halassa MM, Levin JZ, Fu Z, Feng G. Distinct subnetworks of the thalamic reticular nucleus. Nature. Springer US; 2020;583(7818):819–824. PMID: 32699411 [PMC free article: PMC7394718] [PubMed: 32699411]
- 45. Sherman SM. The thalamus is more than just a relay. Curr Opin Neurobiol. 2007;17 (4):417–422. PMID: 17707635 [PMC free article: PMC2753250] [PubMed: 17707635]
- 46. Sohal VS, Huguenard JR. Inhibitory interconnections control burst pattern and emergent network synchrony in reticular thalamus. J Neuros. 2003;23(26):8978–8988. PMID: 14523100 [PMC free article: PMC6740384] [PubMed: 14523100]
- 47. Huntsman MM, Porcello DM, Homanics GE, DeLorey TM, Huguenard JR. Reciprocal Inhibitory Connections and Network Synchrony in the Mammalian Thalamus. Science. 1999;283 (5401):541–543. PMID: 9915702 [PubMed: 9915702]
- 48. Makinson CD, Tanaka BS, Sorokin JM, Wong JC, Christian CA, Goldin AL, Escayg A, Huguenard JR. Regulation of Thalamic and Cortical Network Synchrony by Scn8a. Neuron. Elsevier Inc.; 2017;93 (5):1165–1179.e6. PMID: 28238546 [PMC free article: PMC5393918] [PubMed: 28238546]
- 49. Landisman CE, Long MA, Beierlein M, Deans MR, Paul DL, Connors BW. Electrical synapses in the thalamic reticular nucleus. J Neurosci. 2002;22 (3):1002–1009. PMID: 11826128 [PMC free article: PMC6758490] [PubMed: 11826128]
- 50. Long MA, Landisman CE, Connors BW. Small Clusters of Electrically Coupled Neurons Generate Synchronous Rhythms in the Thalamic Reticular Nucleus. J Neurosci. 2004;24(2):341–349. PMID: 14724232 [PMC free article: PMC6729997] [PubMed: 14724232]
- 51. Deleuze C, Huguenard JR. Distinct electrical and chemical connectivity maps in the thalamic reticular nucleus: Potential roles in synchronization and sensation. J Neurosci. 2006;26(33):8633–8645. PMID: 16914689 [PMC free article: PMC6674339] [PubMed: 16914689]
- 52. Shu Y, McCormick DA. Inhibitory interactions between ferret thalamic reticular neurons. J Neurophysiol. 2002;87 (5):2571–2576. PMID: 11976393 [PubMed: 11976393]
- 53. Hou G, Smith AG, Zhang ZW. Lack of intrinsic GABAergic connections in the thalamic reticular nucleus of the mouse. J Neurosci. 2016;36(27):7246–7252. PMID: 27383598 [PMC free article: PMC4938865] [PubMed: 27383598]
- 54. Cruikshank SJ, Urabe H, Nurmikko A V., Connors BW. Pathway-Specific Feedforward Circuits between Thalamus and Neocortex Revealed by Selective Optical Stimulation of Axons. Neuron. Elsevier Ltd; 2010;65 (2):230–245. PMID: 20152129 [PMC free article: PMC2826223] [PubMed: 20152129]
- 55. Parker PRL, Cruikshank SJ, Connors BW. Stability of electrical coupling despite massive developmental changes of intrinsic neuronal physiology. J Neurosci. 2009;29 (31):9761–9770. PMID: 19657029 [PMC free article: PMC3353772] [PubMed: 19657029]
- 56. Paz JT, Bryant AS, Peng K, Fenno L, Yizhar O, Frankel WN, Deisseroth K, Huguenard JR. A new mode of corticothalamic transmission revealed in the Gria4(-/-) model of absence epilepsy. Nat Neurosci. Nature Publishing Group; 2011;14(9):1167–73. PMID: 21857658 [PMC free article: PMC3308017] [PubMed: 21857658]
- 57. Sohal VS, Huntsman MM, Huguenard JR. Reciprocal Inhibitory Connections Regulate the Spatiotemporal Properties of Intrathalamic Oscillations. J Neurosci. 2000;20 (5):1735–1745. PMID: 10684875 [PMC free article: PMC6772943] [PubMed: 10684875]
- 58. Jahnsen H, Llinás R. Ionic Basis for the Electroresponsiveness. J Physiol. 1984;349:227–247. PMID: 6737293 [PMC free article: PMC1199335] [PubMed: 6737293]
- 59. Huguenard JR. Low-Threshold Calcium Currents in Central Nervous System Neurons. Annu Rev Physiol. 1996;58 (1):329–348. PMID: 8815798 [PubMed: 8815798]
- 60. Sherman SM. A wake-up call from the thalamus. Nat Neurosci. 2001;4 (4):344–346. PMID: 11276218 [PubMed: 11276218]
- 61. Wolfart J, Debay D, Le Masson G, Destexhe A, Bal T. Synaptic background activity controls spike transfer from thalamus to cortex. Nat Neurosci. 2005;8 (12):1760–1767. PMID: 16261132 [PubMed: 16261132]
- 62. Swadlow HA, Gusev AG, Bezdudnaya T. Activation of a cortical column by a thalamocortical impulse. J Neurosci. 2002;22 (17):7766–7773. PMID: 12196600 [PMC free article: PMC6757983] [PubMed: 12196600]
- 63. Llinás RR, Steriade M. Bursting of thalamic neurons and states of vigilance. J Neurophysiol. 2006;95(6):3297–308. PMID: 16554502 [PubMed: 16554502]
- 64. Astori S, Wimmer RD, Prosser HM, Corti C, Corsi M, Liaudet N, Volterra A, Franken P, Adelman JP, Lüthi A. The Ca v3.3 calcium channel is the major sleep spindle pacemaker in thalamus. Proc Natl Acad Sci U S A. 2011;108 (33):13823–13828. PMID: 21808016 [PMC free article: PMC3158184] [PubMed: 21808016]
- 65. Tenney JR, Glauser TA. The current state of absence epilepsy: Can we have your attention? Epilepsy Curr. 2013;13(3):135–140. PMID: 23840175 [PMC free article: PMC3697883] [PubMed: 23840175]
- 66. Crunelli V, Leresche N. Childhood absence epilepsy: Genes, channels, neurons and networks. Nat Rev Neurosci. 2002;3 (5):371–382. PMID: 11988776 [PubMed: 11988776]
- 67. Blumenfeld H. Cellular and Network Mechanisms of Spike-Wave Seizures. Epilepsia. 2005;46:21–33. PMID: 16302873 [PubMed: 16302873]
- 68. Inoue M, Duysens J, Vossen JMH, Coenen AML. Thalamic multiple-unit activity underlying spike-wave discharges in anesthetized rats. Brain Res. 1993;612 (1–2):35–40. PMID: 8330210 [PubMed: 8330210]
- 69. Slaght SJ, Leresche N, Deniau J-M, Crunelli V, Charpier S. Activity of thalamic reticular neurons during spontaneous genetically determined spike and wave discharges. J Neurosci. 2002;22 (6):2323–2334. PMID: 11896171 [PMC free article: PMC6758255] [PubMed: 11896171]
- 70. Huguenard J. Current Controversy: Spikes, Bursts, and Synchrony in Generalized Absence Epilepsy: Unresolved Questions Regarding Thalamocortical Synchrony in Absence Epilepsy. Epilepsy Curr. 2019;19(2):105–111. PMID: 30955423 [PMC free article: PMC6610415] [PubMed: 30955423]
- 71. Yizhar O, Fenno LE, Prigge M, Schneider F, Davidson TJ, O’Shea DJ, Sohal VS, Goshen I, Finkelstein J, Paz JT, Stehfest K, Fudim R, Ramakrishnan C, Huguenard JR, Hegemann P, Deisseroth K. Neocortical excitation/inhibition balance in information processing and social dysfunction. Nature. Nature Publishing Group; 2011;477 (7363):171–178. PMID: 21796121 [PMC free article: PMC4155501] [PubMed: 21796121]
- 72. Ritter-Makinson S, Clemente-Perez A, Higashikubo B, Cho FS, Holden SS, Bennett E, Chkaidze A, Rooda OHJE, Cornet MC, Hoebeek FE, Yamakawa K, Cilio MR, Delord B, Paz JT. Augmented Reticular Thalamic Bursting and Seizures in Scn1a-Dravet Syndrome. Cell Rep. ElsevierCompany.; 2019;26 (1):54–64. PMID: 30673603 [PMC free article: PMC6555418] [PubMed: 30605686]
- 73. McCafferty C, David F, Venzi M, Lorincz ML, Delicata F, Atherton Z, Recchia G, Orban G, Lambert RC, Di Giovanni G, Leresche N, Crunelli V. Cortical drive and thalamic feed-forward inhibition control thalamic output synchrony during absence seizures. Nat Neurosci. Springer US; 2018;21 (5):744–756. PMID: 29662216 [PMC free article: PMC6278913] [PubMed: 29662216]
- 74. Nowycky MC, Fox AP, Tsien RW. Three types of neuronal calcium channel with different calcium agonist sensitivity. Nature. 1985;316 (1):440–443. PMID: 2410796 [PubMed: 2410796]
- 75. Rajakulendran S, Kaski D, Hanna MG. Neuronal P/Q-type calcium channel dysfunction in inherited disorders of the CNS. Nat Rev Neurol. Nature Publishing Group; 2012;8 (2):86–96. PMID: 22249839 [PubMed: 22249839]
- 76. Perez-Reyes E. Molecular physiology of low-voltage-activated T-type calcium channels. Physiol Rev. 2003;83(1):117–161. PMID: 12506128 [PubMed: 12506128]
- 77. Soong TW, Stea A, Hodson CD, Dubel SJ, Vincent SR, Snutch TP. Structure and functional expression of a member of the low voltage - Activated calcium channel family. Science. 1993;260 (5111):1133–1136. PMID: 8388125 [PubMed: 8388125]
- 78. Randall AD, Tsien RW. Contrasting biophysical and pharmacological properties of T-type and R-type calcium channels. Neuropharmacology. 1997;36 (7):879–893. PMID: 9257934 [PubMed: 9257934]
- 79. Catterall WA. Structure and Regulation of Voltage-Gated Ca2+ Channels. Annu Rev Cell Dev Biol. 2000;16 (521):555. PMID: 11031246 [PubMed: 11031246]
- 80. Catterall WA, Few AP. Calcium Channel Regulation and Presynaptic Plasticity. Neuron. 2008;59 (6):882–901. PMID: 18817729 [PubMed: 18817729]
- 81. Cheong E, Shin HS. T-type Ca2+ channels in normal and abnormal brain functions. Physiol Rev. 2013;93 (3):961–992. PMID: 23899559 [PubMed: 23899559]
- 82. 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. PMID: 10066243 [PMC free article: PMC6782581] [PubMed: 10066243]
- 83. Joksovic PM, Nelson MT, Jevtovic-Todorovic V, Patel MK, Perez-Reyes E, Campbell KP, Chen CC, Todorovic SM. CaV3.2 is the major molecular substrate for redox regulation of T-type Ca2+ channels in the rat and mouse thalamus. J Physiol. 2006;574 (2):415–430. PMID: 16644797 [PMC free article: PMC1817755] [PubMed: 16644797]
- 84. Parajuli LK, Fukazawa Y, Watanabe M, Shigemoto R. Subcellular distribution of α1G subunit of T-type calcium channel in the mouse dorsal lateral geniculate nucleus. J Comp Neurol. 2010;518 (21):4362–4374. PMID: 20853512 [PubMed: 20853512]
- 85. Destexhe A, Neubig M, Ulrich D, Huguenard J. Dendritic Low-Threshold Calcium Currents in Thalamic Relay Cells. J Neurosci. 1998;18 (10):3574–3588. PMID: 9570789 [PMC free article: PMC6793130] [PubMed: 9570789]
- 86. Crandall SR, Govindaiah G, Cox CL. Low-threshold Ca2+ current amplifies distal dendritic signaling in thalamic reticular neurons. J Neurosci. 2010;30 (46):15419–15429. PMID: 21084598 [PMC free article: PMC3075467] [PubMed: 21084598]
- 87. Cueni L, Canepari M, Luján R, Emmenegger Y, Watanabe M, Bond CT, Franken P, Adelman JP, Lüthi A. T-type Ca 2+ channels, SK2 channels and SERCAs gate sleep-related oscillations in thalamic dendrites. Nat Neurosci. 2008;11 (6):683–692. PMID: 18488023 [PubMed: 18488023]
- 88. Huguenard JR, Prince DA. A novel T-type current underlies prolonged Ca2+-dependent burst firing in GABAergic neurons of rat thalamic reticular nucleus. J Neurosci. 1992;12 (10):3804–3817. PMID: 1403085 [PMC free article: PMC6575965] [PubMed: 1403085]
- 89. Crunelli V, Cope D, Hughes SW. Thalamic T-type Ca2+ channels and NREM sleep. Cell Calcium. 2006;40 (2):175–190. PMID: 16777223 [PMC free article: PMC3018590] [PubMed: 16777223]
- 90. Lein ES, Hawrylycz MJ, Ao N, Ayres M, Bensinger A, Bernard A, Boe AF, Boguski MS, Brockway KS, Byrnes EJ, Chen L, Chen L, Chen TM, Chin MC, Chong J, Crook BE, Czaplinska A, Dang CN, Datta S, Dee NR, Desaki AL, Desta T, Diep E, Dolbeare TA, Donelan MJ, Dong HW, Dougherty JG, Duncan BJ, Ebbert AJ, Eichele G, Estin LK, Faber C, Facer BA, Fields R, Fischer SR, Fliss TP, Frensley C, Gates SN, Glattfelder KJ, Halverson KR, Hart MR, Hohmann JG, Howell MP, Jeung DP, Johnson RA, Karr PT, Kawal R, Kidney JM, Knapik RH, Kuan CL, Lake JH, Laramee AR, Larsen KD, Lau C, Lemon TA, Liang AJ, Liu Y, Luong LT, Michaels J, Morgan JJ, Morgan RJ, Mortrud MT, Mosqueda NF, Ng LL, Ng R, Orta GJ, Overly CC, Pak TH, Parry SE, Pathak SD, Pearson OC, Puchalski RB, Riley ZL, Rockett HR, Rowland SA, Royall JJ, Ruiz MJ, Sarno NR, Schaffnit K, Shapovalova N V., Sivisay T, Slaughterbeck CR, Smith SC, Smith KA, Smith BI, Sodt AJ, Stewart NN, Stumpf KR, Sunkin SM, Sutram M, Tam A, Teemer CD, Thaller C, Thompson CL, Varnam LR, Visel A, Whitlock RM, Wohnoutka PE, Wolkey CK, Wong VY, Wood M, Yaylaoglu MB, Young RC, Youngstrom BL, Yuan XF, Zhang B, Zwingman TA, Jones AR. Genome-wide atlas of gene expression in the adult mouse brain. Nature. 2007;445 (7124):168–176. PMID: 17151600 [PubMed: 17151600]
- 91. Hughes SW, Lorincz ML, Blethyn K, Kékesi KA, Juhász G, Turmaine M, Parnavelas JG, Crunelli V. Thalamic gap junctions control local neuronal synchrony and influence macroscopic oscillation amplitude during EEG alpha rhythms. Front Psychol. 2011;2 (article 193):1–11. PMID: 22007176 [PMC free article: PMC3187667] [PubMed: 22007176]
- 92. Hoseini MS, Higashikubo B, Cho FS, Chang AH, Clemente-Perez A, Lew I, Ciesielska A, Stryker MP, Paz JT. Gamma rhythms and visual information in mouse V1 specifically modulated by somatostatin+ neurons in reticular thalamus. Elife. 2021;10:1–24. PMID: 33843585 [PMC free article: PMC8064751] [PubMed: 33843585]
- 93. Brown EN, Lydic R, Schiff ND. General Anesthesia, Sleep, and Coma. N Engl J Med. 2010;363(27):2638–2650. PMID: 21190458 [PMC free article: PMC3162622] [PubMed: 21190458]
- 94. McCormick DA, Pape H-C. Properties of a Hyperpolarization-Activated Cation Current and Its Role in Rhythmic Oscillation in Thalamic Relay Neurones. J Physiol. 1990; 431:291–318. PMID: 1712843 [PMC free article: PMC1181775] [PubMed: 1712843]
- 95. Soltesz I, Lightowler S, Leresche N, Jassik-Gerschenfeld D, Pollard CE, Crunelli V. Two inward currents and the transformation of low-frequency oscillations of rat and cat thalamocortical cells. J Physiol. 1991;441 (1):175–197. PMID: 1667794 [PMC free article: PMC1180192] [PubMed: 1667794]
- 96. Nuñez A, Amzica F, Steriade M. Intrinsic and synaptically generated delta (1-4 Hz) rhythms in dorsal lateral geniculate neurons and their modulation by light-induced fast (30-70 Hz) events. Neuroscience. 1992;51 (2):269–284. PMID: 1465192 [PubMed: 1465192]
- 97. Destexhe A, Bal T, Mccormick DA, Sejnowski TJ. Ionic mechanisms underlying synchronized oscillations and propagating waves in a model of ferret thalamic slices. J Neurophysiol. 1996;76 (3):2049–2070. PMID: 8890314 [PubMed: 8890314]
- 98. Steriade M, Dossi RC, Nunez A. Network modulation of a slow intrinsic oscillation of cat thalamocortical neurons implicated in sleep delta waves: Cortically induced synchronization and brainstem cholinergic suppression. J Neurosci. 1991;11 (10):3200–3217. PMID: 1941080 [PMC free article: PMC6575431] [PubMed: 1941080]
- 99. Neske GT. The slow oscillation in cortical and thalamic networks: Mechanisms and functions. Front Neural Circuits. 2016;9(JAN2016):1–25. PMID: 26834569 [PMC free article: PMC4712264] [PubMed: 26834569]
- 100. Lee J, Kim D, Shin HS. Lack of delta waves and sleep disturbances during non-rapid eye movement sleep in mice lacking α1G-subunit of T-type calcium channels. Proc Natl Acad Sci U S A. 2004;101(52):18195–18199. PMID: 15601764 [PMC free article: PMC539778] [PubMed: 15601764]
- 101. Kim D, Song I, Keum S, Lee T, Jeong MJ, Kim SS, McEnery MW, Shin HS. Lack of the burst firing of thalamocortical relay neurons and resistance to absence seizures in mice lacking α1G T-type Ca2+ channels. Neuron. 2001;31 (1):35–45. PMID: 11498049 [PubMed: 11498049]
- 102. Lewis LD, Voigts J, Flores FJ, Schmitt LI, Wilson MA, Halassa MM, Brown EN. Thalamic reticular nucleus induces fast and local modulation of arousal state. Elife. 2015;4(OCTOBER2015):1–23. PMID: 26460547 [PMC free article: PMC4686423] [PubMed: 26460547]
- 103. Fernandez LMJ, Vantomme G, Osorio-Forero A, Cardis R, Béard E, Lüthi A. Thalamic reticular control of local sleep in mouse sensory cortex. Elife. 2018;7:1–25. PMID: 30583750 [PMC free article: PMC6342525] [PubMed: 30583750]
- 104. Notomi T, Shigemoto R. Immunohistochemical Localization of Ih Channel Subunits, HCN1-4, in the Rat Brain. J Comp Neurol. 2004;471 (3):241–276. PMID: 14991560 [PubMed: 14991560]
- 105. Ludwig A, Budde T, Stieber J, Moosmang S, Wahl C, Holthoff K, Langebartels A, Wotjak C, Munsch T, Zong X, Feil S, Feil R, Lancel M, Chien KR, Konnerth A, Pape HC, Biel M, Hofmann F. Absence epilepsy and sinus dysrhythmia in mice lacking the pacemaker channel HCN2. EMBO J. 2003;22 (2):216–224. PMID: 12514127 [PMC free article: PMC140107] [PubMed: 12514127]
- 106. Chung WK, Shin M, Jaramillo TC, Leibel RL, LeDuc CA, Fischer SG, Tzilianos E, Gheith AA, Lewis AS, Chetkovich DM. Absence epilepsy in apathetic, a spontaneous mutant mouse lacking the h channel subunit, HCN2. Neurobiol Dis. Elsevier Inc.; 2009;33 (3):499–508. PMID: 19150498 [PMC free article: PMC2643333] [PubMed: 19150498]
- 107. Heuermann RJ, Jaramillo TC, Ying SW, Suter BA, Lyman KA, Han Y, Lewis AS, Hampton TG, Shepherd GMG, Goldstein PA, Chetkovich DM. Reduction of thalamic and cortical Ih by deletion of TRIP8b produces a mouse model of human absence epilepsy. Neurobiol Dis. Elsevier Inc.; 2016;85:81–92. PMID: 26459112 [PMC free article: PMC4688217] [PubMed: 26459112]
- 108. Ying SW, Jia F, Abbas SY, Hofmann F, Ludwig A, Goldstein PA. Dendritic HCN2 channels constrain glutamate-driven excitability in reticular thalamic neurons. J Neurosci. 2007;27 (32):8719–8732. PMID: 17687049 [PMC free article: PMC6672930] [PubMed: 17687049]
- 109. Kanyshkova T, Meuth P, Bista P, Liu Z, Ehling P, Caputi L, Doengi M, Chetkovich DM, Pape HC, Budde T. Differential regulation of HCN channel isoform expression in thalamic neurons of epileptic and non-epileptic rat strains. Neurobiol Dis. Elsevier Inc.; 2012;45 (1):450–461. PMID: 21945537 [PMC free article: PMC3225716] [PubMed: 21945537]
- 110. Cain SM, Tyson JR, Jones KL, Snutch TP. Thalamocortical neurons display suppressed burst-firing due to an enhanced Ih current in a genetic model of absence epilepsy. Pflugers Arch Eur J Physiol. 2015;467 (6):1367–1382. PMID: 24953239 [PMC free article: PMC4435665] [PubMed: 24953239]
- 111. von Krosigk M, Bal T, Mccormick DA. Cellular Mechanisms of a Synchronized Oscillation in the Thalamus. Science. 1993;261(5119):361–364. PMID: 8392750 [PubMed: 8392750]
- 112. Köhler M, Hirschberg B, Bond CT, Kinzie JM, Marrion N V., Maylie J, Adelman JP. Small-conductance, calcium-activated potassium channels from mammalian brain. Science. 1996;273 (5282):1709–1714. PMID: 8781233 [PubMed: 8781233]
- 113. Wimmer RD, Astori S, Bond CT, Rovó Z, Chatton JY, Adelman JP, Franken P, Lüthi A. Sustaining sleep spindles through enhanced SK2-channel activity consolidates sleep and elevates arousal threshold. J Neurosci. 2012;32 (40):13917–13928. PMID: 23035101 [PMC free article: PMC3658620] [PubMed: 23035101]
- 114. Zaman T, Lee K, Park C, Paydar A, Choi JH, Cheong E, Lee CJ, Shin HS. CaV2.3 Channels Are Critical for Oscillatory Burst Discharges in the Reticular Thalamus and Absence Epilepsy. Neuron. Elsevier Inc.; 2011;70 (1):95–108. PMID: 21482359 [PubMed: 21482359]
- 115. Paz JT, Huguenard JR. R U OK? The novel therapeutic potential of R channels in epilepsy. Epilepsy Curr. 2012;12(2):75–76. PMID: 22473549 [PMC free article: PMC3316366] [PubMed: 22473549]
- 116. Huguenard JR, Prince DA. Intrathalamic rhythmicity studied in vitro: nominal T-current modulation causes robust antioscillatory effects. J Neurosci. 1994a;14 (9):5485–502. PMID: 8083749 [PMC free article: PMC6577071] [PubMed: 8083749]
- 117. Halassa MM, Siegle JH, Ritt JT, Ting JT, Feng G, Moore CI. Selective optical drive of thalamic reticular nucleus generates thalamic bursts and cortical spindles. Nat Neurosci. Nature Publishing Group; 2011;14 (9):1118–1120. PMID: 21785436 [PMC free article: PMC4169194] [PubMed: 21785436]
- 118. Barthó P, Slézia A, Mátyás F, Faradzs-Zade L, Ulbert I, Harris KD, Acsády L. Ongoing network state controls the length of sleep spindles via inhibitory activity [SUPP]. Neuron. 2014;82(6):1367–1379. PMID: 24945776 [PMC free article: PMC4064116] [PubMed: 24945776]
- 119. Leresche N, Lambert RC, Errington AC, Crunelli V. From sleep spindles of natural sleep to spike and wave discharges of typical absence seizures: Is the hypothesis still valid? Pflugers Arch Eur J Physiol. 2012;463 (1):201–212. PMID: 21861061 [PMC free article: PMC3256322] [PubMed: 21861061]
- 120. Kramer MA, Stoyell SM, Chinappen D, Ostrowski LM, Spencer ER, Morgan AK, Emerton BC, Jing J, Westover MB, Eden UT, Stickgold R, Manoach DS, Chu CJ. Focal sleep spindle deficits reveal focal thalamocortical dysfunction and predict cognitive deficits in sleep activated developmental epilepsy. J Neurosci. 2021;41 (8):1816–1829. PMID: 33468567 [PMC free article: PMC8115887] [PubMed: 33468567]
- 121. Holden SS, Grandi FC, Aboubakr O, Higashikubo B, Cho FS, Chang AH, Forero AO, Morningstar AR, Mathur V, Kuhn LJ, Suri P, Sankaranarayanan S, Andrews-Zwilling Y, Tenner AJ, Luthi A, Aronica E, Corces MR, Yednock T, Paz JT. Complement factor C1q mediates sleep spindle loss and epileptic spikes after mild brain injury. Science. 2021;373 (6560):1–9. PMID: 34516796 [PMC free article: PMC8750918] [PubMed: 34516796]
- 122. Hughes SW, Cope DW, Blethyn KL, Crunelli V. Cellular Mechanisms of the Slow (<1 Hz) Oscillation in Thalamocortical Neurons In Vitro. Neuron. 2002;33(6):947–958. PMID: 11906700 [PubMed: 11906700]
- 123. Blethyn KL, Hughes SW, Tóth TI, Cope DW, Crunelli V. Neuronal basis of the slow (<1 Hz) oscillation in neurons of the nucleus reticularis thalami in vitro. J Neurosci. 2006;26 (9):2474–2486. PMID: 16510726 [PMC free article: PMC6793657] [PubMed: 16510726]
- 124. Steriade M, Amzica F. Sleep Oscillations Developing into Seizures in Corticothalamic Systems. Epilepsia. 2003;44(SUPPL.12):9–20. PMID: 14641557 [PubMed: 14641557]
- 125. Scheffer IE, Berkovic S, Capovilla G, Connolly MB, French J, Guilhoto L, Hirsch E, Jain S, Mathern GW, Moshé SL, Nordli DR, Perucca E, Tomson T, Wiebe S, Zhang YH, Zuberi SM. ILAE classification of the epilepsies: Position paper of the ILAE Commission for Classification and Terminology. Epilepsia. 2017;58 (4):512–521. PMID: 28276062 [PMC free article: PMC5386840] [PubMed: 28276062]
- 126. Yang JO, Choi MH, Yoon JY, Lee JJ, Nam SO, Jun SY, Kwon HH, Yun S, Jeon SJ, Byeon I, Halder D, Kong J, Lee B, Lee J, Kang JW, Kim NS. Characteristics of Genetic Variations Associated With Lennox-Gastaut Syndrome in Korean Families. Front Genet. 2021;11(590924):1–11. PMID: 33747056 [PMC free article: PMC7874053] [PubMed: 33584793]
- 127. Dalic LJ, Warren AEL, Young JC, Thevathasan W, Roten A, Bulluss KJ, Archer JS. Cortex leads the thalamic centromedian nucleus in generalized epileptic discharges in Lennox-Gastaut syndrome. Epilepsia. 2020;61(10):2214–2223. PMID: 32944944 [PubMed: 32944944]
- 128. Warren AEL, Dalic LJ, Thevathasan W, Roten A, Bulluss KJ, Archer J. Targeting the centromedian thalamic nucleus for deep brain stimulation. J Neurol Neurosurg Psychiatry. 2020;91(4):339–349. PMID: 31980515 [PubMed: 31980515]
- 129. Velasco M, Velasco F, Alcalá H, Dávila G, Díaz-de-León AE. Epileptiform EEG Activity of the Centromedian Thalamic Nuclei in Children with Intractable Generalized Seizures of the Lennox-Gastaut Syndrome. Epilepsia. 1991;32 (3):310–321. PMID: 1904342 [PubMed: 1904342]
- 130. Velasco AL, Velasco F, Jiménez F, Velasco M, Castro G, Carrillo-Ruiz JD, Fanghänel G, Boleaga B. Neuromodulation of the centromedian thalamic nuclei in the treatment of generalized seizures and the improvement of the quality of life in patients with Lennox-Gastaut syndrome. Epilepsia. 2006;47 (7):1203–1212. PMID: 16886984 [PubMed: 16886984]
- 131. Claes L, Del-Favero J, Ceulemans B, Lagae L, Van Broeckhoven C, De Jonghe P. De novo mutations in the sodium-channel gene SCN1A cause severe myoclonic epilepsy of infancy. Am J Hum Genet. 2001;68 (6):1327–1332. PMID: 11359211 [PMC free article: PMC1226119] [PubMed: 11359211]
- 132. Dravet C. The core Dravet syndrome phenotype. Epilepsia. 2011;52 (SUPPL. 2):3–9. PMID: 21463272 [PubMed: 21463272]
- 133. Andrade DM, Hamani C, Lozano AM, Wennberg RA. Dravet syndrome and deep brain stimulation: Seizure control after 10 years of treatment. Epilepsia. 2010;51 (7):1314–1316. PMID: 19919661 [PubMed: 19919661]
- 134. Maheshwari A, Noebels JL. Monogenic models of absence epilepsy. windows into the complex balance between inhibition and excitation in thalamocortical microcircuits. 1st ed. Progress in Brain Research. Elsevier B.V.; 2014; 213:223–252. PMID: 25194492 [PubMed: 25194492]
- 135. Gobbo D, Scheller A, Kirchhoff F. From Physiology to Pathology of Cortico-Thalamo-Cortical Oscillations: Astroglia as a Target for Further Research. Front Neurol. 2021;12 (June):1–26. PMID: 34177766 [PMC free article: PMC8219957] [PubMed: 34177766]
- 136. Panayiotopoulos CP. Typical absence seizures and related epileptic syndromes: Assessment of current state and directions for future research. Epilepsia. 2008;49 (12):2131–2147. PMID: 19049569 [PubMed: 19049569]
- 137. Jouvenceau A, Eunson LH, Spauschus A, Ramesh V, Zuberi SM, Kullmann DM, Hanna MG. Human epilepsy associated with dysfunction of the brain P/Q-type calcium channel. Lancet. 2001;358 (9284):801–807. PMID: 11564488 [PubMed: 11564488]
- 138. Liang J, Zhang Y, Wang J, Pan H, Wu H, Xu K, Liu X, Jiang Y, Shen Y, Wu X. New variants in the CACNA1H gene identified in childhood absence epilepsy. Neurosci Lett. 2006;406 (1–2):27–32. PMID: 16905256 [PubMed: 16905256]
- 139. Chen Y, Lu J, Pan H, Zhang Y, Wu H, Xu K, Liu X, Jiang Y, Bao X, Yao Z, Ding K, Lo WHY, Qiang B, Chan P, Shen Y, Wu X. Association between genetic variation of CACNA1H and childhood absence epilepsy. Ann Neurol. 2003;54 (2):239–243. PMID: 12891677 [PubMed: 12891677]
- 140. Everett K V., Chioza B, Aicardi J, Aschauer H, Brouwer O, Callenbach P, Covanis A, Dulac O, Eeg-Olofsson O, Feucht M, Friis M, Goutieres F, Guerrini R, Heils A, Kjeldsen M, Lehesjoki AE, Makoff A, Nabbout R, Olsson I, Sander T, Sirén A, McKeigue P, Robinson R, Taske N, Rees M, Gardiner M. Linkage and association analysis of CACNG3 in childhood absence epilepsy. Eur J Hum Genet. 2007;15(4):463–472. PMID: 17264864 [PMC free article: PMC2556708] [PubMed: 17264864]
- 141. Wallace RH, Marini C, Petrou S, Harkin LA, Bowser DN, Panchal RG, Williams DA, Sutherland GR, Mulley JC, Scheffer IE, Berkovic SF. Mutant GABA A receptor γ2-subunit in childhood absence epilepsy and febrile seizures. Nat Genet. 2001;28 (1):49–52. PMID: 11326275 [PubMed: 11326275]
- 142. Baulac S, Huberfeld G, Gourfinkel-An I, Mitropoulou G, Beranger A, Prud’homme JF, Baulac M, Brice A, Bruzzone R, LeGuern E. First genetic evidence of GABAA receptor dysfunction in epilepsy: A mutation in the γ2-subunit gene. Nat Genet. 2001;28 (1):46–48. PMID: 11326274 [PubMed: 11326274]
- 143. Farrello K. Benzodiazepines in the Treatment of Children with Epilepsy. Epilepsia. 1986;27:S45–S52. PMID: 3743524 [PubMed: 3743524]
- 144. Huguenard JR, Prince DA. Clonazepam suppresses GABAB-mediated inhibition in thalamic relay neurons through effects in nucleus reticularis. J Neurophysiol. 1994b;71 (6):2576–2581. PMID: 7931539 [PubMed: 7931539]
- 145. Huguenard JR. Neuronal circuitry of thalamocortical epilepsy and mechanisms of antiabsence drug action. Adv Neurol. 1999;79:991–999. PMID: 10514881 [PubMed: 10514881]
- 146. Vitko I, Chen Y, Arias JM, Shen Y, Wu XR, Perez-Reyes E. Functional characterization and neuronal modeling of the effects of childhood absence epilepsy variants of CACNA1H, a T-type calcium channel. J Neurosci. 2005;25 (19):4844–4855. PMID: 15888660 [PMC free article: PMC6724770] [PubMed: 15888660]
- 147. Klassen T, Davis C, Goldman A, Burgess D, Chen T, Wheeler D, McPherson J, Bourquin T, Lewis L, Villasana D, Morgan M, Muzny D, Gibbs R, Noebels J. Exome sequencing of ion channel genes reveals complex profiles confounding personal risk assessment in epilepsy. Cell. Elsevier Inc.; 2011;145 (7):1036–1048. PMID: 21703448 [PMC free article: PMC3131217] [PubMed: 21703448]
- 148. Noebels J. Precision physiology and rescue of brain ion channel disorders. J Gen Physiol. 2017;149(5):533–546. PMID: 28428202 [PMC free article: PMC5412535] [PubMed: 28428202]
- 149. Tsakiridou E, Bertollini L, De Curtis M, Avanzini G, Pape HC. Selective increase in T-type calcium conductance of reticular thalamic neurons in a rat model of absence epilepsy. J Neurosci. 1995;15 (4):3110–3117. PMID: 7722649 [PMC free article: PMC6577780] [PubMed: 7722649]
- 150. Talley EM, Solórzano G, Depaulis A, Perez-Reyes E, Bayliss DA. Low-voltage-activated calcium channel subunit expression in a genetic model of absence epilepsy in the rat. Mol Brain Res. 2000;75 (1):159–165. PMID: 10648900 [PubMed: 10648900]
- 151. Powell KL, Cain SM, Ng C, Sirdesai S, David LS, Kyi M, Garcia E, Tyson JR, Reid CA, Bahlo M, Foote SJ, Snutch TP, O’Brien TJ. A Cav3.2 T-type calcium channel point mutation has splice-variant-specific effects on function and segregates with seizure expression in a polygenic rat model of absence epilepsy. J Neurosci. 2009;29 (2):371–380. PMID: 19144837 [PMC free article: PMC6664949] [PubMed: 19144837]
- 152. Ernst WL, Zhang Y, Yoo JW, Ernst SJ, Noebels JL. Genetic enhancement of thalamocortical network activity by elevating α1G-mediated low-voltage-activated calcium current induces pure absence epilepsy. J Neurosci. 2009;29 (6):1615–1625. PMID: 19211869 [PMC free article: PMC2660673] [PubMed: 19211869]
- 153. Fletcher CF, Lutz CM, O’Sullivan TN, Shaughnessy JD, Hawkes R, Frankel WN, Copeland NG, Jenkins NA. Absence epilepsy in tottering mutant mice is associated with calcium channel defects. Cell. 1996;87 (4):607–617. PMID: 8929530 [PubMed: 8929530]
- 154. Zhang Y, Mori M, Burgess DL, Noebels JL. Mutations in high-voltage-activated calcium channel genes stimulate low-voltage-activated currents in mouse thalamic relay neurons. J Neurosci. 2002;22(15):6362–6371. PMID: 12151514 [PMC free article: PMC6758149] [PubMed: 12151514]
- 155. Song I, Kim D, Choi S, Sun M, Kim Y, Shin HS. Role of the α1G T-type calcium channel in spontaneous absence seizures in mutant mice. J Neurosci. 2004;24 (22):5249–5257. PMID: 15175395 [PMC free article: PMC6729205] [PubMed: 15175395]
- 156. Polack P, Charpier S. Ethosuximide converts ictogenic neurons initiating absence seizures into normal neurons in a genetic model. Epilepsia. 2009;50 (7):1816–1820. PMID: 19260940 [PubMed: 19260940]
- 157. Lee SE, Lee J, Latchoumane C, Lee B, Oh SJ, Saud ZA, Park C, Sun N, Cheong E, Chen CC, Choi EJ, Lee CJ, Shin HS. Rebound burst firing in the reticular thalamus is not essential for pharmacological absence seizures in mice. Proc Natl Acad Sci U S A. 2014;111 (32):11828–11833. PMID: 25071191 [PMC free article: PMC4136605] [PubMed: 25071191]
- 158. Helbig KL, Lauerer RJ, Bahr J C, Souza IA, Myers CT, Uysal B, Schwarz N, Gandini MA, Huang S, Keren B, Mignot C, Afenjar A, Billette de Villemeur T, Héron D, Nava C, Valence S, Buratti J, Fagerberg CR, Soerensen KP, Kibaek M, Kamsteeg EJ, Koolen DA, Gunning B, Schelhaas HJ, Kruer MC, Fox J, Bakhtiari S, Jarrar R, Padilla-Lopez S, Lindstrom K, Jin SC, Zeng X, Bilguvar K, Papavasileiou A, Xin Q, Zhu C, Boysen K, Vairo F, Lanpher BC, Klee EW, Tillema JM, Payne ET, Cousin MA, Kruisselbrink TM, Wick MJ, Baker J, Haan E, Smith N, Corbett MA, MacLennan AH, Gecz J, Biskup S, Goldmann E, Rodan LH, Kichula E, Segal E, Jackson KE, Asamoah A, Dimmock D, McCarrier J, Botto LD, Filloux F, Tvrdik T, Cascino GD, Klingerman S, Neumann C, Wang R, Jacobsen JC, Nolan MA, Snell RG, Lehnert K, Sadleir LG, Anderlid BM, Kvarnung M, Guerrini R, Friez MJ, Lyons MJ, Leonhard J, Kringlen G, Casas K, El Achkar CM, Smith LA, Rotenberg A, Poduri A, Sanchis-Juan A, Carss KJ, Rankin J, Zeman A, Raymond FL, Blyth M, Kerr B, Ruiz K, Urquhart J, Hughes I, Banka S, Hedrich UBS, Scheffer IE, Helbig I, Zamponi GW, Lerche H, Mefford HC. De Novo Pathogenic Variants in CACNA1E Cause Developmental and Epileptic Encephalopathy with Contractures, Macrocephaly, and Dyskinesias. Am J Hum Genet. 2018;103(5):666–678. PMID: 30343943 [PMC free article: PMC6216110] [PubMed: 30343943]
- 159. Weiergräber M, Henry M, Krieger A, Kamp M, Radhakrishnan K, Hescheler J, Schneider T. Altered seizure susceptibility in mice lacking the Cav2.3 E-type Ca2+ channel. Epilepsia. 2006;47(5):839–850. PMID: 16686648 [PubMed: 16686648]
- 160. Noebels JL. The Voltage-Gated Calcium Channel and Absence Epilepsy. In: Noebels JL, Avoli M, Rogawski MA, Olsen RW, Delgado-Escueta A V, editors. Jasper’s Basic Mech Epilepsies. 4th ed. Bethesda: National Center for Biotechnology Information (US); 2012. p. 702–713. PMID: 22787604 [PubMed: 22787604]
- 161. Burgess DL, Biddlecome GH, McDonough SI, Diaz ME, Zilinski CA, Bean BP, Campbell KP, Noebels JL. β subunit reshuffling modifies N- and P/Q-type Ca2+ channel subunit compositions in lethargic mouse brain. Mol Cell Neurosci. 1999;13 (4):293–311. PMID: 10328888 [PubMed: 10328888]
- 162. Han Z, Chen C, Christiansen A, Ji S, Lin Q, Anumonwo C, Liu C, Leiser SC, Meena, Aznarez I, Liau G, Isom LL. Antisense oligonucleotides increase Scn1a expression and reduce seizures and SUDEP incidence in a mouse model of Dravet syndrome. Sci Transl Med. 2020;12 (558):1–14. PMID: 32848094 [PubMed: 32848094]
- 163. Rubinstein M, Han S, Tai C, Westenbroek RE, Hunker A, Scheuer T, Catterall WA. Dissecting the phenotypes of Dravet syndrome by gene deletion. Brain. 2015;138 (8):2219–2233. PMID: 26017580 [PMC free article: PMC5022661] [PubMed: 26017580]
- 164. Dutton SB, Makinson CD, Papale LA, Shankar A, Balakrishnan B, Nakazawa K, Escayg A. Preferential inactivation of SCN1A in parvalbumin interneurons increases seizure susceptibility. Neurobiol Dis. Elsevier Inc.; 2013;49 (1):211–220. PMID: 22926190 [PMC free article: PMC3740063] [PubMed: 22926190]
- 165. Favero M, Sotuyo NP, Lopez E, Kearney JA, Goldberg EM. A transient developmental window of fast-spiking interneuron dysfunction in a mouse model of dravet syndrome. J Neurosci. 2018;38 (36):7912–7927. PMID: 30104343 [PMC free article: PMC6125809] [PubMed: 30104343]
- 166. Cho FS, Clemente A, Holden S, Paz JT. Thalamic Models of Seizures In Vitro, chapter in Pitkänen, Buckmaster, Galanopoulou, Moshé Model Seizures Epilepsy, Second Ed. Academic Press; 2017. p. 273–284.
- 167. Grossman EJ, Inglese M. The role of thalamic damage in mild traumatic brain injury. J Neurotrauma. 2016;33 (2):163–167. PMID: 26054745 [PMC free article: PMC4722574] [PubMed: 26054745]
- 168. Scott G, Hellyer PJ, Ramlackhansingh AF, Brooks DJ, Matthews PM, Sharp DJ. Thalamic inflammation after brain trauma is associated with thalamo-cortical white matter damage. J Neuroinflammation. Journal of Neuroinflammation; 2015;12 (1):1–5. PMID: 26627199 [PMC free article: PMC4666189] [PubMed: 26627199]
- 169. Pappata S, Levasseur M, Gunn RN, Myers R, Crouzel C, Syrota A, Jones T, Kreutzberg GW, Banati RB. Thalamic microglial activation in ischemic stroke detected in vivo by PET and [(11)C]PK11195. Neurology. 2000;55 (7):1052–1054. PMID: 11061271 [PubMed: 11061271]
- 170. Paz JT, Christian CA, Parada I, Prince DA, Huguenard JR. Focal Cortical Infarcts Alter Intrinsic Excitability and Synaptic Excitation in the Reticular Thalamic Nucleus. J Neurosci. 2010;30 (15):5465–5479. PMID: 20392967 [PMC free article: PMC2861582] [PubMed: 20392967]
- 171. Hazra A, Macolino C, Elliott MB, Chin J. Delayed thalamic astrocytosis and disrupted sleep-wake patterns in a preclinical model of traumatic brain injury. J Neurosci Res. 2014;92 (11):1434–1445. [PubMed: 24964253]
- 172. Necula D, Cho FS, He A, Paz JT. Secondary thalamic neuroinflammation after focal cortical stroke and traumatic injury mirrors corticothalamic functional connectivity. J Comp Neurol. 2021; 530(7):998–1019. PMID: 34633669 [PMC free article: PMC8957545] [PubMed: 34633669]
- 173. Cao Z, Harvey SS, Bliss TM, Cheng MY, Steinberg GK. Inflammatory Responses in the Secondary Thalamic Injury After Cortical Ischemic Stroke. Front Neurol. 2020;11(April):1–12. PMID: 32318016 [PMC free article: PMC7154072] [PubMed: 32318016]
- 174. Immonen R, Kharatishvili I, Gröhn O, Pitkänen A. MRI biomarkers for post-traumatic epileptogenesis. J Neurotrauma. 2013;30 (14):1305–1309. PMID: 23469770 [PMC free article: PMC3713441] [PubMed: 23469770]
- 175. Pitkänen A, Paananen T, Kyyriäinen J, Das Gupta S, Heiskanen M, Vuokila N, Bañuelos-Cabrera I, Lapinlampi N, Kajevu N, Andrade P, Ciszek R, Lara-Valderrábano L, Ekolle Ndode-Ekane X, Puhakka N. Biomarkers for posttraumatic epilepsy. Epilepsy Behav. Elsevier Inc.; 2021;121:107080. PMID: 32317161 [PubMed: 32317161]
- 176. Manninen E, Chary K, Lapinlampi N, Andrade P, Paananen T, Sierra A, Tohka J, Gröhn O, Pitkänen A. Acute thalamic damage as a prognostic biomarker for post- traumatic epileptogenesis. Epilepsia. 2021; 62(8):1852–1864. PMID: 34245005 [PubMed: 34245005]
- 177. Ramlackhansingh AF, Brooks DJ, Greenwood RJ, Bose SK, Turkheimer FE, Kinnunen KM, Gentleman S, Heckemann RA, Gunanayagam K, Gelosa G, Sharp DJ. Inflammation after trauma: Microglial activation and traumatic brain injury. Ann Neurol. 2011;70 (3):374–383. PMID: 21710619 [PubMed: 21710619]
- 178. Grossman E J, Ge Y, Jensen JH, Babb JS, Miles L, Reaume J, Silver JM, Grossman RI, Inglese M. Thalamus and cognitive impairment in mild traumatic brain injury: A diffusional kurtosis imaging study. J Neurotrauma. 2012;29 (13):2318–2327. PMID: 21639753 [PMC free article: PMC3430483] [PubMed: 21639753]
- 179. Fasano A, Eliashiv D, Herman ST, Lundstrom BN, Polnerow D, Henderson JM, Fisher RS. Experience and consensus on stimulation of the anterior nucleus of thalamus for epilepsy. Epilepsia. 2021; 62(12):2883–2898. PMID: 34697794 [PubMed: 34697794]
- 180. Vezzani A, Balosso S, Ravizza T. Neuroinflammatory pathways as treatment targets and biomarkers in epilepsy. Nat Rev Neurol. Springer US; 2019;15 (8):459–472. PMID: 31263255 [PubMed: 31263255]
- 181. Patel DC, Tewari BP, Chaunsali L, Sontheimer H. Neuron–glia interactions in the pathophysiology of epilepsy. Nat Rev Neurosci. Springer US; 2019;20 (5):282–297. [PMC free article: PMC8558781] [PubMed: 30792501]
- 182. Klein P, Dingledine R, Aronica E, Bernard C, Blümcke I, Boison D, Brodie MJ, Brooks-Kayal AR, Engel J, Forcelli PA, Hirsch LJ, Kaminski RM, Klitgaard H, Kobow K, Lowenstein DH, Pearl PL, Pitkänen A, Puhakka N, Rogawski MA, Schmidt D, Sillanpää M, Sloviter RS, Steinhäuser C, Vezzani A, Walker MC, Löscher W. Commonalities in epileptogenic processes from different acute brain insults: Do they translate? Epilepsia. 2018;59 (1):37–66. PMID: 29247482 [PMC free article: PMC5993212] [PubMed: 29247482]
- 183. Aronica E, Bauer S, Bozzi Y, Caleo M, Dingledine R, Gorter JA, Henshall DC, Kaufer D, Koh S, Löscher W, Louboutin JP, Mishto M, Norwood BA, Palma E, Poulter MO, Terrone G, Vezzani A, Kaminski RM. Neuroinflammatory targets and treatments for epilepsy validated in experimental models. Epilepsia. 2017;58:27–38. PMID: 28675563 [PMC free article: PMC5873599] [PubMed: 28675563]
- 184. Winnubst J, Bas E, Ferreira T A, Wu Z, Economo MN, Edson P, Arthur BJ, Bruns C, Rokicki K, Schauder D, Olbris DJ, Murphy SD, Ackerman DG, Arshadi C, Baldwin P, Blake R, Elsayed A, Hasan M, Ramirez D, Dos Santos B, Weldon M, Zafar A, Dudman JT, Gerfen CR, Hantman AW, Korff W, Sternson SM, Spruston N, Svoboda K, Chandrashekar J. Reconstruction of 1,000 Projection Neurons Reveals New Cell Types and Organization of Long-Range Connectivity in the Mouse Brain. Cell. Elsevier Inc.; 2019;179(1):268–281.e13. PMID: 31495573 [PMC free article: PMC6754285] [PubMed: 31495573]
- 185. Siegle JH, Jia X, Durand S, Gale S, Bennett C, Graddis N, Heller G, Ramirez TK, Choi H, Luviano JA, Groblewski PA, Ahmed R, Arkhipov A, Bernard A, Billeh YN, Brown D, Buice MA, Cain N, Caldejon S, Casal L, Cho A, Chvilicek M, Cox TC, Dai K, Denman DJ, de Vries SEJ, Dietzman R, Esposito L, Farrell C, Feng D, Galbraith J, Garrett M, Gelfand EC, Hancock N, Harris JA, Howard R, Hu B, Hytnen R, Iyer R, Jessett E, Johnson K, Kato I, Kiggins J, Lambert S, Lecoq J, Ledochowitsch P, Lee JH, Leon A, Li Y, Liang E, Long F, Mace K, Melchior J, Millman D, Mollenkopf T, Nayan C, Ng L, Ngo K, Nguyen T, Nicovich PR, North K, Ocker GK, Ollerenshaw D, Oliver M, Pachitariu M, Perkins J, Reding M, Reid D, Robertson M, Ronellenfitch K, Seid S, Slaughterbeck C, Stoecklin M, Sullivan D, Sutton B, Swapp J, Thompson C, Turner K, Wakeman W, Whitesell JD, Williams D, Williford A, Young R, Zeng H, Naylor S, Phillips JW, Reid RC, Mihalas S, Olsen SR, Koch C. Survey of spiking in the mouse visual system reveals functional hierarchy. Nature. Springer US; 2021;592 (7852):86–92. PMID: 33473216 [PMC free article: PMC10399640] [PubMed: 33473216]
- 186. Krol A, Wimmer RD, Halassa MM, Feng G. Thalamic Reticular Dysfunction as a Circuit Endophenotype in Neurodevelopmental Disorders. Neuron. Elsevier Inc.; 2018;98 (2):282–295. PMID: 29673480 [PMC free article: PMC6886707] [PubMed: 29673480]
- 187. Cho FS, Vainchtein ID, Voskobiynkyk Y, Morningstar AR, Aparicio F, Ciesielska A, Broekaart DWM, Anink JJ, van Vliet EA, Yu X, Khakh BS, Aronica E, Molofsky AV, Paz JT. Enhancing GAT-3 in thalamic astrocytes promotes resilience to brain injury in rodents. Sci Transl Med. 2022;14 (652):1–14. PMID: 35857628 [PMC free article: PMC9491689] [PubMed: 35857628]
- Abstract
- Introduction
- Thalamic Organization and Rhythmogenesis
- Structural Elements of Thalamic and Thalamocortical Circuits: From Gross Anatomy to Cell Types
- Thalamic Firing
- Calcium Channels and Thalamic Firing
- Rhythmogenesis in the Thalamus: Strengths and Weaknesses
- The Thalamus in Absence Epilepsy and Beyond
- The Emerging Role of the Thalamus in Acquired Epilepsies
- Conclusions
- Acknowledgments
- References
- Convergence of Thalamic Mechanisms in Genetic Epilepsies - Jasper's Basic Mechan...Convergence of Thalamic Mechanisms in Genetic Epilepsies - Jasper's Basic Mechanisms of the Epilepsies
Your browsing activity is empty.
Activity recording is turned off.
See more...