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/9780197549469.003.0004
Abstract
Mutations in genes regulating the mechanistic target of rapamycin (mTOR) signaling pathway have emerged as a common cause of childhood epilepsy, including focal cortical dysplasia, hemimegalencephaly, and tuberous sclerosis complex. Dysregulation of mTOR signaling is also implicated in nongenetic causes of epilepsy, such as following status epilepticus or traumatic brain injury. In this chapter, we describe the evidence linking the mTOR pathway to genetic and acquired epilepsies, the variety of animal models developed to examine mTOR pathway signaling in the brain, and the general patterns of brain pathology produced by pathway disruption. We also examine potential mechanisms by which disrupted mTOR signaling regulates brain development, neuronal excitability, and epileptogenesis. The variety of tissues and cellular functions regulated by mTOR, and the breadth of epilepsies linked to disrupted mTOR signaling, all implicate the pathway as an important modulator of epileptogenesis.
The mTOR Signaling Pathway
The mechanistic target of rapamycin (mTOR) is a serine/threonine-protein kinase that is expressed throughout the body in mammals. It is evolutionarily ancient, being present from yeast to human (Heitman et al., 1991). The mTOR protein itself is a key component of the larger mTOR signaling pathway (Fig. 4–1), which can be activated by both extracellular trophic factors and intracellular factors, such as nutrients. It mediates its effects through multiple levels, from protein-protein interactions to regulation of autophagy to translation. The pathway is a critical regulator of cell metabolism and generally promotes growth (Sabatini, 2017). It is implicated in the regulation of lifespan in mammals, following a pattern in which enhanced growth is associated with reduced longevity (Harrison et al., 2009, Leontieva et al., 2012, Garratt et al., 2016). In the periphery, the pathway also promotes immune function (Jones and Pearce, 2017) and wound healing (Ekici et al., 2007). In the brain, the pathway regulates neurogenesis, neuronal development, neuronal survival, microglial function, and synaptic plasticity (Sarbassov et al., 2005, Switon et al., 2017). While interest in the mTOR pathway in the neurosciences has exploded in recent years, mTOR pathway genes have long been known to be among the most commonly mutated in cancer (Zou et al., 2020). In general, single and double hits to the pathway lead to neurological disease and noninvasive tumors (Martin et al., 2017), whereas multiple hits produce malignant tumors.

Figure 4–1.
The mTOR signaling pathway.
The mTOR pathway acts through two signaling arms mediated by mTOR complex 1 (mTORC1) and mTOR complex 2 (mTORC2). mTORC1 signaling requires the adaptor protein Raptor (Regulatory Associated Protein of mTOR) to function, while mTORC2 requires Rictor (Rapamycin Insensitive Component of mTOR). The two arms are important, as they mediate distinct functions and show differential sensitivity to pharmacological inhibitors—in particular, mTOR’s namesake, rapamycin. Rapamycin originates from a bacterium discovered on Rapa Nui (Easter Island) in the 1970s (Vézina et al., 1975). Rapamycin inhibits mTOR signaling by preventing assembly of mTORC1, although the drug is not a pure mTORC1 antagonist and can also inhibit mTORC2 under certain conditions (Beretta et al., 1996, Sarbassov et al., 2006, Urbanska et al., 2012). Rapamycin (sirolimus) and its analogs (e.g., everolimus) serve both as experimental tools to identify mTOR dependent cellular functions and to treat diseases caused by mTOR pathway hyperactivation.
Clinical Disorders Caused by mTOR Pathway Mutations
mTORopathy Genes
Given the importance and ubiquity of the mTOR pathway for normal cellular function, it is not surprising that an expanding number of clinical diseases have been identified involving mTOR dysfunction, particularly due to genetic mutations in mTOR pathway components (“mTORopathies”). Mutations in at least a dozen genes associated with the mTOR pathway have been found in patients with neurological disorders. These mutations can involve both negative (e.g., TSC1/2, DEPDC5, PTEN) and positive (e.g., MTOR, AKT, PI3K) regulators of the mTOR pathway, but the net result typically leads to mTORC1 hyperactivation due to loss-of-function mutations in negative regulators and gain-of-function mutations in positive regulators. The clinical manifestations of mTORopathies are variable, but many involve pathological features of different malformations of cortical development and neurological symptoms of drug-resistant epilepsy and intellectual disability. Notably, while the current review focuses on the pathological consequences of mTOR hyperactivation, which is strongly linked to epilepsy, other neurological diseases, such as bipolar disorder, can be associated with mTOR hypoactivation (Vanderplow et al., 2021).
Tuberous Sclerosis Complex and Related Malformations of Cortical Development
Tuberous sclerosis complex (TSC), the prototypical mTORopathy, is an autosomal dominant disorder involving tumor formation in multiple organs, including the brain, eyes, skin, kidneys, heart, and lungs (Orlova and Crino, 2010). The neurological manifestations of TSC are typically the most disabling, with a majority of patients having drug-resistant epilepsy (Chu-Shore et al., 2010). Intellectual disability and autism are also common comorbidities. Seizures in TSC are usually thought to originate from cortical tubers, which are focal cortical malformations featuring loss of cortical lamination and the presence of a variety of abnormal cell types, such as dysmorphic neurons and giant cells. Surgical resection of tubers can sometimes eliminate seizures in some TSC patients, but seizures are often difficult to localize to a single tuber and epileptogenesis may also involve tuber-independent cellular and molecular mechanisms (Wong, 2008).
Two genes, TSC1 and TSC2, have been identified that cause TSC, with the TSC1 and TSC2 protein products (also called hamartin and tuberin, respectively) binding together to form a complex, which has the major function of inhibiting the mTOR pathway (Orlova and Crino, 2010). Mutation of either TSC1 or TSC2 results in a disinhibition or hyperactivation of mTORC1, which leads to excessive cell growth and proliferation, promoting tumor formation in various organs in TSC. Resulting mTORC1 hyperactivation likely also promotes cortical tuber formation in early brain development and the subsequent development of epilepsy, though the precise mechanisms of epileptogenesis in TSC are still incompletely understood.
While the link between mTOR and TSC was discovered two decades ago, more recently other related malformations of cortical development strongly associated with intractable epilepsy have been attributed to mTOR pathway mutations. In particular, hemimegalencephaly (HME) and focal cortical dysplasia (FCD) represent two malformations of cortical development that are classified as being associated with abnormal glioneuronal proliferation and share pathological features with tubers of TSC, such as abnormal cortical lamination, dysmorphic neurons, and balloon cells (akin to giant cells of TSC) (Barkovich et al., 2012). In fact, FCD type IIB is often pathologically indistinguishable from tubers of TSC, although isolated FCD lacks the other systemic manifestations and tumors typical of TSC. Mutations in multiple mTOR pathway components (e.g., MTOR, TSC1, TSC2, RHEB, DEPDC5, PI3K, AKT) have been identified in up to 50% of patients with HME or FCD (Lee et al., 2012, D’Gama et al., 2015, Lim et al., 2015, Jansen et al., 2015, Lim et al., 2017).
The genetic mechanisms linking mTOR pathway mutations to malformations of cortical development are not uniform. TSC typically involves a germline heterozygous TSC1 or TSC2 mutation present in all cells of the body, thus predisposing to tumor formation in multiple organs. In addition, an acquired somatic mutation of the other allele (a “second hit”) is thought to be necessary for certain tumors to form in TSC, such as the kidney tumors, although the necessity and prevalence of second hits in cortical tubers are less consistent and often debated (Crino et al., 2010, Qin et al., 2010). In contrast, germline mutations in mTOR pathway genes are relatively rare in FCD and HME, but instead somatic mutations primarily in the brain, but not in other organs, during early cortical development are more common in these other cortical malformations, resulting in somatic mosaicism, where only a subset of cells carry the mutation. In fact, even in the brain itself, only a small percentage (often less than 5%) of brain cells in FCDs may carry an mTOR pathway mutation (Baldassari et al., 2019).
Other Syndromes Associated with mTOR
While a majority of mTOR-related epilepsies involve cortical malformations, some nonlesional focal epilepsies have also been attributed to mTORopathies, such as DEPDC5-related autosomal dominant nocturnal frontal lobe epilepsy or familial focal epilepsy with variable foci (Dibbens et al., 2013; Ishida et al., 2013). Although drug-resistant epilepsy is often the most prominent, disabling symptom of mTORopathies, some mTOR-related syndromes predominantly feature other neurological comorbidities, such as autism in PTEN hamartoma tumor syndrome (Busch et al., 2019). Furthermore, abnormal mTOR signaling has also been implicated in the pathophysiology of nonsyndromic or “idiopathic” autism (Ganesan et al., 2019).
Clinical Treatment of mTORopathies with mTOR Inhibitors
Assuming epilepsy and other clinical manifestations are primarily driven by mTOR pathway hyperactivation, in theory mTOR inhibitors represent a rational treatment for mTORopathies. mTOR inhibitors are proven treatments for brain, kidney, and lung tumors in TSC by inhibiting cell growth and proliferation (Franz et al., 2013). In terms of epilepsy, the efficacy of mTOR inhibitors is best established for everolimus as an adjunctive treatment for TSC patients with refractory focal seizures in placebo-controlled trials (French et al., 2016), leading to the recent FDA approval of everolimus for this indication. Limited case series suggest that mTOR inhibitors could also be effective for other mTORopathies (Parker et al., 2013), but controlled clinical trials are required to establish the role of mTOR inhibitors for non-TSC epilepsies. Furthermore, even in TSC, the effects of mTOR inhibitors on drug-resistant epilepsy are somewhat limited and comparable to other standard antiseizure medications, as most patients do not become seizure-free. Given the known mechanism of action of mTOR inhibitors, which does not involve direct modulation of neuronal excitability, but rather may regulate underlying mechanisms of epileptogenesis, mTOR inhibitors may be more effective as an early antiepileptogenic therapy to prevent epilepsy rather than conventional antiseizure treatment for drug-resistant seizures.
Genetic Models of mTORopathies
An Overview of Models
Understanding mechanisms of epileptogenesis and seizure generation in mTORopathies depends largely on having relevant preclinical and animal models of these disorders. Given the expanding number of mTORopathies and variety of associated genetic, pathological, and epileptogenic mechanisms, there has been a recent explosion in animal models of mTORopathies. These include over 20 brain-specific models of TSC and an increasing number of models targeting other mTOR pathway genes (Table 4–1). While most models of mTORopathies share the central feature of demonstrating increased mTOR pathway activation, no animal model perfectly recapitulates all aspects of human disease. However, each model has specific advantages and limitations in mimicking the various phenotypes of mTORopathies, including genetic mechanisms, pathological features, and seizure semiology. Rather than attempting to describe all relevant models, we provide an overview of the different strategies for generating models of mTORopathies. The corresponding advantages and limitations of different models, as well as the key lessons learned, will also be discussed.

Table 4–1
Summary of Genetic Rodent Models of mTORopathies.
Constitutive Knockout Models
Given the role of mTOR in critical physiological and developmental functions, homozygous constitutive knockout of most mTOR pathway genes in mice is embryonic lethal, thus limiting their usefulness as models of mTORopathies. Heterozygous constitutive knockouts are viable and represent realistic models on the genetic level for disorders involving germline mutations, particularly TSC. However, the pathological and behavioral phenotypes of heterozygous TSC models, such as the Eker rat and Tsc1+/– or Tsc2+/– mice, are relatively limited. The Eker rat with a spontaneous Tsc2 mutation has been reported to have lesions resembling cortical tubers and subependymal hamartomas, but these are very rare (one rat with a tuber, two rats with subependymal hamartomas, out of 19 Eker rats) (Mizuguchi et al., 2000). In terms of an epilepsy phenotype, there are no known reports of the Eker rat having spontaneous seizures, though they do have a modestly decreased threshold to chemically induced seizure kindling (Waltereit et al., 2006). Similarly, adult Tsc1+/– and Tsc2+/– mice have minimal to no pathological abnormalities or spontaneous seizures, although they do have learning deficits (Goorden et al., 2007, Ehninger et al., 2008), and pre-weanling Tsc1+/– mice have been reported to have seizures in a limited developmental period that apparently resolve with age (Gataullina et al., 2016). The relatively mild neurological phenotypes of heterozygous models of TSC may reflect the absence of “second hits,” which may be required to induce more robust pathological manifestation and seizures, at least within animal models.
Conditional Knockout Models
Conditional knockout models target gene deletion selectively to subsets of cells (e.g., neurons, astrocytes), thus limiting the effects of the genetic manipulation and potentially testing the role of that gene in the targeted cell type. In addition, the timing of the gene inactivation can be varied by using promoters expressed at specific developmental time points or by utilizing inducible gene inactivation systems. There have been numerous conditional knockout mouse models of mTORopathies, most frequently involving Tsc1 or Tsc2 gene inactivation, but also other mTOR pathway genes, particularly negative regulators of mTOR, such as Pten and Depdc5 (Table 4–1). The phenotypic features depend on the spatial and temporal properties and targeted cells of the gene inactivation, but most models exhibit evidence of abnormal mTOR pathway activation, similar pathological features, and epilepsy, regardless of which mTOR pathway gene is inactivated. Conditional knockout mice targeting neurons, glia, or neuroglial progenitor cells typically exhibit cytomegaly, astrogliosis, and megalencephaly, which tend to be diffuse, without focal tuber-like abnormalities (Meikle et al., 2007, Uhlmann et al., 2004, Goto et al., 2011, Way et al., 2009, Yuskaitis et al., 2018, Backman et al., 2001). Most of these conditional knockout mice also exhibit epilepsy, which is often progressive, leading to premature death. mTOR inhibitors are uniformly effective against seizures in these models, including prevention of epilepsy when started early, which along with prevention of the underlying pathological abnormalities, is most likely consistent with an antiepileptogenic rather than simply seizure-suppressing effect (Meikle et al., 2008a, Yuskaitis et al., 2019, Ljungberg et al., 2009, Zeng et al., 2008).
The similarities in the phenotype of these conditional knockout models regardless of the mTOR regulator involved confirm the clinical overlap between the mTORopathies and suggest that mTOR activation is the primary pathophysiological mechanism causing the phenotype. Furthermore, as only homozygous, not heterozygous, conditional knockout mice develop severe neurological manifestations, these models support the importance of a “two-hit” mechanism for brain manifestations of mTORopathies, at least those involving negative regulators of the mTOR pathway, such as Tsc1/Tsc2, Pten, and Depdc5. A major limitation of the CKO knockout strategy in modeling human disease is the diffuse nature of the pathological abnormalities in animals and the absence of focal lesions resembling tubers or FCD, which is better addressed by viral or in utero electroporation approaches.
Viral Deletion and In Utero Electroporation Strategies
These models provide a means to recapitulate focal lesions more precisely in TSC and FCDII. One such TSC mouse model combined exogenous injection of a Cre-expressing viral vector in Tsc1floxed/floxed pups. By controlling the dose of the viral vector, investigators achieved mosaic Tsc1 loss in a localized region similar to those observed in human TSC (Prabhakar et al., 2013). Another approach that has emerged within the last decade is based on in utero electroporation (IUE). This technique allows for in vivo manipulation of specific cell types in select brain regions (LoTurco et al., 2009). Using this technique, plasmid DNA targeting PI3K-mTOR or GATOR1 components was microinjected into the ventricles of embryonic rodent brains and transfected via electrical pulses into neural progenitor cells (i.e., radial glia) that line the ventricles. The position of the electrodes that generate the electrical pulses, which is manually determined by the investigator, directs plasmid expression into the brain region of interest. Current IUE-based models of mTORopathies include expression of Pi3k, Akt, Rheb, and mTOR gain-of- function variants (mTORC1 activators) or suppression of Pten, Tsc1, Tsc2, Stradα, Depdc5, and Nprl3 (mTORC1 negative regulators) (Table 4–2). Some of the gain-of-function variants are mutations that have been identified in patients with malformations of cortical development (MCD) and epilepsy while others are experimental mutations predicted to be pathogenic. Suppression of the negative regulators is achieved by CRISPR/Cas9 gene editing, which leads to gene knockout, or short hairpin RNA (shRNA)-mediated methods, which leads to gene knockdown. Additionally, two studies used a two-hit model where a Cre plasmid was expressed by IUE in Tsc1floxed/mutant or Depdc5floxed/mutant mice. The studies all targeted cortical pyramidal neurons of layer (L) 2/3 as these principal cells are proposed to be most severely affected in MCDs (Blumcke et al., 2009; Hadjivassiliou et al., 2010; D’Gama et al., 2017).

Table 4–2
Summary of IUE-Based Rodent Models of mTORopathies.
It is important to note that although IUE targets radial glia that give rise to both neurons and astrocytes, episomal plasmids that are used in standard IUE will only be expressed in neurons due to a plasmid dilution effect that occurs in dividing cells (Chen and LoTurco, 2012, Feliciano et al., 2011). In contrast, CRISPR/Cas9 plasmids integrate into the genome of the electroporated radial glia and are expected to label all cell types in the radial glial cell lineage (i.e., neurons and astrocytes). IUE has allowed for quantification of the cytoarchitectural abnormalities, including cell misplacement leading to mislamination, increased neuronal soma size, and dendritic arborization. In addition, multinucleated cells have also been reported (Feliciano et al., 2011). In a few studies, IUE has been followed by video-EEG recordings in >2-month-old mice. Mice displayed Racine grade 4/5 convulsive seizures that occurred almost daily at a rate of 3–6 per day (as a mean) depending on the degree of mTORC1 activation. Overall, IUE offers the flexibility to vary the region, timing, and cell-type specificity of plasmid expression in a relatively simple manner, facilitating analyses of their impact on cellular alterations and seizure activity.
Patient-Derived Induced Pluripotent Stem Cell and Organoid Models
Given some of the limitations of rodent models in recapitulating human disease, patient-derived cellular models offer some advantages for mechanistic and preclinical studies, including preservation of patient-specific genetic backgrounds and mechanisms. Induced pluripotent stem cell (iPSC)-derived neurons in culture or three-dimensional organoids have been generated from patients with TSC, STRADA, or DEPDC5 mutations (Dooves et al., 2021, Nadadhur et al., 2019, Winden et al., 2019, Klofas et al., 2020, Dang et al., 2021). While there are limitations in the ability of these reduced systems to generate seizures, they do show evidence of hyperexcitable neuronal network activity, as well as pathological features of cytomegaly (Nadadhur et al., 2019, Winden et al., 2019, Dooves et al., 2021). Further studies may better define mechanisms related to epileptogenesis in these patient-specific models.
General Lessons from Animal Models
As a disease class, mTORopathies are diverse, involving numerous genes with variable involvement of different cell types and tissues. Studies from animal models, however, have produced several broad findings predictive of disease severity, including the size of the cell population carrying the causal mutation, the developmental stage when mutations first occur, and the types of affected cells.
Size of the Affected Cell Population
An important feature of mTORopathies is that they can be driven both by germline and somatic mutations. Somatic mutations in brain cell progenitors likely underlie some of the diversity in phenotypes seen in mTORopathies. A mutation occurring in an early-stage progenitor, for example, can impact an entire hemisphere, likely underling severe conditions such as hemimegalencephaly. Mutations in late-stage progenitors, on the other hand, are hypothesized to underlie smaller lesions, such as those observed in FCD. Consistent with this interpretation, FCD tends to be sporadic, and brain mosaicism rate is positively correlated with pathology in patients with mTORopathies (Marsan and Baulac, 2018; Lee et al., 2021). An interesting implication of this spectrum is that it predicts even more subtle disruptions are possible if somatic mutations occur very late in development or even among postnatally/adult-generated neuronal populations. Consistent with this prediction, postnatal deletion of the mTOR inhibitor PTEN from hippocampal granule cells in mice causes epilepsy and behavioral deficits (Amiri et al., 2012; Pun et al., 2012). Indeed, loss of Pten from only 5% of granule cells rendered the hippocampus hyperexcitable. Moreover, experimentally manipulating the number of knockout cells regulates disease severity, with greater cell numbers producing more severe pathology (LaSarge et al., 2021). The cellular threshold for mTOR mutations to cause disease in humans is not known; however, it is conceivable that low-level mosaicism could underlie some epilepsies of unknown cause or other cognitive deficits (Mirzaa et al., 2016).
Earlier Onset of mTOR Hyperactivation Produces More Severe Pathology
Somatic disease-causing mTOR pathway mutations in neurons would seem to invariably involve progenitor cells, as a somatic mutation in a single postmitotic neuron seems unlikely to produce circuit level abnormalities (excluding mutations that lead to tumor formation). Daughter cells of affected progenitors will experience hyperactive mTOR signaling throughout their maturation. Because the mTOR pathway is active throughout neuronal development, early disruption accounts for many of the pathologies observed among mTORopathies, including deficits in neuronal proliferation, neuronal migration, axonal growth, and dendritic development (Kwon et al., 2006; Magri et al., 2011; Lafourcade et al., 2013; LaSarge et al., 2015; Santos et al., 2017). The pathology does suggest that the types of progenitors affected introduce some variability. For example, mTORopathies can present with both cortical and subcortical dysplasia, or with only cortical dysplasia, perhaps reflecting disruption of ventricular progenitors in the former, or secondary progenitors that have already successfully migrated into cortex in the latter.
Even though clinical mTORopathies target neural progenitors, ascertaining how hyperactive mTOR signaling disrupts different neuronal stages is important for understanding mTOR function and predicting therapeutic targets. The mTOR pathway continues to play important roles in mature neurons, regulating neuronal metabolism, autophagy, neuronal physiology, and plasticity (Tavazoie et al., 2005; Bateup et al., 2011; Takeuchi et al., 2013; Skelton et al., 2019; Skelton et al., 2020). While changes occurring early in development, such as migration defects, might be irreversible, other changes, such as mTOR regulated plasticity, might respond to treatment. Numerous investigators have made use of conditional gene deletion approaches to address this question by disrupting mTOR at different developmental stages. Conditional deletion of mTOR regulators from progenitor cells in rodent models using promoters such as Nestin, Emx and Gli1 invariably produce a full range of pathologies (Table 4–1). Targeted deletions using promoters for postmitotic neurons (e.g., CamkII, synapsin) produce phenotypes that range from subtle to severe, likely reflecting combined effects of later gene loss and lesion size (Table 4–1; Kwon et al., 2006; Meikle et al., 2007; Luikart et al., 2011; Abs et al., 2013; Bateup et al., 2013; Koene et al., 2019). The findings indicate that mTORopathies reflect a combination of disrupted developmental processes and mature neuronal functions. Consistent with this interpretation, delayed rapamycin treatment of rodents with mTOR pathway mutations can improve behavioral phenotypes and reverse neuronal hypertrophy (Ehninger et al., 2008; Getz et al., 2016). Migration defects, on the other hand, were irreversible (Getz et al., 2016). The extent to which developmentally induced structural abnormalities drive disease remains an area of intense research. Notably, seizures can occur in the absence of migration defects, suggesting that changes in cell physiology may be more important than some of the gross structural defects (Hsieh et al., 2016).
Target Cell Type Matters
Germline mTOR pathway mutations have the potential to impact every cell in the body, while somatic mutations can impact homogenous or mixed cell types, depending on the progenitor affected. In clinical samples from patients with TSC, pathology and mTOR pathway hyperactivation support involvement of excitatory neurons, inhibitory neurons, astrocytes, and oligodendrocytes (Wong, 2019; Gelot and Represa, 2020; Zimmer et al., 2020). The types of cells involved presumably reflect the nature of the progenitor cell affected in the case of somatic mutations, and likely secondary disruption of wildtype cells. Animal studies, however, suggest that not all cell types contribute equally to pathology.
mTOR Hyperactivation in Excitatory versus Inhibitory Neurons
Deletion of mTOR pathway genes from excitatory neurons produces morphological abnormalities, increased excitability, behavioral deficits, and seizures in rodent models. Although models vary, severe disease is produced following gene deletion from a variety of excitatory neurons (Table 4–1). By contrast, mTOR hyperactivation among interneurons does not produce gross brain hypertrophy or seizures (Gregorian et al., 2009; D’Gama et al., 2017; Zhao and Yoshii, 2019). Reductions in interneuron number and changes in interneuron phenotype have been observed following mTOR pathway gene deletion from interneuron progenitors (Vogt et al., 2015; D’Gama et al., 2017; Ka et al., 2017; Malik et al., 2019). Such interneuron deficits could lead to subtle changes in behavior or seizure susceptibility. Indeed, Fu and colleagues (Fu et al., 2012) observed a paradoxical decreased latency to flurothyl-evoked myoclonic jerks, but increased latency to clonic seizures following Tsc1 deletion from medial ganglionic eminence-derived inhibitory neuron progenitors. Moreover, deletion of the mTOR inhibitor Pten from interneurons increased inhibitory tone (Vogt et al., 2015), raising the possibility that mTOR hyperactivation in interneurons may oppose pro-excitatory effects following hyperactivation among excitatory neurons. In children with TSC, recent studies suggest that interneurons can harbor TSC mutations, evident as enhanced pS6 immunoreactivity in autopsy specimens (Gelot and Represa, 2020), but their significance is unclear. Favoring enhanced function, GABAergic signaling was found to be increased in iPSCs from a patient with a TSC2 mutation (Alsaqati et al., 2020). In general, while mTOR mutations appear to follow a pattern of increasing excitability among excitatory neurons (Santos et al., 2017; Skelton et al., 2019), co-involvement of interneurons—and potential enhancement of interneuron function—adds additional complexity to predicting the net effects of mTORopathies on brain excitability.
mTOR Hyperactivation among Non-Neuronal Cells in the CNS
Animal models demonstrate that mTOR pathway mutations targeted to non-neuronal cells can produce severe brain malformations and disease. Mutations targeted to astrocytes consistently produce severe disease in animal models (Table 4–1), and histopathological studies implicate astrocytes and oligodendrocytes in TSC (Wong, 2019, Zimmer et al., 2020). mTOR hyperactivation in microglia is also sufficient to cause epilepsy (Zhao et al., 2018), while inhibition makes acquired epilepsy worse (Zhao et al., 2020), indicative of a complex biological role for mTOR in microglia. In addition, the brain vasculature appears to be an important target. Tsc and Pten mutations drive angiogenesis and vascular abnormalities (Wen et al., 2001; Martinez-Lopez et al., 2019), and animal models of TSC replicate these findings (Zhang et al., 2019; Kútna et al., 2020). Abnormalities could develop by direct disruption of components of the vasculature, such as endothelial cells, or by altered neurovascular signaling (Parker et al., 2011; Feliciano et al., 2013; Brugarolas et al., 2003). Vascular abnormalities can contribute to blood-brain leakage, which is important, because leakage can cause seizures (Gorter et al., 2019), while studies with rapamycin implicate the mTOR pathway in blood–brain barrier disruption following stimulation-induced status epilepticus (van Vliet et al., 2012; van Vliet et al., 2016b; van Vliet et al., 2016a).
Finally, while the epilepsy field naturally focuses on the brain, we would be remiss to not discuss potential mechanisms by which mTOR could modulate epilepsy from outside the CNS. The mTOR signaling pathway is a critical regulator of systemic immune function, and the role for mTOR in energy homeostasis may also be relevant. mTOR signaling is important for regulating adipose tissue (Cai et al., 2016) and controls ketogenesis (Sengupta et al., 2010; McDaniel et al., 2011). Germline mutations in mTOR pathway genes and systemic treatment with rapalogues, therefore, have the potential to impact peripheral systems that could modulate neurological phenotypes.
Molecular Mechanisms of Epileptogenesis due to mTOR Hyperactivation
Given the diversity of physiological functions controlled by mTOR, abnormal mTOR pathway hyperactivation in mTORopathies has the potential to promote epileptogenesis by a variety of cellular and molecular mechanisms. First, mTOR activation can cause cytomegaly and other pathological changes that lead to formation of focal cortical malformations (e.g., tubers, FCD, HME) during early brain development, and the lesion itself can secondarily lead to seizures, potentially independent of mTOR activation. Secondly, mTOR signaling may stimulate other specific molecular mechanisms that directly promote epileptogenesis, either within a cortical malformation or potentially completely independent of the lesion. In addition, the molecular mechanisms involved may vary depending on the specific gene and mTOR pathway component that is mutated. In this section, molecular mechanisms regulated by mTOR that may directly mediate epileptogenesis are reviewed.
Ion Channel/Neurotransmitter Regulation
As a major function of the mTOR pathway, particularly mTORC1, is to regulate protein synthesis, mTORC1 hyperactivation in mTORopathies has the potential to cause dysregulated expression of numerous proteins. Of direct relevance to neuronal excitability and epileptogenesis, several ion channel subunits and neurotransmitter receptors or transporters have been found to have abnormal (increased or decreased) expression in brain samples from patients or animal models of mTORopathies. In animal models, most of these changes have been shown to be reversible with rapamycin, establishing their mTOR dependence. For example, ectopic expression of the nonspecific cation HCN4 channel is increased in cortical neurons in a Rheb-activating in utero electroporation model and in brain samples from TSC and FCDII patients, which is mTOR dependent and promotes neuronal depolarization and seizures in the mouse model (Hsieh et al., 2020). Conversely, decreased expression of specific potassium channel subtypes, either in neurons or astrocytes, has been found to be due to mTOR activation in TSC or other models, which could also contribute to increased neuronal excitability (Jansen et al., 2005; Raab-Graham et al., 2006). Abnormal expression of glutamate and GABA receptor channel subunits has also been identified in clinical samples from TSC and FCDII patients (Talos et al., 2012a; Talos et al., 2008), though their causal role in epilepsy is difficult to prove in human studies. In a mouse model of TSC, a specific NMDA glutamate receptor subunit is upregulated in cortical neurons in an mTOR dependent manner, and a specific NMDA receptor antagonist blocks seizures in these mice (Lozovaya et al., 2014). In another animal model of TSC focusing on the role of astrocytes in epileptogenesis, decreased astrocyte glutamate transporters lead to elevated extracellular glutamate levels and seizures, which can be reversed by rapamycin and other modulators of glutamate transporters (Wong et al., 2003; Zeng et al., 2010). Overall, there appears to be a spectrum of neuronal and glial ion channels and neurotransmitter receptors/transporters whose protein expression is dysregulated by mTOR hyperactivation and which collectively may contribute to epileptogenesis in mTORopathies.
Inflammatory Mechanisms
Neuroinflammation and innate immunity in the brain have received increasing attention as potential mechanisms contributing to epilepsy (Vezzani, 2020). With mTORopathies, inflammatory and immune processes also represent a rational mechanism of epileptogenesis, given the normal role of mTOR signaling in regulating the immune system and the established action of mTOR inhibitors as immunosuppressant drugs (Nedredal et al., 2020). A number of inflammatory molecules have been identified in human brain pathological specimens or animal models of TSC, including a variety of cytokines, chemokines, and complement (Boer et al., 2008, Boer et al., 2010, Zhang et al., 2015). The cellular localization and contribution of these inflammatory cascades may vary, but predominantly involve non-neuronal populations, such as astrocytes, microglia, and systemic immune cells. In a mouse model of TSC, elevated proinflammatory cytokines and chemokines are inhibited by rapamycin, again supporting mTOR dependence. Furthermore, an anti-inflammatory drug that suppresses cytokine and chemokine production partially decreases seizures in these mice (Zhang et al., 2015). There is also a strong possibility that mTOR antagonists impact epilepsy by regulating systemic immune function in tissues outside the brain. Notably, mTOR acts as a metabolic sensor and regulates the function of cells in the innate and adaptive immune systems (Jones and Pearce, 2017). While the specific mechanisms of action of mTOR inhibitors in decreasing seizures are unknown, similar to other immune-modulatory drugs like corticosteroids that can be effective for epilepsy, the efficacy of mTOR inhibitors in epilepsy may, in part, involve their immunosuppressant properties.
Metabolism and Autophagy
The mTOR pathway normally plays an essential role in regulating nutrient processing, growth, and metabolism in response to environmental conditions. Under anabolic states of nutrient surplus, the mTOR pathway is activated by upstream regulators, such as the insulin/PI3K/AKT growth factor pathway, to stimulate cellular growth and metabolism. Conversely, under catabolic conditions of relative nutrient or oxygen deprivation, the mTOR pathway is inhibited by upstream regulators, such as the GATOR complex or AMPK, to conserve energy and inhibit metabolic processes. Under pathological states of mTOR hyperactivation with mTORopathies, dysregulated metabolic processes could contribute to neurological dysfunction, such as epilepsy.
mTOR pathway hyperactivation in mTORopathies promotes anabolic metabolism, which is highly dependent on mitochondria. In Tsc2-deficient cellular models, mitochondrial function and bioenergetics have been found to be impaired, with decreased respiration and ATP turnover (Ebrahimi-Fakhari et al., 2016). In addition, the turnover of mitochondria through the process of mitophagy is also decreased, leading to accumulation of dysfunctional mitochondria. While the contribution of impaired mitochondrial homeostasis and function to epileptogenesis in mTORopathies has not been tested directly, it is rational to hypothesize that disturbances in brain energetics would affect epilepsy given the metabolic dependence of seizures. In fact, while the ketogenic diet, an effective dietary therapy for epilepsy, has an unknown mechanism of action, the ketogenic diet clearly modulates the metabolic and energetic status of the brain and promotes a catabolic state, which may relate to inhibition of the mTOR pathway (McDaniel et al., 2011).
A related process affecting metabolism is autophagy, which normally serves to degrade damaged organelles within lysosomes and recycle cellular components for energy conservation. Without autophagy, intracellular debris could accumulate leading to the release of harmful reactive oxidative species or other toxic effects on cells that could promote epileptogenesis. mTORC1 is a potent inhibitor of autophagy and mTOR inhibitors stimulate autophagy in most cells. However, in neuronal models of mTORopathies, loss of Tsc1 or Tsc2 paradoxically leads to accumulation of autophagic markers and increased autophagy via mTORC1-regulation of a ULK kinase (Di Nardo et al., 2014). Thus, the relationship of autophagy to epilepsy is complex and not well-established, potentially being a double-edged sword. Nevertheless, conditional knockout of Pten or Tsc1 in mouse cortical neurons in vivo demonstrates reduced autophagy, which is reversed by rapamycin and is at least correlated with seizure development (McMahon et al., 2012). While a direct causal link between mTOR, autophagy, and epilepsy was not established in those mouse models, in separate experiments knockout of Atg7, a molecular promotor of autophagy, caused impaired autophagy and induced epilepsy in mice, indicating that inhibition of autophagy is sufficient to cause epilepsy (McMahon et al., 2012). Thus, metabolic and energetic processes in the brain are likely important factors that at least modulate, if not directly promote, epileptogenesis in mTORopathies.
Morphological and Pathological Effects of mTOR Hyperactivation on Cell Structure
Ectopic neuronal placement and the presence of dysmorphic, cytomegalic neurons are major histopathological findings in TSC and FCDII (Blumcke et al., 2011; Wong, 2008). Animal models based on IUE provide a precise method to quantify the cellular alterations that are outlined below. All gene variants (see Table 4–2) consistently reproduce these phenotypes (Fig. 4–2), supporting the notion that the underlying mechanisms for these alterations involve the mTOR pathway and its downstream processes. Additional histopathological phenotypes, such as the presence of balloon cells, have not been commonly reported in murine studies. Gliosis, including astrocyte and microglia reactivity, is regularly observed but seems to track with seizure activity (Feliciano et al., 2011, Nguyen et al., 2019) and is not discussed here.

Figure 4–2.
Biocytin-filled hippocampal dentate granule cells from a control and a Pten knockout (KO) mouse. The KO granule cell shows pronounced hypertrophy. Scale = 100 µm.
Cortical pyramidal neurons destined to L2/3 are generated from radial glia in the ventricular zone at E14-15 in mice. Newborn neurons migrate from the ventricular zone shortly thereafter and reach their final position in the cortex by postnatal day (P) 7 (Greig et al., 2013; Kast and Levitt, 2019). In line with this timeframe, IUE at E13.5–15.5 results in neuronal migration defects and misplacement for all evaluated gene variants (Table 4–2). Despite being misplaced in other cortical layers, these neurons preserved the molecular identity of L2/3 pyramidal neurons, as evidenced by positive staining for the upper layer markers Cux1 or Satb2 and negative staining for the deeper layer marker Ctip2 (Baek et al., 2015; Iffland et al., 2020; Onori et al., 2020; Park et al., 2018; Tarkowski et al., 2019; Tsai et al., 2014; Zhong et al., 2021; Moon et al., 2015; Lin et al., 2016). Prenatal treatment with rapamycin to reduce mTORC1 activity prevented the migration defects and led to correct placement of the neurons in the cortex, supporting that mTORC1 hyperactivity contributes to these defects (Baek et al., 2015; Tsai et al., 2014; Onori et al., 2020; Kim et al., 2019; Parker et al., 2013; Ribierre et al., 2018; Iffland et al., 2020; Kassai et al., 2014b). Postnatal rapamycin treatment after P7, when cortical neuronal migration is complete, is not expected to rescue neuronal misplacement. Indeed, this expectation was confirmed in the Pi3k E545 variant (Zhong et al., 2021). Additionally, two studies used a conditional Rheb mutant plasmid with a tamoxifen-inducible Cre plasmid to show that mTORC1 hyperactivation has no effect on neuronal placement once migration is complete (Hsieh et al., 2016; Onori et al., 2020). Thus, the impact of mTORC1 hyperactivation on neuronal placement occurs during a specific developmental timeframe, and cortical mislamination can be prevented in a limited temporal window.
Neuronal cytomegaly is consistently reported for the evaluated gene variants (Table 4–2). Furthermore, increasing mTORC1 activity levels with increasing concentrations of the activating Rheb S16H plasmid resulted in dose-dependent neuronal hypertrophy, supporting a direct relationship between mTORC1 activity level and soma size (Nguyen et al., 2019). In contrast to migration defects, mTORC1-induced neuronal dysmorphogenesis is not restricted to an early developmental window, as hyperactivating mTORC1 signaling via conditional Rheb S16H expression at P7 resulted in enlarged neurons (Hsieh et al., 2016). Neuron soma size was reduced with both pre- and postnatal rapamycin (or everolimus) treatment (Baek et al., 2015; Lim et al., 2015; Tsai et al., 2014; Zhong et al., 2021; Hsieh et al., 2016; Lim et al., 2017; Zhang et al., 2019; Park et al., 2018; Hu et al., 2018; Kassai et al., 2014b; Iffland et al., 2020). Moreover, postnatal rapamycin treatment reduced soma size regardless of whether treatment began in neonatal or adult ages, suggesting mTORC1 hyperactivation impacts neuron size via a dynamic, modifiable process. Dendrite morphology was assessed in the IUE studies targeting Pi3k, Pten, Rheb, and Depdc5, and all revealed dendritic overgrowth involving (for some) increased dendrite thickness and complexity (Zhong et al., 2021; Hsieh et al., 2016; Lin et al., 2016; Chen et al., 2015; Sokolov et al., 2018; Zhang et al., 2019; Zhang et al., 2020; Onori et al., 2020; Ribierre et al., 2018; Dawson et al., 2020). Rescue of dendrite morphology by postnatal rapamycin treatment was reported in the Pi3k E545 and Rheb S16H variants, and rapamycin likely has the same effect for the other variants (Zhong et al., 2021; Zhang et al., 2019). Data on axons in the IUE models are more limited. One study reported accelerated axon growth in L2/3 neurons at P0 following Rheb S16H expression (Gong et al., 2015). Axon overgrowth to the contralateral cortex of L2/3 neurons at P45 has also been reported in mice expressing the Rheb P37L variant (Onori et al., 2020).
The effectiveness of rapamycin in restoring neuronal migration and morphology defects supports that these phenotypes are mTORC1-dependent processes. mTORC1 regulates several downstream intracellular pathways, in particular altered translational control via 4E-BP1/2 and S6K1/2, and defective neuronal ciliogenesis in the context of autophagy, that contribute to some of the cellular alterations. mTORC1 controls mRNA translation via direct phosphorylation of two major translational regulators, S6K1/2 and 4E-BP1/2 (Ma and Blenis, 2009). 4E-BP1/2 serves as a translational repressor that inhibits eIF4F function. Expressing a mutant 4E-BP1/2 which resists inactivation by mTORC1 or knockdown of eIF4E prevented mutant Rheb and mTOR-induced neuronal misplacement (Lin et al., 2016; Kim et al., 2019). Furthermore, knockdown of 4E-BP2 alone, which leads to increased cap-dependent translation, led to ectopic neuronal placement, suggesting that increased cap-dependent translation is necessary and sufficient to induce neuronal misplacement (Lin et al., 2016). With regards to morphology, expression of constitutively active 4E-BP1 or knockdown of eIF4E reduced soma enlargement caused by Rheb or mTOR mutants. Additionally, knockdown of S6K1/2 reduced the increased soma size associated with a hyperactivating rat mTOR mutant (Kassai et al., 2014b). In terms of axons, studies by Gong et al. showed that enhancing 4E-BP1 or reducing S6K1/2 activity prevented Rheb mutant-induced axon overgrowth in vivo (Gong et al., 2015), supporting in vitro studies demonstrating both 4E-BP1/2 and S6K1/2 involvement in axon development (Choi et al., 2008; Li et al., 2008; Morita and Sobue, 2009). With respect to another major mTORC1-dependent pathway, it was reported that neuronal autophagy-mediated primary cilia defects cause mTORC1-induced neuronal misplacement (Park et al., 2018). More specifically, impaired autophagy following mTOR mutant expression led to OFD1 accumulation, a regulator of neuronal ciliogenesis, and disrupted cilia formation. Neuronal cilia are essential organelles that integrate many important signaling pathways, such as Wnt, and defective neuron cilia-mediated signaling can lead to abnormal migration and ectopic neuronal placement (Guemez-Gamboa et al., 2014). Knockdown of Ofd1 or expression of Wnt5a restored neuronal placement in the mTOR mutant condition, supporting that disrupted ciliogenesis underlies these alterations. Importantly, defective ciliogenesis was sensitive to rapamycin treatment. Overall, these studies demonstrate that defects in neuronal migration and morphology are conserved phenotypes across distinct PI3K-mTOR and GATOR1 gene variants, and the mechanisms that underlie these alterations likely converge on mTORC1 and its downstream effectors.
Genetic Effects: Not All mTORopathy Genes Are Equal
While pathological variants in the PI3K-mTOR pathway and GATOR1 complex all lead to mTORC1 hyperactivity, they can also differentially modulate other intracellular signaling pathways independently of mTORC1. These parallel pathways include (but are not limited to) AKT-GSK3β, AKT- FOXG1-reelin, and TSC/RHEB-MAPK/ERK-FLNA. Other examples, such as the noncanonical function of TSC/RHEB on the notch signaling pathway, have been reviewed elsewhere (Neuman and Henske, 2011). One of the best established functions of AKT is phosphorylation and inhibition of the glycogen synthase GSK3β (Bellacosa et al., 2004). GSK3β regulates many neuronal processes, including neuronal polarization, axon growth, and axon branching (Kim et al., 2011). Studies in the cancer field have shown that gain-of-function mutations in Pi3k and Akt, as well as Pten loss-of-function mutations, upregulate AKT and lead to decreased GSK3β activity (Duda et al., 2020). By contrast, Tsc1 loss-of-function or Rheb gain-of-function mutations indirectly downregulate AKT via a negative homeostatic feedback loop through mTORC1-S6K1/2, thereby increasing GSK3β activity (Meikle et al., 2008a; Gong et al., 2015; Shah et al., 2004). It is expected that variants of the GATOR1 complex would lead to decreased AKT activity through the same feedback mechanism, but this needs to be examined. Interestingly, despite Pten and Rheb mutants having opposite effects on GSK3b activity, Pten–/– and Rheb S16H neurons both displayed increased axon growth (Kwon et al., 2006), suggesting that too little or too much GSK3β activity leads to a similar axonal phenotype in hyperactive mTORC1 conditions. It was also reported that Rheb S6H leads to MAPK activation independent of mTOR that ultimately increases an actin-crosslinking molecule FLNA (Zhang et al., 2020). Knockdown of FLNA prevented neuronal misplacement and reduced soma and dendrite hypertrophy in Rheb S16H-expressing mice, supporting a role for FLNA in cortical mislamination and neuronal dysmorphogenesis in mTORopathies (Zhang et al., 2020). Further, this reduced seizure activity. Although alterations in FLNA were observed in Rheb and Tsc–/– conditions, it was not conserved with mTOR variants.
Perhaps the greatest divergence across gene variants is for synaptic activity. The diverging effects of Pi3k, Pten, Tsc1, Rheb, mTOR, and Depdc5 variants on synaptic activity emphasize a critical need to systematically investigate the functional effects of these variants side by side in the same system (i.e., IUE), cell type, cortical region, and age to parse out the sources of the differences (i.e., whether the observed effects are due to variant-specific functional differences giving rise to distinct effects or variable experimental conditions). Such side-by-side comparisons to investigate the effects of Pten and Tsc1 knockout on hippocampal neurons have been performed in vitro by Weston and colleagues (Weston et al., 2014). The authors found that Pten deletion led to an increase in both excitatory and inhibitory synaptic transmission, while Tsc1 deletion reduced inhibitory but did not affect excitatory synaptic transmission within the same culture system. These findings support diverging effects of Pten and Tsc1 on synaptic transmission, which may be accounted for by the differential impact of Pten and Tsc1 deletion on AKT activation (discussed above). In addition, Pten is present in the nucleus, where it modulates a range of functions, including chromosome stability and DNA replication (Chow and Salmena, 2020). Pten is also present at synapses and in synaptic spines, where it has been shown to regulate long-term depression (Wang et al., 2006; Jurado et al., 2010; Arendt et al., 2014; Sánchez-Puelles et al., 2020) and may play a role in memory impairment in Alzheimer’s disease (Knafo and Esteban, 2017; Knafo et al., 2016).
mTOR Hyperactivation in Acquired Epilepsy
Evidence establishing mTOR pathway mutations as a cause of genetic epilepsy is unambiguous and compelling. Somewhat unexpectedly, however, the mTOR signaling pathway is also implicated in the development of acquired epilepsy. Acquired epilepsy can result from a range of neurological injuries, such as status epilepticus, head trauma, or stroke. A role for mTOR signaling in acquired epilepsy is not necessarily intuitive, because acquired epilepsies are not associated with mTOR pathway mutations. However, enhanced mTOR signaling is evident in almost all acquired epilepsy models—presumably as part of the injury response. For example, mTOR hyperactivation occurs following systemic kainic acid-induced status epilepticus (Zeng et al., 2009; Macias et al., 2013), intrahippocampal kainic acid-induced status epilepticus (Tang et al., 2018; Gericke et al., 2020), systemic pilocarpine-induced status epilepticus (Huang et al., 2010; Okamoto et al., 2010), electrically induced status epilepticus (Drion et al., 2016), traumatic brain injury (Chen et al., 2006; Guo et al., 2013), and hypoxia (Talos et al., 2012b; Sun et al., 2013). Increased mTOR pathway activation following these injuries is typically assessed by immunostaining for proteins activated by the mTOR signaling cascade, such as phosphorylated S6 (Fig. 4–1). Increased signaling is evident within hours of injury, and elevated levels can persist for weeks (Zeng et al., 2008). Although human data for acute periods of epileptogenesis are not available, enhanced mTOR signaling is evident in tissue from patients with chronic idiopathic temporal lobe epilepsy (Sha et al., 2012; Sosunov et al., 2012; Talos et al., 2018).
Potential physiological and morphological effects of enhanced mTOR signaling after epileptogenic brain injury are less pronounced than the changes observed following genetic hyperactivation of mTOR but follow the same trends. Specifically, acquired epilepsy is associated with neuronal hypertrophy, dendritic overgrowth, spine changes, axonal sprouting, altered neuroplasticity and network hyperexcitability—all of which are features disrupted in mTOR pathway mutants. Use of rapamycin in animal models is beginning to reveal which changes are mTOR dependent. Most prominently, rapamycin has consistently been found to block hippocampal granule cell mossy fiber sprouting (Buckmaster et al., 2009; Yamawaki et al., 2015; Hester et al., 2016), in which granule cell axons form de novo connections with granule cell dendrites in the dentate inner molecular layer. Mossy fiber sprouting is a consistent pathological finding in human temporal lobe epilepsy, and in animal models of the disease. The recurrent excitatory connections created by sprouted mossy fibers have long been proposed as a mechanism of epileptogenesis (Sutula and Dudek, 2007; Hendricks et al., 2019), although their functional significance is debated (Buckmaster, 2014). Rapamycin treatment has also been found to prevent dendritic change in rodent models of epilepsy (Brewster et al., 2013), regulate changes in ion channel expression controlling neuronal excitability (Sosanya et al., 2015), and block increases in spontaneous excitatory postsynaptic current frequency (Tang et al., 2012).
Evidence for mTOR hyperactivation in animal models of acquired epilepsy has led to a host of studies aimed at determining whether blocking mTOR signaling with mTOR antagonists, like rapamycin, can suppress spontaneous seizures. Encouragingly, antagonists have been found to reduce spontaneous seizure frequency in many of these studies (Zeng et al., 2009; Huang et al., 2010; Raffo et al., 2011; Talos et al., 2012b; Russo et al., 2013; Guo et al., 2013; Butler et al., 2015; Drion et al., 2016; Chi et al., 2017). The literature also includes, however, many negative results (Buckmaster and Lew, 2011; Sliwa et al., 2012; Heng et al., 2013; Shima et al., 2015; Gericke et al., 2020). An additional unresolved question regards the extent to which rapamycin exerts true antiepileptogenic/disease modifying effects versus transient seizure-suppressive effects, like all existing antiseizure medications (ASMs). Studies showing recurrence of seizures after rapamycin withdrawal suggest it may act more as an ASM (Huang et al., 2010; Drion et al., 2016; Löscher, 2020); however, interpretation is complicated, as treatment may not eliminate underlying disease mechanisms (i.e., cell loss or injury) that could drive excess mTOR signaling, thus allowing the disease processes to resume once the drug is withdrawn. This would certainly appear to be the case for genetic mTORopathies, where treatment cannot fix the causal genetic lesion, requiring chronic treatment of patients with TSC (Franz and Capal, 2017).
The consistent ability of rapamycin to prevent disease in genetic models of mTORopathies, versus the variable effects in acquired epilepsies, is notable and suggests differences in underlying disease mechanisms. That treatment works well in models with genetic hyperactivating mutations of the mTOR pathway is not particularly surprising and supports the conclusion that excess mTOR signaling is the principal disease-driving mechanism. Mixed effects in acquired epilepsy, on the other hand, suggest that excess mTOR signaling is one of several epileptogenic mechanisms. Indeed, rapid loss of interneurons (hours to days) is prominent in many acquired epilepsy models and is hypothesized to be a key mechanism of epileptogenesis (Houser, 2014; Buckmaster et al., 2017). There is no evidence that mTOR inhibition could restore this nonregenerative cell population. By contrast, interneuron loss has been observed as a delayed, secondary effect in a Pten knockout mouse model of epilepsy (LaSarge et al., 2021), and therefore potentially preventable with early mTOR inhibition.
Inhibition of compensatory interneuron sprouting may be another mechanism that limits the efficacy of rapamycin in acquired epilepsy. Parvalbumin-expressing interneurons undergo extensive sprouting in epileptic rodents (Christenson Wick et al., 2017) and humans (Ábrahám et al., 2020). Similarly, somatostatin-expressing interneurons exhibit increased soma size, longer dendrites, axonal sprouting, and enhanced connectivity with excitatory cells in rodent epilepsy models (Buckmaster and Dudek, 1997; Zhang and Buckmaster, 2009; Halabisky et al., 2010; Hunt et al., 2011; Long et al., 2011; Drexel et al., 2012; Peng et al., 2013; Butler et al., 2017). In human temporal lobe epilepsy, somatostatin neurons also undergo sprouting (Mathern et al., 1995; DeLanerolle, 2012). Evidence suggests that interneuron sprouting and plasticity are mTOR dependent. In the pilocarpine model, for example, rapamycin blocks sprouting of inhibitory somatostatin neurons (Buckmaster and Wen, 2011). Rapamycin also reduces excitatory input to somatostatin neurons in the closed cortical impact model (Butler et al., 2017), and it reduces inhibitory input to granule cells (Butler et al., 2016). Directly supporting the idea that rapamycin can block protective interneuron sprouting, Jiang and colleagues (Jiang et al., 2018) demonstrated that rapamycin treatment blocked sprouting of somatostatin neurons in a mutant calcium channel model of epilepsy, increasing seizure severity. Together, these findings provide an instructive example of how systemic inhibition of mTOR could exert simultaneous anti- and pro-epileptogenic effects. Given the many roles and ubiquitous expression of mTOR, competing effects are likely to be common and represent a challenge for developing effective therapies.
Challenges and Opportunities
The mTOR pathway is emerging as an important regulator of epileptogenesis. Hyperactivating mutations in mTOR pathway genes consistently cause epilepsy in animal models and are linked to numerous childhood syndromes causing a range of neurological deficits, including epilepsy. Enhanced mTOR signaling is also evident in acquired epilepsy in both animal models and human temporal lobe epilepsy. Studies using mTOR inhibitors in animal models of acquired epilepsy support the conclusion that the pathway could promote epileptogenesis in some cases.
Significant challenges and unanswered questions, however, remain. While disease-causing mTOR mutations produce many common phenotypes, differential gene effects are also evident and remain incompletely characterized. The number of affected cells, mosaicism, and cell type-specific effects are also likely sources of significant variation. Given the diverse functions and ubiquitous expression of mTOR, involvement of organ systems outside the CNS may also impact epilepsy phenotypes. Finally, while many pro-excitatory effects of enhanced mTOR signaling have been identified, whether and how specific changes promote epileptogenesis remains an area of active investigation. Despite these challenges, the mTOR signaling pathway is an extremely promising target for the development of antiepileptogenic therapies. Initial successes in treating TSC with mTOR antagonists are encouraging, and clinical trials assessing the ability of early treatment with mTOR inhibitors to prevent epilepsy, not just reduce existing seizures, in TSC patients are in progress. As knowledge of mTOR signaling in epilepsy is expanded and refined, expansion of treatments to other epilepsies may be on the horizon.
Acknowledgments
This work was supported by the National Institute of Neurological Disorders and Stroke (SCD: R01NS065020, R01NS062806; MW: R01NS056872; AB: R01NS111980, R01NS093704). We would like to thank Christin Godale for help with the preparation of Figure 4–1. We would also like to thank Natasha Mayer and Maria Ashton for editorial assistance.
Disclosure Statement
The authors have no relevant conflicts of interests to report.
References
- Ábrahám, H., Molnár, J. E., Sóki, N., Gyimesi, C., Horváth, Z., Janszky, J., Dóczi, T. & Seress, L. 2020. Etiology-related degree of sprouting of parvalbumin-immunoreactive axons in the human dentate gyrus in temporal lobe epilepsy. Neuroscience, 448, 55–70. [PubMed: 32931846]
- Abs, E., Goorden, S. M., Schreiber, J., Overwater, I. E., Hoogeveen-westerveld, M., Bruinsma, C. F., Aganović, E., Borgesius, N. Z., Nellist, M. & Elgersma, Y. 2013. TORC1-dependent epilepsy caused by acute biallelic Tsc1 deletion in adult mice. Ann Neurol, 74, 569–79. [PubMed: 23720219]
- Alsaqati, M., Heine, V. M. & Harwood, A. J. 2020. Pharmacological intervention to restore connectivity deficits of neuronal networks derived from ASD patient iPSC with a TSC2 mutation. Mol Autism, 11, 80. [PMC free article: PMC7574213] [PubMed: 33076974]
- Amiri, A., Cho, W., Zhou, J., Birnbaum, S. G., Sinton, C. M., Mckay, R. M. & Parada, L. F. 2012. Pten deletion in adult hippocampal neural stem/progenitor cells causes cellular abnormalities and alters neurogenesis. J Neurosci, 32, 5880–90. [PMC free article: PMC6703621] [PubMed: 22539849]
- Arendt, K. L., Benoist, M., Lario, A., Draffin, J. E., Muñoz, M. & Esteban, J. A. 2014. PTEN counteracts PIP3 upregulation in spines during NMDA-receptor-dependent long-term depression. J Cell Sci, 127, 5253–60. [PubMed: 25335889]
- Backman, S. A., Stambolic, V., Suzuki, A., Haight, J., Elia, A., Pretorius, J., Tsao, M. S., Shannon, P., Bolon, B., Ivy, G. O. & Mak, T. W. 2001. Deletion of Pten in mouse brain causes seizures, ataxia and defects in soma size resembling Lhermitte-Duclos disease. Nat Genet, 29, 396–403. [PubMed: 11726926]
- Baek, S. T., Copeland, B., Yun, E. J., Kwon, S. K., Guemez-gamboa, A., Schaffer, A. E., Kim, S., Kang, H. C., Song, S., Mathern, G. W. & Gleeson, J. G. 2015. An AKT3-FOXG1-reelin network underlies defective migration in human focal malformations of cortical development. Nat Med, 21, 1445–54. [PMC free article: PMC4955611] [PubMed: 26523971]
- Baldassari, S., Ribierre, T., Marsan, E., Adle-biassette, H., Ferrand-sorbets, S., Bulteau, C., Dorison, N., Fohlen, M., Polivka, M., Weckhuysen, S., Dorfmüller, G., Chipaux, M. & Baulac, S. 2019. Dissecting the genetic basis of focal cortical dysplasia: a large cohort study. Acta Neuropathol, 138, 885–900. [PMC free article: PMC6851393] [PubMed: 31444548]
- Barkovich, A. J., Guerrini, R., Kuzniecky, R. I., Jackson, G. D. & Dobyns, W. B. 2012. A developmental and genetic classification for malformations of cortical development: update 2012. Brain, 135, 1348–69. [PMC free article: PMC3338922] [PubMed: 22427329]
- Bateup, H. S., Johnson, C. A., Denefrio, C. L., Saulnier, J. L., Kornacker, K. & Sabatini, B. L. 2013. Excitatory/inhibitory synaptic imbalance leads to hippocampal hyperexcitability in mouse models of tuberous sclerosis. Neuron, 78, 510–22. [PMC free article: PMC3690324] [PubMed: 23664616]
- Bateup, H. S., Takasaki, K. T., Saulnier, J. L., Denefrio, C. L. & Sabatini, B. L. 2011. Loss of Tsc1 in vivo impairs hippocampal mGluR-LTD and increases excitatory synaptic function. J Neurosci, 31, 8862–9. [PMC free article: PMC3133739] [PubMed: 21677170]
- Bellacosa, A., Testa, J. R., Moore, R. & Larue, L. 2004. A portrait of AKT kinases: human cancer and animal models depict a family with strong individualities. Cancer Biol Ther, 3, 268–75. [PubMed: 15034304]
- Beretta, L., Gingras, A. C., Svitkin, Y. V., Hall, M. N. & Sonenberg, N. 1996. Rapamycin blocks the phosphorylation of 4E-BP1 and inhibits cap-dependent initiation of translation. The EMBO journal, 15, 658–664. [PMC free article: PMC449984] [PubMed: 8599949]
- Blumcke, I., Thom, M., Aronica, E., Armstrong, D. D., Vinters, H. V., Palmini, A., Jacques, T. S., Avanzini, G., Barkovich, A. J., Battaglia, G., Becker, A., Cepeda, C., Cendes, F., Colombo, N., Crino, P., Cross, J. H., Delalande, O., Dubeau, F., Duncan, J., Guerrini, R., Kahane, P., Mathern, G., Najm, I., Ozkara, C., Raybaud, C., Represa, A., Roper, S. N., Salamon, N., Schulze-bonhage, A., Tassi, L., Vezzani, A. & Spreafico, R. 2011. The clinicopathologic spectrum of focal cortical dysplasias: a consensus classification proposed by an ad hoc Task Force of the ILAE Diagnostic Methods Commission. Epilepsia, 52, 158–74. [PMC free article: PMC3058866] [PubMed: 21219302]
- Blumcke, I., Vinters, H. V., Armstrong, D., Aronica, E., Thom, M. & Spreafico, R. 2009. Malformations of cortical development and epilepsies: neuropathological findings with emphasis on focal cortical dysplasia. Epileptic Disord, 11, 181–93. [PubMed: 19736171]
- Boer, K., Crino, P. B., Gorter, J. A., Nellist, M., Jansen, F. E., Spliet, W. G., Van rijen, P. C., Wittink, F. R., Breit, T. M., Troost, D., Wadman, W. J. & Aronica, E. 2010. Gene expression analysis of tuberous sclerosis complex cortical tubers reveals increased expression of adhesion and inflammatory factors. Brain Pathol, 20, 704–19. [PMC free article: PMC2888867] [PubMed: 19912235]
- Boer, K., Jansen, F., Nellist, M., Redeker, S., Van den ouweland, A. M., Spliet, W. G., Van nieuwenhuizen, O., Troost, D., Crino, P. B. & Aronica, E. 2008. Inflammatory processes in cortical tubers and subependymal giant cell tumors of tuberous sclerosis complex. Epilepsy Res, 78, 7–21. [PubMed: 18023148]
- Brewster, A. L., Lugo, J. N., Patil, V. V., Lee, W. L., Qian, Y., Vanegas, F. & Anderson, A. E. 2013. Rapamycin reverses status epilepticus-induced memory deficits and dendritic damage. PLoS One, 8, e57808. [PMC free article: PMC3594232] [PubMed: 23536771]
- Brugarolas, J. B., Vazquez, F., Reddy, A., Sellers, W. R. & Kaelin, W. G. 2003. TSC2 regulates VEGF through mTOR-dependent and -independent pathways. Cancer Cell, 4, 147–158. [PubMed: 12957289]
- Buckmaster, P. S. 2014. Does mossy fiber sprouting give rise to the epileptic state? Adv Exp Med Biol, 813, 161–8. [PubMed: 25012375]
- Buckmaster, P. S., Abrams, E. & Wen, X. 2017. Seizure frequency correlates with loss of dentate gyrus GABAergic neurons in a mouse model of temporal lobe epilepsy. J Comp Neurol, 525, 2592–2610. [PMC free article: PMC5963263] [PubMed: 28425097]
- Buckmaster, P. S. & Dudek, F. E. 1997. Neuron loss, granule cell axon reorganization, and functional changes in the dentate gyrus of epileptic kainate-treated rats. Journal of Comparative Neurology, 385, 385–404. [PubMed: 9300766]
- Buckmaster, P. S., Ingram, E. A. & Wen, X. 2009. Inhibition of the mammalian target of rapamycin signaling pathway suppresses dentate granule cell axon sprouting in a rodent model of temporal lobe epilepsy. J Neurosci, 29, 8259–69. [PMC free article: PMC2819377] [PubMed: 19553465]
- Buckmaster, P. S. & Lew, F. H. 2011. Rapamycin Suppresses Mossy Fiber Sprouting But Not Seizure Frequency in a Mouse Model of Temporal Lobe Epilepsy. The Journal of Neuroscience, 31, 2337–2347. [PMC free article: PMC3073836] [PubMed: 21307269]
- Buckmaster, P. S. & Wen, X. 2011. Rapamycin suppresses axon sprouting by somatostatin interneurons in a mouse model of temporal lobe epilepsy. Epilepsia, 52, 2057–2064. [PMC free article: PMC3204172] [PubMed: 21883182]
- Busch, R. M., Srivastava, S., Hogue, O., Frazier, T. W., Klaas, P., Hardan, A., Martinez-agosto, J. A., Sahin, M. & Eng, C. 2019. Neurobehavioral phenotype of autism spectrum disorder associated with germline heterozygous mutations in PTEN. Transl Psychiatry, 9, 253. [PMC free article: PMC6783427] [PubMed: 31594918]
- Butler, C. R., Boychuk, J. A. & Smith, B. N. 2015. Effects of Rapamycin Treatment on Neurogenesis and Synaptic Reorganization in the Dentate Gyrus after Controlled Cortical Impact Injury in Mice. Front Syst Neurosci, 9, 163. [PMC free article: PMC4661228] [PubMed: 26640431]
- Butler, C. R., Boychuk, J. A. & Smith, B. N. 2016. Differential effects of rapamycin treatment on tonic and phasic GABAergic inhibition in dentate granule cells after focal brain injury in mice. Exp Neurol, 280, 30–40. [PMC free article: PMC4860056] [PubMed: 27018320]
- Butler, C. R., Boychuk, J. A. & Smith, B. N. 2017. Brain Injury-Induced Synaptic Reorganization in Hilar Inhibitory Neurons Is Differentially Suppressed by Rapamycin. eneuro, 4, ENEURO.0134-17.2017. [PMC free article: PMC5659239] [PubMed: 29085896]
- Cai, H., Dong, L. Q. & Liu, F. 2016. Recent Advances in Adipose mTOR Signaling and Function: Therapeutic Prospects. Trends Pharmacol Sci, 37, 303–317. [PMC free article: PMC4811695] [PubMed: 26700098]
- Carson, R. P., Van nielen, D. L., Winzenburger, P. A. & Ess, K. C. 2012. Neuronal and glia abnormalities in Tsc1-deficient forebrain and partial rescue by rapamycin. Neurobiol Dis, 45, 369–80. [PMC free article: PMC3225598] [PubMed: 21907282]
- Chen, F. & Loturco, J. 2012. A method for stable transgenesis of radial glia lineage in rat neocortex by piggyBac mediated transposition. J Neurosci Methods, 207, 172–80. [PMC free article: PMC3972033] [PubMed: 22521325]
- Chen, F., Rosiene, J., Che, A., Becker, A. & Loturco, J. 2015. Tracking and transforming neocortical progenitors by CRISPR/Cas9 gene targeting and piggyBac transposase lineage labeling. Development, 142, 3601–11. [PMC free article: PMC4631763] [PubMed: 26400094]
- Chen, S., Atkins, C. M., Liu, C. L., Alonso, O. F., Dietrich, W. D. & Hu, B. R. 2006. Alterations in Mammalian Target of Rapamycin Signaling Pathways after Traumatic Brain Injury. Journal of Cerebral Blood Flow & Metabolism, 27, 939–949. [PubMed: 16955078]
- Chi, X., Huang, C., Li, R., Wang, W., Wu, M., Li, J. & Zhou, D. 2017. Inhibition of mTOR Pathway by Rapamycin Decreases P-glycoprotein Expression and Spontaneous Seizures in Pharmacoresistant Epilepsy. J Mol Neurosci, 61, 553–562. [PubMed: 28229367]
- Choi, Y. J., Di nardo, A., Kramvis, I., Meikle, L., Kwiatkowski, D. J., Sahin, M. & He, X. 2008. Tuberous sclerosis complex proteins control axon formation. Genes Dev, 22, 2485–95. [PMC free article: PMC2546692] [PubMed: 18794346]
- Chow, J. T. & Salmena, L. 2020. Recent advances in PTEN signalling axes in cancer. Fac Rev, 9, 31. [PMC free article: PMC7886076] [PubMed: 33659963]
- Christenson wick, Z., Leintz, C. H., Xamonthiene, C., Huang, B. H. & Krook-magnuson, E. 2017. Axonal sprouting in commissurally projecting parvalbumin-expressing interneurons. Journal of neuroscience research, 95, 2336–2344. [PMC free article: PMC5540851] [PubMed: 28151564]
- Chu-shore, C. J., Major, P., Camposano, S., Muzykewicz, D. & Thiele, E. A. 2010. The natural history of epilepsy in tuberous sclerosis complex. Epilepsia, 51, 1236–41. [PMC free article: PMC3065368] [PubMed: 20041940]
- Clipperton-allen, A. E. & Page, D. T. 2014. Pten haploinsufficient mice show broad brain overgrowth but selective impairments in autism-relevant behavioral tests. Hum Mol Genet, 23(13): 3490–505. [PubMed: 24497577]
- Crino, P. B., Aronica, E., Baltuch, G. & Nathanson, K. L. 2010. Biallelic TSC gene inactivation in tuberous sclerosis complex. Neurology, 74, 1716–23. [PMC free article: PMC2882213] [PubMed: 20498439]
- D’gama, A. M., Geng, Y., Couto, J. A., Martin, B., Boyle, E. A., Lacoursiere, C. M., Hossain, A., Hatem, N. E., Barry, B. J., Kwiatkowski, D. J., Vinters, H. V., Barkovich, A. J., Shendure, J., Mathern, G. W., Walsh, C. A. & Poduri, A. 2015. Mammalian target of rapamycin pathway mutations cause hemimegalencephaly and focal cortical dysplasia. Ann Neurol, 77, 720–5. [PMC free article: PMC4471336] [PubMed: 25599672]
- D’gama, A. M., Woodworth, M. B., Hossain, A. A., Bizzotto, S., Hatem, N. E., Lacoursiere, C. M., Najm, I., Ying, Z., Yang, E., Barkovich, A. J., Kwiatkowski, D. J., Vinters, H. V., Madsen, J. R., Mathern, G. W., Blumcke, I., Poduri, A. & Walsh, C. A. 2017. Somatic Mutations Activating the mTOR Pathway in Dorsal Telencephalic Progenitors Cause a Continuum of Cortical Dysplasias. Cell Rep, 21(13): 3754–3766. [PMC free article: PMC5752134] [PubMed: 29281825]
- Dang, L. T., Vaid, S., Lin, G., Swaminathan, P., Safran, J., Loughman, A., Lee, M., Glenn, T., Majolo, F., Crino, P. B. & Parent, J. M. 2021. STRADA-mutant human cortical organoids model megalencephaly and exhibit delayed neuronal differentiation. Dev Neurobiol. [PMC free article: PMC8364481] [PubMed: 33619909]
- Dawson, R. E., Nieto guil, A. F., Robertson, L. J., Piltz, S. G., Hughes, J. N. & Thomas, P. Q. 2020. Functional screening of GATOR1 complex variants reveals a role for mTORC1 deregulation in FCD and focal epilepsy. Neurobiol Dis, 134, 104640. [PubMed: 31639411]
- Delanerolle, N. 2012. Histopathology of Human Epilepsy. In: noebels jl, A. M., Rogawski MA, olsen RW, delgado-escueta AV (ed.) Jasper’s Basic Mechanisms of the Epilepsies. 4th ed. Bethesda (MD): National Center for Biotechnology Information (US), 567–593. [PubMed: 22787592]
- Di nardo, A., Wertz, M. H., Kwiatkowski, E., Tsai, P. T., Leech, J. D., Greene-colozzi, E., Goto, J., Dilsiz, P., Talos, D. M., Clish, C. B., Kwiatkowski, D. J. & Sahin, M. 2014. Neuronal Tsc1/2 complex controls autophagy through AMPK-dependent regulation of ULK1. Hum Mol Genet, 23, 3865–74. [PMC free article: PMC4065158] [PubMed: 24599401]
- Dibbens, L. M., De vries, B., Donatello, S., Heron, S. E., Hodgson, B. L., Chintawar, S., Crompton, D. E., Hughes, J. N., Bellows, S. T., Klein, K. M., Callenbach, P. M., Corbett, M. A., Gardner, A. E., Kivity, S., Iona, X., Regan, B. M., Weller, C. M., Crimmins, D., O’brien, T. J., Guerrero-lópez, R., Mulley, J. C., Dubeau, F., Licchetta, L., Bisulli, F., Cossette, P., Thomas, P. Q., Gecz, J., Serratosa, J., Brouwer, O. F., Andermann, F., Andermann, E., Van den maagdenberg, A. M., Pandolfo, M., Berkovic, S. F. & Scheffer, I. E. 2013. Mutations in DEPDC5 cause familial focal epilepsy with variable foci. Nat Genet, 45, 546–51. [PubMed: 23542697]
- Dooves, S., Van velthoven, A. J. H., Suciati, L. G. & Heine, V. M. 2021. Neuron-Glia Interactions in Tuberous Sclerosis Complex Affect the Synaptic Balance in 2D and Organoid Cultures. Cells, 10. [PMC free article: PMC7826837] [PubMed: 33445520]
- Drexel, M., Preidt, A. P. & Sperk, G. 2012. Sequel of spontaneous seizures after kainic acid-induced status epilepticus and associated neuropathological changes in the subiculum and entorhinal cortex. Neuropharmacology, 63, 806–817. [PMC free article: PMC3409872] [PubMed: 22722023]
- Drion, C. M., Borm, L. E., Kooijman, L., Aronica, E., Wadman, W. J., Hartog, A. F., Van vliet, E. A. & Gorter, J. A. 2016. Effects of rapamycin and curcumin treatment on the development of epilepsy after electrically induced status epilepticus in rats. Epilepsia, 57, 688–697. [PubMed: 26924447]
- Duda, P., Akula, S. M., Abrams, S. L., Steelman, L. S., Martelli, A. M., Cocco, L., Ratti, S., Candido, S., Libra, M., Montalto, G., Cervello, M., Gizak, A., Rakus, D. & Mccubrey, J. A. 2020. Targeting GSK3 and Associated Signaling Pathways Involved in Cancer. Cells, 9. [PMC free article: PMC7290852] [PubMed: 32365809]
- Ebrahimi-fakhari, D., Saffari, A., Wahlster, L., Di nardo, A., Turner, D., Lewis, T. L., JR., Conrad, C., Rothberg, J. M., Lipton, J. O., Kölker, S., Hoffmann, G. F., Han, M. J., Polleux, F. & Sahin, M. 2016. Impaired Mitochondrial Dynamics and Mitophagy in Neuronal Models of Tuberous Sclerosis Complex. Cell Rep, 17, 1053–1070. [PMC free article: PMC5078873] [PubMed: 27760312]
- Ehninger, D., Han, S., Shilyansky, C., Zhou, Y., Li, W., Kwiatkowski, D. J., Ramesh, V. & Silva, A. J. 2008. Reversal of learning deficits in a Tsc2+/- mouse model of tuberous sclerosis. Nat Med, 14, 843–8. [PMC free article: PMC2664098] [PubMed: 18568033]
- Ekici, Y., Emiroglu, R., Ozdemir, H., Aldemir, D., Karakayali, H. & Haberal, M. 2007. Effect of rapamycin on wound healing: an experimental study. Transplant Proc, 39, 1201–3. [PubMed: 17524932]
- Feliciano, D. M., Su, T., Lopez, J., Platel, J. C. & Bordey, A. 2011. Single-cell Tsc1 knockout during corticogenesis generates tuber-like lesions and reduces seizure threshold in mice. J Clin Invest, 121, 1596–607. [PMC free article: PMC3069783] [PubMed: 21403402]
- Feliciano, D. M., Zhang, S., Quon, J. L. & Bordey, A. 2013. Hypoxia-inducible factor 1a is a Tsc1-regulated survival factor in newborn neurons in tuberous sclerosis complex. Hum Mol Genet, 22, 1725–34. [PMC free article: PMC3613161] [PubMed: 23349360]
- Franz, D. N., Belousova, E., Sparagana, S., Bebin, E. M., Frost, M., Kuperman, R., Witt, O., Kohrman, M. H., Flamini, J. R., Wu, J. Y., Curatolo, P., De vries, P. J., Whittemore, V. H., Thiele, E. A., Ford, J. P., Shah, G., Cauwel, H., Lebwohl, D., Sahmoud, T. & Jozwiak, S. 2013. Efficacy and safety of everolimus for subependymal giant cell astrocytomas associated with tuberous sclerosis complex (EXIST-1): a multicentre, randomised, placebo-controlled phase 3 trial. Lancet, 381, 125–32. [PubMed: 23158522]
- Franz, D. N. & Capal, J. K. 2017. mTOR inhibitors in the pharmacologic management of tuberous sclerosis complex and their potential role in other rare neurodevelopmental disorders. Orphanet J Rare Dis, 12, 51. [PMC free article: PMC5348752] [PubMed: 28288694]
- French, J. A., Lawson, J. A., Yapici, Z., Ikeda, H., Polster, T., Nabbout, R., Curatolo, P., De vries, P. J., Dlugos, D. J., Berkowitz, N., Voi, M., Peyrard, S., Pelov, D. & Franz, D. N. 2016. Adjunctive everolimus therapy for treatment-resistant focal-onset seizures associated with tuberous sclerosis (EXIST-3): a phase 3, randomised, double-blind, placebo-controlled study. Lancet, 388, 2153–2163. [PubMed: 27613521]
- Fu, C., Cawthon, B., Clinkscales, W., Bruce, A., Winzenburger, P. & Ess, K. C. 2012. GABAergic interneuron development and function is modulated by the Tsc1 gene. Cereb Cortex, 22, 2111–9. [PMC free article: PMC3412444] [PubMed: 22021912]
- Ganesan, H., Balasubramanian, V., Iyer, M., Venugopal, A., Subramaniam, M. D., Cho, S. G. & Vellingiri, B. 2019. mTOR signalling pathway - A root cause for idiopathic autism? BMB Rep, 52, 424–433. [PMC free article: PMC6675248] [PubMed: 31186084]
- Garratt, M., Nakagawa, S. & Simons, M. J. 2016. Comparative idiosyncrasies in life extension by reduced mTOR signalling and its distinctiveness from dietary restriction. Aging Cell, 15, 737–43. [PMC free article: PMC4933670] [PubMed: 27139919]
- Gataullina, S., Lemaire, E., Wendling, F., Kaminska, A., Watrin, F., Riquet, A., Ville, D., Moutard, M. L., De saint martin, A., Napuri, S., Pedespan, J. M., Eisermann, M., Bahi-buisson, N., Nabbout, R., Chiron, C., Dulac, O. & Huberfeld, G. 2016. Epilepsy in young Tsc1(+/-) mice exhibits age-dependent expression that mimics that of human tuberous sclerosis complex. Epilepsia, 57, 648–59. [PubMed: 26873267]
- Gelot, A. B. & Represa, A. 2020. Progression of Fetal Brain Lesions in Tuberous Sclerosis Complex. Front Neurosci, 14, 899. [PMC free article: PMC7472962] [PubMed: 32973442]
- Gericke, B., Brandt, C., Theilmann, W., Welzel, L., Schidlitzki, A., Twele, F., Kaczmarek, E., Anjum, M., Hillmann, P. & Löscher, W. 2020. Selective inhibition of mTORC1/2 or PI3K/mTORC1/2 signaling does not prevent or modify epilepsy in the intrahippocampal kainate mouse model. Neuropharmacology, 162, 107817. [PubMed: 31654704]
- Getz, S. A., Despenza, T., Jr., Li, M. & Luikart, B. W. 2016. Rapamycin prevents, but does not reverse, aberrant migration in Pten knockout neurons. Neurobiol Dis, 93, 12–20. [PMC free article: PMC5409093] [PubMed: 26992888]
- Gong, X., Zhang, L., Huang, T., Lin, T. V., Miyares, L., Wen, J., Hsieh, L. & Bordey, A. 2015. Activating the translational repressor 4E-BP or reducing S6K-GSK3beta activity prevents accelerated axon growth induced by hyperactive mTOR in vivo. Hum Mol Genet, 24, 5746–58. [PMC free article: PMC4581604] [PubMed: 26220974]
- Goorden, S. M., Van woerden, G. M., Van der weerd, L., Cheadle, J. P. & Elgersma, Y. 2007. Cognitive deficits in Tsc1+/- mice in the absence of cerebral lesions and seizures. Ann Neurol, 62, 648–55. [PubMed: 18067135]
- Gorter, J. A., Aronica, E. & Van vliet, E. A. 2019. The Roof is Leaking and a Storm is Raging: Repairing the Blood-Brain Barrier in the Fight Against Epilepsy. Epilepsy Curr, 19, 177–181. [PMC free article: PMC6610387] [PubMed: 31037960]
- Goto, J., Talos, D. M., Klein, P., Qin, W., Chekaluk, Y. I., Anderl, S., Malinowska, I. A., Di nardo, A., Bronson, R. T., Chan, J. A., Vinters, H. V., Kernie, S. G., Jensen, F. E., Sahin, M. & Kwiatkowski, D. J. 2011. Regulable neural progenitor-specific Tsc1 loss yields giant cells with organellar dysfunction in a model of tuberous sclerosis complex. Proc Natl Acad Sci U S A, 108, E1070–9. [PMC free article: PMC3214999] [PubMed: 22025691]
- Goz, R. U., Akgul, G. & Loturco, J. J. 2020. BRAFV600E expression in neural progenitors results in a hyperexcitable phenotype in neocortical pyramidal neurons. J Neurophysiol, 123, 2449–2464. [PMC free article: PMC7311733] [PubMed: 32401131]
- Gregorian, C., Nakashima, J., Le belle, J., Ohab, J., Kim, R., Liu, A., Smith, K. B., Groszer, M., Garcia, A. D., Sofroniew, M. V., Carmichael, S. T., Kornblum, H. I., Liu, X. & Wu, H. 2009. Pten deletion in adult neural stem/progenitor cells enhances constitutive neurogenesis. J Neurosci, 29, 1874–86. [PMC free article: PMC2754186] [PubMed: 19211894]
- Greig, L. C., Woodworth, M. B., Galazo, M. J., Padmanabhan, H. & Macklis, J. D. 2013. Molecular logic of neocortical projection neuron specification, development and diversity. Nat Rev Neurosci, 14, 755–69. [PMC free article: PMC3876965] [PubMed: 24105342]
- Guemez-gamboa, A., Coufal, N. G. & Gleeson, J. G. 2014. Primary cilia in the developing and mature brain. Neuron, 82, 511–21. [PMC free article: PMC4104280] [PubMed: 24811376]
- Guo, D., Zeng, L., Brody, D. L. & Wong, M. 2013. Rapamycin attenuates the development of posttraumatic epilepsy in a mouse model of traumatic brain injury. PloS one, 8, e64078–e64078. [PMC free article: PMC3653881] [PubMed: 23691153]
- Hadjivassiliou, G., Martinian, L., Squier, W., Blumcke, I., Aronica, E., Sisodiya, S. M. & Thom, M. 2010. The application of cortical layer markers in the evaluation of cortical dysplasias in epilepsy. Acta Neuropathol, 120, 517–28. [PMC free article: PMC2923329] [PubMed: 20411268]
- Haji, N., Riebe, I., Aguilar-valles, A., Artinian, J., Laplante, I. & Lacaille, J. C. 2020. Tsc1 haploinsufficiency in Nkx2.1 cells upregulates hippocampal interneuron mTORC1 activity, impairs pyramidal cell synaptic inhibition, and alters contextual fear discrimination and spatial working memory in mice. Mol Autism, 11, 29. [PMC free article: PMC7201610] [PubMed: 32375878]
- Halabisky, B., Parada, I., Buckmaster, P. S. & Prince, D. A. 2010. Excitatory Input Onto Hilar Somatostatin Interneurons Is Increased in a Chronic Model of Epilepsy. Journal of Neurophysiology, 104, 2214–2223. [PMC free article: PMC3774571] [PubMed: 20631216]
- Hanai, S., Sukigara, S., Dai, H., Owa, T., Horike, S. I., Otsuki, T., Saito, T., Nakagawa, E., Ikegaya, N., Kaido, T., Sato, N., Takahashi, A., Sugai, K., Saito, Y., Sasaki, M., Hoshino, M., Goto, Y. I., Koizumi, S. & Itoh, M. 2017. Pathologic Active mTOR Mutation in Brain Malformation with Intractable Epilepsy Leads to Cell-Autonomous Migration Delay. Am J Pathol, 187, 1177–1185. [PubMed: 28427592]
- Harrison, D. E., Strong, R., Sharp, Z. D., Nelson, J. F., Astle, C. M., Flurkey, K., Nadon, N. L., Wilkinson, J. E., Frenkel, K., Carter, C. S., Pahor, M., Javors, M. A., Fernandez, E. & Miller, R. A. 2009. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature, 460, 392–5. [PMC free article: PMC2786175] [PubMed: 19587680]
- Heitman, J., Movva, N. R. & Hall, M. N. 1991. Targets for cell cycle arrest by the immunosuppressant rapamycin in yeast. Science, 253, 905–9. [PubMed: 1715094]
- Hendricks, W. D., Westbrook, G. L. & Schnell, E. 2019. Early detonation by sprouted mossy fibers enables aberrant dentate network activity. Proc Natl Acad Sci U S A, 116, 10994–10999. [PMC free article: PMC6561181] [PubMed: 31085654]
- Heng, K., Haney, M. M. & Buckmaster, P. S. 2013. High-dose rapamycin blocks mossy fiber sprouting but not seizures in a mouse model of temporal lobe epilepsy. Epilepsia, 54, 1535–1541. [PMC free article: PMC3769425] [PubMed: 23848506]
- Hester, M. S., Hosford, B. E., Santos, V. R., Singh, S. P., Rolle, I. J., Lasarge, C. L., Liska, J. P., Garcia-cairasco, N. & Danzer, S. C. 2016. Impact of rapamycin on status epilepticus induced hippocampal pathology and weight gain. Exp Neurol, 280, 1–12. [PMC free article: PMC4860167] [PubMed: 26995324]
- Houser, C. R. 2014. Do structural changes in GABA neurons give rise to the epileptic state? Advances in experimental medicine and biology, 813, 151–160. [PMC free article: PMC4634888] [PubMed: 25012374]
- Hsieh, L. S., Wen, J. H., Claycomb, K., Huang, Y., Harrsch, F. A., Naegele, J. R., Hyder, F., Buchanan, G. F. & Bordey, A. 2016. Convulsive seizures from experimental focal cortical dysplasia occur independently of cell misplacement. Nat Commun, 7, 11753. [PMC free article: PMC4895394] [PubMed: 27249187]
- Hsieh, L. S., Wen, J. H., Nguyen, L. H., Zhang, L., Getz, S. A., Torres-reveron, J., Wang, Y., Spencer, D. D. & Bordey, A. 2020. Ectopic HCN4 expression drives mTOR-dependent epilepsy in mice. Sci Transl Med. 12(570):eabc1492. [PMC free article: PMC9888000] [PubMed: 33208499]
- Hu, S., Knowlton, R. C., Watson, B. O., Glanowska, K. M., Murphy, G. G., Parent, J. M. & Wang, Y. 2018. Somatic Depdc5 deletion recapitulates electroclinical features of human focal cortical dysplasia type IIA. Ann Neurol, 84, 140–146. [PMC free article: PMC6119494] [PubMed: 30080265]
- Huang, X., Zhang, H., Yang, J., Wu, J., Mcmahon, J., Lin, Y., Cao, Z., Gruenthal, M. & Huang, Y. 2010. Pharmacological inhibition of the mammalian target of rapamycin pathway suppresses acquired epilepsy. Neurobiology of Disease, 40, 193–199. [PMC free article: PMC2926303] [PubMed: 20566381]
- Hunt, R. F., Scheff, S. W. & Smith, B. N. 2011. Synaptic Reorganization of Inhibitory Hilar Interneuron Circuitry after Traumatic Brain Injury in Mice. The Journal of Neuroscience, 31, 6880–6890. [PMC free article: PMC3102318] [PubMed: 21543618]
- Iffland, P. H., Everett, M. E., Cobb-pitstick, K. M., Bowser, L. E., Barnes, A. E., Babus, J. K., Romanowski, A., Baybis, M., Puffenberger, E. G., Gonzaga-jauregui, C., Poulopoulos, A., Carson, V. J. & Crino, P. B. 2022. NPRL3 loss alters neuronal morphology, mTOR localization, cortical lamination and seizure threshold. Brain. 21;145(11):3872–3885. [PMC free article: PMC10200289] [PubMed: 35136953]
- Ishida, S., Picard, F., Rudolf, G., Noé, E., Achaz, G., Thomas, P., Genton, P., Mundwiller, E., Wolff, M., Marescaux, C., Miles, R., Baulac, M., Hirsch, E., Leguern, E. & Baulac, S. 2013. Mutations of DEPDC5 cause autosomal dominant focal epilepsies. Nat Genet, 45, 552–5. [PMC free article: PMC5010101] [PubMed: 23542701]
- Itoh, Y., Higuchi, M., Oishi, K., Kishi, Y., Okazaki, T., Sakai, H., Miyata, T., Nakajima, K. & Gotoh, Y. 2016. PDK1-Akt pathway regulates radial neuronal migration and microtubules in the developing mouse neocortex. Proc Natl Acad Sci U S A, 113, E2955–64. [PMC free article: PMC4889351] [PubMed: 27170189]
- Jansen, L. A., Mirzaa, G. M., Ishak, G. E., O’roak, B. J., Hiatt, J. B., Roden, W. H., Gunter, S. A., Christian, S. L., Collins, S., Adams, C., Riviere, J. B., St-onge, J., Ojemann, J. G., Shendure, J., Hevner, R. F. & Dobyns, W. B. 2015. PI3K/AKT pathway mutations cause a spectrum of brain malformations from megalencephaly to focal cortical dysplasia. Brain, 138, 1613–28. [PMC free article: PMC4614119] [PubMed: 25722288]
- Jansen, L. A., Uhlmann, E. J., Crino, P. B., Gutmann, D. H. & Wong, M. 2005. Epileptogenesis and reduced inward rectifier potassium current in tuberous sclerosis complex-1-deficient astrocytes. Epilepsia, 46, 1871–80. [PubMed: 16393152]
- Jiang, X., Lupien-meilleur, A., Tazerart, S., Lachance, M., Samarova, E., Araya, R., Lacaille, J.-C. & Rossignol, E. 2018. Remodeled cortical inhibition prevents motor seizures in generalized epilepsy. Annals of Neurology, 84, 436–451. [PubMed: 30048010]
- Jones, R. G. & Pearce, E. J. 2017. MenTORing Immunity: mTOR Signaling in the Development and Function of Tissue-Resident Immune Cells. Immunity, 46, 730–742. [PMC free article: PMC5695239] [PubMed: 28514674]
- Jurado, S., Benoist, M., Lario, A., Knafo, S., Petrok, C. N. & Esteban, J. A. 2010. PTEN is recruited to the postsynaptic terminal for NMDA receptor-dependent long-term depression. Embo j, 29, 2827–40. [PMC free article: PMC2924645] [PubMed: 20628354]
- Ka, M., Smith, A. L. & Kim, W. Y. 2017. MTOR controls genesis and autophagy of GABAergic interneurons during brain development. Autophagy, 13, 1348–1363. [PMC free article: PMC5584862] [PubMed: 28598226]
- Kassai, H., Sugaya, Y., Noda, S., Nakao, K., Maeda, T., Kano, M. & Aiba, A. 2014a. Selective activation of mTORC1 signaling recapitulates microcephaly, tuberous sclerosis, and neurodegenerative diseases. Cell Rep, 7, 1626–1639. [PubMed: 24857653]
- Kassai, H., Sugaya, Y., Noda, S., Nakao, K., Maeda, T., Kano, M. & Aiba, A. 2014b. Selective activation of mTORC1 signaling recapitulates microcephaly, tuberous sclerosis, and neurodegenerative diseases. Cell Rep, 7, 1626–39. [PubMed: 24857653]
- Kast, R. J. & Levitt, P. 2019. Precision in the development of neocortical architecture: From progenitors to cortical networks. Prog Neurobiol, 175, 77–95. [PMC free article: PMC6402587] [PubMed: 30677429]
- Kelly, E., Schaeffer, S. M., Dhamne, S. C., Lipton, J. O., Lindemann, L., Honer, M., Jaeschke, G., Super, C. E., Lammers, S. H., Modi, M. E., Silverman, J. L., Dreier, J. R., Kwiatkowski, D. J., Rotenberg, A. & Sahin, M. 2018. mGluR5 Modulation of Behavioral and Epileptic Phenotypes in a Mouse Model of Tuberous Sclerosis Complex. Neuropsychopharmacology, 43, 1457–1465. [PMC free article: PMC5916364] [PubMed: 29206810]
- Kim, J. K., Cho, J., Kim, S. H., Kang, H. C., Kim, D. S., Kim, V. N. & Lee, J. H. 2019. Brain somatic mutations in MTOR reveal translational dysregulations underlying intractable focal epilepsy. J Clin Invest, 129, 4207–4223. [PMC free article: PMC6763223] [PubMed: 31483294]
- Kim, Y. T., Hur, E. M., Snider, W. D. & Zhou, F. Q. 2011. Role of GSK3 Signaling in Neuronal Morphogenesis. Front Mol Neurosci, 4, 48. [PMC free article: PMC3222852] [PubMed: 22131966]
- Klofas, L. K., Short, B. P., Snow, J. P., Sinnaeve, J., Rushing, G. V., Westlake, G., Weinstein, W., Ihrie, R. A., Ess, K. C. & Carson, R. P. 2020. DEPDC5 haploinsufficiency drives increased mTORC1 signaling and abnormal morphology in human iPSC-derived cortical neurons. Neurobiol Dis, 143, 104975. [PMC free article: PMC7462127] [PubMed: 32574724]
- Knafo, S. & Esteban, J. A. 2017. PTEN: Local and Global Modulation of Neuronal Function in Health and Disease. Trends Neurosci, 40, 83–91. [PubMed: 28081942]
- Knafo, S., Sánchez-puelles, C., Palomer, E., Delgado, I., Draffin, J. E., Mingo, J., Wahle, T., Kaleka, K., Mou, L., Pereda-perez, I., Klosi, E., Faber, E. B., Chapman, H. M., Lozano-montes, L., Ortega-molina, A., Ordóñez-gutiérrez, L., Wandosell, F., Viña, J., Dotti, C. G., Hall, R. A., Pulido, R., Gerges, N. Z., Chan, A. M., Spaller, M. R., Serrano, M., Venero, C. & Esteban, J. A. 2016. PTEN recruitment controls synaptic and cognitive function in Alzheimer’s models. Nat Neurosci, 19, 443–53. [PubMed: 26780512]
- Koene, L. M. C., Van grondelle, S. E., Proietti onori, M., Wallaard, I., Kooijman, N. H. R. M., Van oort, A., Schreiber, J. & Elgersma, Y. 2019. Effects of antiepileptic drugs in a new TSC/mTOR-dependent epilepsy mouse model. Annals of Clinical and Translational Neurology, 6, 1273–1291. [PMC free article: PMC6649373] [PubMed: 31353861]
- Konno, D., Yoshimura, S., Hori, K., Maruoka, H. & Sobue, K. 2005. Involvement of the phosphatidylinositol 3-kinase/rac1 and cdc42 pathways in radial migration of cortical neurons. J Biol Chem, 280, 5082–8. [PubMed: 15557338]
- Kútna, V., Uttl, L., Waltereit, R., Krištofiková, Z., Kaping, D., Petrásek, T., Hoschl, C. & Ovsepian, S. V. 2020. Tuberous Sclerosis (tsc2+/-) Model Eker Rats Reveals Extensive Neuronal Loss with Microglial Invasion and Vascular Remodeling Related to Brain Neoplasia. Neurotherapeutics, 17, 329–339. [PMC free article: PMC7007483] [PubMed: 31820275]
- Kwon, C.-H., Zhu, X., Zhang, J. & Baker, S. J. 2003. mTor is required for hypertrophy of Pten-deficient neuronal soma <em>in vivo</em>. Proceedings of the National Academy of Sciences, 100, 12923–12928. [PMC free article: PMC240720] [PubMed: 14534328]
- Kwon, C. H., Luikart, B. W., Powell, C. M., Zhou, J., Matheny, S. A., Zhang, W., Li, Y., Baker, S. J. & Parada, L. F. 2006. Pten regulates neuronal arborization and social interaction in mice. Neuron, 50, 377–88. [PMC free article: PMC3902853] [PubMed: 16675393]
- Lafourcade, C. A., Lin, T. V., Feliciano, D. M., Zhang, L., Hsieh, L. S. & Bordey, A. 2013. Rheb Activation in Subventricular Zone Progenitors Leads to Heterotopia, Ectopic Neuronal Differentiation, and Rapamycin-Sensitive Olfactory Micronodules and Dendrite Hypertrophy of Newborn Neurons. The Journal of Neuroscience, 33, 2419–2431. [PMC free article: PMC3711634] [PubMed: 23392671]
- Lasarge, C. L., Pun, R. Y. K., Gu, Z., Riccetti, M. R., Namboodiri, D. V., Tiwari, D., Gross, C. & Danzer, S. C. 2021. mTOR-driven neural circuit changes initiate an epileptogenic cascade. Prog Neurobiol, 200, 101974. [PMC free article: PMC8026598] [PubMed: 33309800]
- Lasarge, C. L., Santos, V. R. & Danzer, S. C. 2015. PTEN deletion from adult-generated dentate granule cells disrupts granule cell mossy fiber axon structure. Neurobiol Dis, 75, 142–50. [PMC free article: PMC4351143] [PubMed: 25600212]
- Lee, J. H., Huynh, M., Silhavy, J. L., Kim, S., Dixon-salazar, T., Heiberg, A., Scott, E., Bafna, V., Hill, K. J., Collazo, A., Funari, V., Russ, C., Gabriel, S. B., Mathern, G. W. & Gleeson, J. G. 2012. De novo somatic mutations in components of the PI3K-AKT3-mTOR pathway cause hemimegalencephaly. Nat Genet, 44, 941–5. [PMC free article: PMC4417942] [PubMed: 22729223]
- Lee, W. S., Baldassari, S., Chipaux, M., Adle-biassette, H., Stephenson, S. E. M., Maixner, W., Harvey, A. S., Lockhart, P. J., Baulac, S. & Leventer, R. J. 2021. Gradient of brain mosaic RHEB variants causes a continuum of cortical dysplasia. Ann Clin Transl Neurol, 8, 485–490. [PMC free article: PMC7886042] [PubMed: 33434304]
- Leontieva, O. V., Paszkiewicz, G. M. & Blagosklonny, M. V. 2012. Mechanistic or mammalian target of rapamycin (mTOR) may determine robustness in young male mice at the cost of accelerated aging. Aging (Albany NY), 4, 899–916. [PMC free article: PMC3615157] [PubMed: 23443503]
- Li, Y. H., Werner, H. & Puschel, A. W. 2008. Rheb and mTOR regulate neuronal polarity through Rap1B. J Biol Chem, 283, 33784–92. [PMC free article: PMC2662285] [PubMed: 18842593]
- Lim, J. S., Gopalappa, R., Kim, S. H., Ramakrishna, S., Lee, M., Kim, W. I., Kim, J., Park, S. M., Lee, J., Oh, J. H., Kim, H. D., Park, C. H., Lee, J. S., Kim, S., Kim, D. S., Han, J. M., Kang, H. C., Kim, H. H. & Lee, J. H. 2017. Somatic Mutations in TSC1 and TSC2 Cause Focal Cortical Dysplasia. Am J Hum Genet, 100, 454–472. [PMC free article: PMC5339289] [PubMed: 28215400]
- Lim, J. S., Kim, W. I., Kang, H. C., Kim, S. H., Park, A. H., Park, E. K., Cho, Y. W., Kim, S., Kim, H. M., Kim, J. A., Kim, J., Rhee, H., Kang, S. G., Kim, H. D., Kim, D., Kim, D. S. & Lee, J. H. 2015. Brain somatic mutations in MTOR cause focal cortical dysplasia type II leading to intractable epilepsy. Nat Med, 21, 395–400. [PubMed: 25799227]
- Lin, T. V., Hsieh, L., Kimura, T., Malone, T. J. & Bordey, A. 2016. Normalizing translation through 4E-BP prevents mTOR-driven cortical mislamination and ameliorates aberrant neuron integration. Proc Natl Acad Sci U S A, 113, 11330–11335. [PMC free article: PMC5056085] [PubMed: 27647922]
- Ljungberg, M. C., Sunnen, C. N., Lugo, J. N., Anderson, A. E. & D’arcangelo, G. 2009. Rapamycin suppresses seizures and neuronal hypertrophy in a mouse model of cortical dysplasia. Dis Model Mech, 2, 389–98. [PMC free article: PMC2707106] [PubMed: 19470613]
- Long, L., Xiao, B., Feng, L., Yi, F., Li, G., Li, S., Mutasem, M. A., Chen, S., Bi, F. & Li, Y. 2011. Selective Loss and Axonal Sprouting of GABAergic Interneurons in the Sclerotic Hippocampus Induced by LiCl–Pilocarpine. International Journal of Neuroscience, 121, 69–85. [PubMed: 21142829]
- Löscher, W. 2020. The holy grail of epilepsy prevention: Preclinical approaches to antiepileptogenic treatments. Neuropharmacology, 167, 107605. [PubMed: 30980836]
- Loturco, J., Manent, J. B. & Sidiqi, F. 2009. New and improved tools for in utero electroporation studies of developing cerebral cortex. Cereb Cortex, 19 Suppl 1, i120–5. [PMC free article: PMC3859267] [PubMed: 19395528]
- Lozovaya, N., Gataullina, S., Tsintsadze, T., Tsintsadze, V., Pallesi-pocachard, E., Minlebaev, M., Goriounova, N. A., Buhler, E., Watrin, F., Shityakov, S., Becker, A. J., Bordey, A., Milh, M., Scavarda, D., Bulteau, C., Dorfmuller, G., Delalande, O., Represa, A., Cardoso, C., Dulac, O., Ben-ari, Y. & Burnashev, N. 2014. Selective suppression of excessive GluN2C expression rescues early epilepsy in a tuberous sclerosis murine model. Nat Commun, 5, 4563. [PMC free article: PMC4143949] [PubMed: 25081057]
- Luikart, B. W., Schnell, E., Washburn, E. K., Bensen, A. L., Tovar, K. R. & Westbrook, G. L. 2011. Pten knockdown in vivo increases excitatory drive onto dentate granule cells. J Neurosci, 31, 4345–54. [PMC free article: PMC3113533] [PubMed: 21411674]
- Ma, X. M. & Blenis, J. 2009. Molecular mechanisms of mTOR-mediated translational control. Nat Rev Mol Cell Biol, 10, 307–18. [PubMed: 19339977]
- Macias, M., Blazejczyk, M., Kazmierska, P., Caban, B., Skalecka, A., Tarkowski, B., Rodo, A., Konopacki, J. & Jaworski, J. 2013. Spatiotemporal characterization of mTOR kinase activity following kainic acid induced status epilepticus and analysis of rat brain response to chronic rapamycin treatment. PloS one, 8, e64455–e64455. [PMC free article: PMC3665782] [PubMed: 23724051]
- Magri, L., Cambiaghi, M., Cominelli, M., Alfaro-cervello, C., Cursi, M., Pala, M., Bulfone, A., Garcìa-verdugo, J. M., Leocani, L., Minicucci, F., Poliani, P. L. & Galli, R. 2011. Sustained activation of mTOR pathway in embryonic neural stem cells leads to development of tuberous sclerosis complex-associated lesions. Cell Stem Cell, 9, 447–62. [PubMed: 22056141]
- Magri, L., Cominelli, M., Cambiaghi, M., Cursi, M., Leocani, L., Minicucci, F., Poliani, P. L. & Galli, R. 2013. Timing of mTOR activation affects tuberous sclerosis complex neuropathology in mouse models. Dis Model Mech, 6, 1185–97. [PMC free article: PMC3759338] [PubMed: 23744272]
- Malik, R., Pai, E. L., Rubin, A. N., Stafford, A. M., Angara, K., Minasi, P., Rubenstein, J. L., Sohal, V. S. & Vogt, D. 2019. Tsc1 represses parvalbumin expression and fast-spiking properties in somatostatin lineage cortical interneurons. Nat Commun, 10, 4994. [PMC free article: PMC6825152] [PubMed: 31676823]
- Marsan, E. & Baulac, S. 2018. Review: Mechanistic target of rapamycin (mTOR) pathway, focal cortical dysplasia and epilepsy. Neuropathology and Applied Neurobiology, 44, 6–17. [PubMed: 29359340]
- Martin, K. R., Zhou, W., Bowman, M. J., Shih, J., Au, K. S., Dittenhafer-reed, K. E., Sisson, K. A., Koeman, J., Weisenberger, D. J., Cottingham, S. L., Deroos, S. T., Devinsky, O., Winn, M. E., Cherniack, A. D., Shen, H., Northrup, H., Krueger, D. A. & Mackeigan, J. P. 2017. The genomic landscape of tuberous sclerosis complex. Nat Commun, 8, 15816. [PMC free article: PMC5481739] [PubMed: 28643795]
- Martinez-lopez, A., Salvador-rodriguez, L., Montero-vilchez, T., Molina-leyva, A., Tercedor-sanchez, J. & Arias-santiago, S. 2019. Vascular malformations syndromes: an update. Curr Opin Pediatr, 31, 747–753. [PubMed: 31693582]
- Mathern, G. W., Pretorius, J. K. & Babb, T. L. 1995. Quantified patterns of mossy fiber sprouting and neuron densities in hippocampal and lesional seizures. J Neurosurg, 82(2):211–9. [PubMed: 7815148]
- Matsushita, Y., Sakai, Y., Shimmura, M., Shigeto, H., Nishio, M., Akamine, S., Sanefuji, M., Ishizaki, Y., Torisu, H., Nakabeppu, Y., Suzuki, A., Takada, H. & Hara, T. 2016. Hyperactive mTOR signals in the proopiomelanocortin-expressing hippocampal neurons cause age-dependent epilepsy and premature death in mice. Sci Rep, 6, 22991. [PMC free article: PMC4785342] [PubMed: 26961412]
- Mcdaniel, S. S., Rensing, N. R., Thio, L. L., Yamada, K. A. & Wong, M. 2011. The ketogenic diet inhibits the mammalian target of rapamycin (mTOR) pathway. Epilepsia, 52, e7–11. [PMC free article: PMC3076631] [PubMed: 21371020]
- Mcmahon, J., Huang, X., Yang, J., Komatsu, M., Yue, Z., Qian, J., Zhu, X. & Huang, Y. 2012. Impaired autophagy in neurons after disinhibition of mammalian target of rapamycin and its contribution to epileptogenesis. J Neurosci, 32, 15704–14. [PMC free article: PMC3501684] [PubMed: 23136410]
- Meikle, L., Pollizzi, K., Egnor, A., Kramvis, I., Lane, H., Sahin, M. & Kwiatkowski, D. J. 2008a. Response of a neuronal model of tuberous sclerosis to mammalian target of rapamycin (mTOR) inhibitors: effects on mTORC1 and Akt signaling lead to improved survival and function. J Neurosci, 28, 5422–32. [PMC free article: PMC2633923] [PubMed: 18495876]
- Meikle, L., Pollizzi, K., Egnor, A., Kramvis, I., Lane, H., Sahin, M. & Kwiatkowski, D. J. 2008b. Response of a Neuronal Model of Tuberous Sclerosis to Mammalian Target of Rapamycin (mTOR) Inhibitors: Effects on mTORC1 and Akt Signaling Lead to Improved Survival and Function. The Journal of Neuroscience, 28, 5422–5432. [PMC free article: PMC2633923] [PubMed: 18495876]
- Meikle, L., Talos, D. M., Onda, H., Pollizzi, K., Rotenberg, A., Sahin, M., Jensen, F. E. & Kwiatkowski, D. J. 2007. A mouse model of tuberous sclerosis: neuronal loss of Tsc1 causes dysplastic and ectopic neurons, reduced myelination, seizure activity, and limited survival. J Neurosci, 27, 5546–58. [PMC free article: PMC6672762] [PubMed: 17522300]
- Mirzaa, G. M., Campbell, C. D., Solovieff, N., Goold, C. P., Jansen, L. A., Menon, S., Timms, A. E., Conti, V., Biag, J. D., Olds, C., Boyle, E. A., Collins, S., Ishak, G., Poliachik, S. L., Girisha, K. M., Yeung, K.-S., Chung, B. H. Y., Rahikkala, E., Gunter, S. A., Mcdaniel, S. S., Macmurdo, C. F., Bernstein, J. A., Martin, B., Leary, R. J., Mahan, S., Liu, S., Weaver, M., Dorschner, M. O., Jhangiani, S., Muzny, D. M., Boerwinkle, E., Gibbs, R. A., Lupski, J. R., Shendure, J., Saneto, R. P., Novotny, E. J., Wilson, C. J., Sellers, W. R., Morrissey, M. P., Hevner, R. F., Ojemann, J. G., Guerrini, R., Murphy, L. O., Winckler, W. & Dobyns, W. B. 2016. Association of MTOR Mutations With Developmental Brain Disorders, Including Megalencephaly, Focal Cortical Dysplasia, and Pigmentary Mosaicism. JAMA Neurology, 73, 836–845. [PMC free article: PMC4979321] [PubMed: 27159400]
- Mizuguchi, M., Takashima, S., Yamanouchi, H., Nakazato, Y., Mitani, H. & Hino, O. 2000. Novel cerebral lesions in the Eker rat model of tuberous sclerosis: cortical tuber and anaplastic ganglioglioma. J Neuropathol Exp Neurol, 59, 188–96. [PubMed: 10744057]
- Moon, U. Y., Park, J. Y., Park, R., Cho, J. Y., Hughes, L. J., Mckenna, J., 3rd, goetzl, L., Cho, S. H., Crino, P. B., Gambello, M. J. & Kim, S. 2015. Impaired Reelin-Dab1 Signaling Contributes to Neuronal Migration Deficits of Tuberous Sclerosis Complex. Cell Rep, 12, 965–78. [PMC free article: PMC4536164] [PubMed: 26235615]
- Morita, T. & Sobue, K. 2009. Specification of neuronal polarity regulated by local translation of CRMP2 and Tau via the mTOR-p70S6K pathway. J Biol Chem, 284, 27734–45. [PMC free article: PMC2785701] [PubMed: 19648118]
- Nadadhur, A. G., Alsaqati, M., Gasparotto, L., Cornelissen-steijger, P., Van hugte, E., Dooves, S., Harwood, A. J. & Heine, V. M. 2019. Neuron-Glia Interactions Increase Neuronal Phenotypes in Tuberous Sclerosis Complex Patient iPSC-Derived Models. Stem Cell Reports, 12, 42–56. [PMC free article: PMC6335594] [PubMed: 30581017]
- Nedredal, G. I., Picon, R. V., Chedid, M. F. & Foss, A. 2020. Immunosuppression in Liver Transplantation: State of the Art and Future Perspectives. Curr Pharm Des, 26, 3389–3401. [PubMed: 32520679]
- Neuman, N. A. & Henske, E. P. 2011. Non-canonical functions of the tuberous sclerosis complex-Rheb signalling axis. EMBO Mol Med, 3, 189–200. [PMC free article: PMC3377068] [PubMed: 21412983]
- Nguyen, L. H., Mahadeo, T. & Bordey, A. 2019. mTOR Hyperactivity Levels Influence the Severity of Epilepsy and Associated Neuropathology in an Experimental Model of Tuberous Sclerosis Complex and Focal Cortical Dysplasia. J Neurosci, 39, 2762–2773. [PMC free article: PMC6445990] [PubMed: 30700531]
- Normand, E. A., Crandall, S. R., Thorn, C. A., Murphy, E. M., Voelcker, B., Browning, C., Machan, J. T., Moore, C. I., Connors, B. W. & Zervas, M. 2013. Temporal and mosaic Tsc1 deletion in the developing thalamus disrupts thalamocortical circuitry, neural function, and behavior. Neuron, 78, 895–909. [PMC free article: PMC4529124] [PubMed: 23664552]
- Okamoto, O. K., Janjoppi, L., Bonone, F. M., Pansani, A. P., Da silva, A. V., Scorza, F. A. & Cavalheiro, E. A. 2010. Whole transcriptome analysis of the hippocampus: toward a molecular portrait of epileptogenesis. BMC Genomics, 11, 230. [PMC free article: PMC2859406] [PubMed: 20377889]
- Onori, M. P., Koene, L. M. C., Schafer, C. B., Nellist, M., De brito van velze, M., Gao, Z., Elgersma, Y. & Van woerden, G. M. 2021. RHEB/mTOR-hyperactivity causing cortical malformations drives seizures through increased axonal connectivity. PLoS Biol, 226;19(5):e3001279. [PMC free article: PMC8186814] [PubMed: 34038402]
- Orlova, K. A. & Crino, P. B. 2010. The tuberous sclerosis complex. Ann N Y Acad Sci, 1184, 87–105. [PMC free article: PMC2892799] [PubMed: 20146692]
- Orlova, K. A., Parker, W. E., Heuer, G. G., Tsai, V., Yoon, J., Baybis, M., Fenning, R. S., Strauss, K. & Crino, P. B. 2010. STRADalpha deficiency results in aberrant mTORC1 signaling during corticogenesis in humans and mice. J Clin Invest, 120, 1591–602. [PMC free article: PMC2860905] [PubMed: 20424326]
- Park, S. M., Lim, J. S., Ramakrishina, S., Kim, S. H., Kim, W. K., Lee, J., Kang, H. C., Reiter, J. F., Kim, D. S., Kim, H. H. & Lee, J. H. 2018. Brain Somatic Mutations in MTOR Disrupt Neuronal Ciliogenesis, Leading to Focal Cortical Dyslamination. Neuron, 99, 83–97 e7. [PubMed: 29937275]
- Parker, W. E., Orlova, K. A., Heuer, G. G., Baybis, M., Aronica, E., Frost, M., Wong, M. & Crino, P. B. 2011. Enhanced epidermal growth factor, hepatocyte growth factor, and vascular endothelial growth factor expression in tuberous sclerosis complex. Am J Pathol, 178, 296–305. [PMC free article: PMC3069836] [PubMed: 21224066]
- Parker, W. E., Orlova, K. A., Parker, W. H., Birnbaum, J. F., Krymskaya, V. P., Goncharov, D. A., Baybis, M., Helfferich, J., Okochi, K., Strauss, K. A. & Crino, P. B. 2013. Rapamycin prevents seizures after depletion of STRADA in a rare neurodevelopmental disorder. Sci Transl Med, 5, 182ra53. [PMC free article: PMC3720125] [PubMed: 23616120]
- Pelorosso, C., Watrin, F., Conti, V., Buhler, E., Gelot, A., Yang, X., Mei, D., Mcevoy-venneri, J., Manent, J. B., Cetica, V., Ball, L. L., Buccoliero, A. M., Vinck, A., Barba, C., Gleeson, J. G., Guerrini, R. & Represa, A. 2019. Somatic double-hit in MTOR and RPS6 in hemimegalencephaly with intractable epilepsy. Hum Mol Genet, 28, 3755–3765. [PMC free article: PMC6935386] [PubMed: 31411685]
- Peng, Z., Zhang, N., Wei, W., Huang, C. S., Cetina, Y., Otis, T. S. & Houser, C. R. 2013. A Reorganized GABAergic Circuit in a Model of Epilepsy: Evidence from Optogenetic Labeling and Stimulation of Somatostatin Interneurons. The Journal of Neuroscience, 33, 14392–14405. [PMC free article: PMC3761049] [PubMed: 24005292]
- Prabhakar, S., Goto, J., Zhang, X., Sena-esteves, M., Bronson, R., Brockmann, J., Gianni, D., Wojtkiewicz, G. R., Chen, J. W., Stemmer-rachamimov, A., Kwiatkowski, D. J. & Breakefield, X. O. 2013. Stochastic model of Tsc1 lesions in mouse brain. PLoS One, 8, e64224. [PMC free article: PMC3655945] [PubMed: 23696872]
- Pun, R. Y., Rolle, I. J., Lasarge, C. L., Hosford, B. E., Rosen, J. M., Uhl, J. D., Schmeltzer, S. N., Faulkner, C., Bronson, S. L., Murphy, B. L., Richards, D. A., Holland, K. D. & Danzer, S. C. 2012. Excessive activation of mTOR in postnatally generated granule cells is sufficient to cause epilepsy. Neuron, 75, 1022–34. [PMC free article: PMC3474536] [PubMed: 22998871]
- Qin, W., Chan, J. A., Vinters, H. V., Mathern, G. W., Franz, D. N., Taillon, B. E., Bouffard, P. & Kwiatkowski, D. J. 2010. Analysis of TSC cortical tubers by deep sequencing of TSC1, TSC2 and KRAS demonstrates that small second-hit mutations in these genes are rare events. Brain Pathol, 20, 1096–105. [PMC free article: PMC2951479] [PubMed: 20633017]
- Raab-graham, K. F., Haddick, P. C., Jan, Y. N. & Jan, L. Y. 2006. Activity- and mTOR-dependent suppression of Kv1.1 channel mRNA translation in dendrites. Science, 314, 144–8. [PubMed: 17023663]
- Raffo, E., Coppola, A., Ono, T., Briggs, S. W. & Galanopoulou, A. S. 2011. A pulse rapamycin therapy for infantile spasms and associated cognitive decline. Neurobiol Dis, 43, 322–9. [PMC free article: PMC3114281] [PubMed: 21504792]
- Reijnders, M. R. F., Kousi, M., Van woerden, G. M., Klein, M., Bralten, J., Mancini, G. M. S., Van essen, T., Proietti-onori, M., Smeets, E. E. J., Van gastel, M., Stegmann, A. P. A., Stevens, S. J. C., Lelieveld, S. H., Gilissen, C., Pfundt, R., Tan, P. L., Kleefstra, T., Franke, B., Elgersma, Y., Katsanis, N. & Brunner, H. G. 2017. Variation in a range of mTOR-related genes associates with intracranial volume and intellectual disability. Nat Commun, 8, 1052. [PMC free article: PMC5648772] [PubMed: 29051493]
- Ribierre, T., Deleuze, C., Bacq, A., Baldassari, S., Marsan, E., Chipaux, M., Muraca, G., Roussel, D., Navarro, V., Leguern, E., Miles, R. & Baulac, S. 2018. Second-hit mosaic mutation in mTORC1 repressor DEPDC5 causes focal cortical dysplasia-associated epilepsy. J Clin Invest, 128, 2452–2458. [PMC free article: PMC5983335] [PubMed: 29708508]
- Roy, A., Skibo, J., Kalume, F., Ni, J., Rankin, S., Lu, Y., Dobyns, W. B., Mills, G. B., Zhao, J. J., Baker, S. J. & Millen, K. J. 2015. Mouse models of human PIK3CA-related brain overgrowth have acutely treatable epilepsy. Elife, 4:e12703. [PMC free article: PMC4744197] [PubMed: 26633882]
- Russo, E., Citraro, R., Donato, G., Camastra, C., Iuliano, R., Cuzzocrea, S., Constanti, A. & De sarro, G. 2013. mTOR inhibition modulates epileptogenesis, seizures and depressive behavior in a genetic rat model of absence epilepsy. Neuropharmacology, 69, 25–36. [PubMed: 23092918]
- Sabatini, D. M. 2017. Twenty-five years of mTOR: Uncovering the link from nutrients to growth. Proc Natl Acad Sci U S A, 114, 11818–11825. [PMC free article: PMC5692607] [PubMed: 29078414]
- Sánchez-puelles, C., Calleja-felipe, M., Ouro, A., Bougamra, G., Arroyo, A., Diez, I., Erramuzpe, A., Cortés, J., Martínez-hernández, J., Luján, R., Navarrete, M., Venero, C., Chan, A., Morales, M., Esteban, J. A. & Knafo, S. 2020. PTEN Activity Defines an Axis for Plasticity at Cortico-Amygdala Synapses and Influences Social Behavior. Cereb Cortex, 30, 505–524. [PubMed: 31240311]
- Santos, V. R., Pun, R. Y. K., Arafa, S. R., Lasarge, C. L., Rowley, S., Khademi, S., Bouley, T., Holland, K. D., Garcia-cairasco, N. & Danzer, S. C. 2017. PTEN deletion increases hippocampal granule cell excitability in male and female mice. Neurobiol Dis, 108, 339–351. [PMC free article: PMC5675774] [PubMed: 28855130]
- Sarbassov, D. D., Ali, S. M. & Sabatini, D. M. 2005. Growing roles for the mTOR pathway. Current Opinion in Cell Biology, 17, 596–603. [PubMed: 16226444]
- Sarbassov, D. D., Ali, S. M., Sengupta, S., Sheen, J.-H., Hsu, P. P., Bagley, A. F., Markhard, A. L. & Sabatini, D. M. 2006. Prolonged Rapamycin Treatment Inhibits mTORC2 Assembly and Akt/PKB. Molecular Cell, 22, 159–168. [PubMed: 16603397]
- Sengupta, S., Peterson, T. R., Laplante, M., Oh, S. & Sabatini, D. M. 2010. mTORC1 controls fasting-induced ketogenesis and its modulation by ageing. Nature, 468, 1100–4. [PubMed: 21179166]
- Sha, L. Z., Xing, X. L., Zhang, D., Yao, Y., Dou, W. C., Jin, L. R., Wu, L. W. & Xu, Q. 2012. Mapping the spatio-temporal pattern of the mammalian target of rapamycin (mTOR) activation in temporal lobe epilepsy. PLoS One, 7, e39152. [PMC free article: PMC3384628] [PubMed: 22761730]
- Shah, O. J., Wang, Z. & Hunter, T. 2004. Inappropriate activation of the TSC/Rheb/mTOR/S6K cassette induces IRS1/2 depletion, insulin resistance, and cell survival deficiencies. Curr Biol, 14, 1650–6. [PubMed: 15380067]
- Shi, X., Lim, Y., Myers, A. K., Stallings, B. L., Mccoy, A., Zeiger, J., Scheck, J., Cho, G., Marsh, E. D., Mirzaa, G. M., Tao, T. & Golden, J. A. 2020. PIK3R2/Pik3r2 Activating Mutations Result in Brain Overgrowth and EEG Changes. Ann Neurol, 88, 1077–1094. [PMC free article: PMC8176885] [PubMed: 32856318]
- Shima, A., Nitta, N., Suzuki, F., Laharie, A.-M., Nozaki, K. & Depaulis, A. 2015. Activation of mTOR signaling pathway is secondary to neuronal excitability in a mouse model of mesio-temporal lobe epilepsy. European Journal of Neuroscience, 41, 976–988. [PubMed: 25605420]
- Skelton, P. D., Frazel, P. W., Lee, D., Suh, H. & Luikart, B. W. 2019. Pten loss results in inappropriate excitatory connectivity. Mol Psychiatry, 24, 1627–1640. [PMC free article: PMC6785382] [PubMed: 30967683]
- Skelton, P. D., Poquerusse, J., Salinaro, J. R., Li, M. & Luikart, B. W. 2020. Activity-dependent dendritic elaboration requires Pten. Neurobiol Dis, 134, 104703. [PMC free article: PMC7014817] [PubMed: 31838155]
- Sliwa, A., Plucinska, G., Bednarczyk, J. & Lukasiuk, K. 2012. Post-treatment with rapamycin does not prevent epileptogenesis in the amygdala stimulation model of temporal lobe epilepsy. Neuroscience Letters, 509, 105–109. [PubMed: 22227620]
- Sokolov, A. M., Seluzicki, C. M., Morton, M. C. & Feliciano, D. M. 2018. Dendrite growth and the effect of ectopic Rheb expression on cortical neurons. Neurosci Lett, 671, 140–147. [PMC free article: PMC5889745] [PubMed: 29447953]
- Sosanya, N. M., Brager, D. H., Wolfe, S., Niere, F. & Raab-graham, K. F. 2015. Rapamycin reveals an mTOR-independent repression of Kv1.1 expression during epileptogenesis. Neurobiol Dis, 73, 96–105. [PubMed: 25270294]
- Sosunov, A. A., Wu, X., Mcgovern, R. A., Coughlin, D. G., Mikell, C. B., Goodman, R. R. & Mckhann, G. M., 2nd 2012. The mTOR pathway is activated in glial cells in mesial temporal sclerosis. Epilepsia, 53 Suppl 1, 78–86. [PubMed: 22612812]
- Sun, H., Kosaras, B., Klein, P. M. & Jensen, F. E. 2013. Mammalian target of rapamycin complex 1 activation negatively regulates Polo-like kinase 2-mediated homeostatic compensation following neonatal seizures. Proceedings of the National Academy of Sciences, 110, 5199–5204. [PMC free article: PMC3612683] [PubMed: 23479645]
- Sutula, T. P. & Dudek, F. E. 2007. Unmasking recurrent excitation generated by mossy fiber sprouting in the epileptic dentate gyrus: an emergent property of a complex system. Prog Brain Res, 163, 541–63. [PubMed: 17765737]
- Switon, K., Kotulska, K., Janusz-kaminska, A., Zmorzynska, J. & Jaworski, J. 2017. Molecular neurobiology of mTOR. Neuroscience, 341, 112–153. [PubMed: 27889578]
- Takeuchi, K., Gertner, M. J., Zhou, J., Parada, L. F., Bennett, M. V. & Zukin, R. S. 2013. Dysregulation of synaptic plasticity precedes appearance of morphological defects in a Pten conditional knockout mouse model of autism. Proc Natl Acad Sci U S A, 110, 4738–43. [PMC free article: PMC3607034] [PubMed: 23487788]
- Talos, D. M., Jacobs, L. M., Gourmaud, S., Coto, C. A., Sun, H., Lim, K.-C., Lucas, T. H., Davis, K. A., Martinez-lage, M. & Jensen, F. E. 2018. Mechanistic target of rapamycin complex 1 and 2 in human temporal lobe epilepsy. Annals of neurology, 83, 311–327. [PMC free article: PMC5821556] [PubMed: 29331082]
- Talos, D. M., Kwiatkowski, D. J., Cordero, K., Black, P. M. & Jensen, F. E. 2008. Cell-specific alterations of glutamate receptor expression in tuberous sclerosis complex cortical tubers. Ann Neurol, 63, 454–65. [PMC free article: PMC2701384] [PubMed: 18350576]
- Talos, D. M., Sun, H., Kosaras, B., Joseph, A., Folkerth, R. D., Poduri, A., Madsen, J. R., Black, P. M. & Jensen, F. E. 2012a. Altered inhibition in tuberous sclerosis and type IIb cortical dysplasia. Ann Neurol, 71, 539–51. [PMC free article: PMC3334406] [PubMed: 22447678]
- Talos, D. M., Sun, H., Zhou, X., Fitzgerald, E. C., Jackson, M. C., Klein, P. M., Lan, V. J., Joseph, A. & Jensen, F. E. 2012b. The interaction between early life epilepsy and autistic-like behavioral consequences: a role for the mammalian target of rapamycin (mTOR) pathway. PloS one, 7, e35885–e35885. [PMC free article: PMC3342334] [PubMed: 22567115]
- Tang, H., Long, H., Zeng, C., Li, Y., Bi, F., Wang, J., Qian, H. & Xiao, B. 2012. Rapamycin suppresses the recurrent excitatory circuits of dentate gyrus in a mouse model of temporal lobe epilepsy. Biochem Biophys Res Commun, 420, 199–204. [PubMed: 22414694]
- Tang, S., Qin, F., Wang, X., Liang, Z., Cai, H., Mo, L., Huang, Y., Liang, B., Wei, X., Ao, Q., Xu, Y., Liu, Y., Xiao, D., Guo, S., Lu, C. & Li, X. 2018. H2O2 induces PP2A demethylation to downregulate mTORC1 signaling in HEK293 cells. Cell Biology International, 42, 1182–1191. [PubMed: 29752834]
- Tarkowski, B., Kuchcinska, K., Blazejczyk, M. & Jaworski, J. 2019. Pathological mTOR mutations impact cortical development. Hum Mol Genet, 28, 2107–2119. [PubMed: 30789219]
- Tavazoie, S. F., Alvarez, V. A., Ridenour, D. A., Kwiatkowski, D. J. & Sabatini, B. L. 2005. Regulation of neuronal morphology and function by the tumor suppressors Tsc1 and Tsc2. Nat Neurosci, 8, 1727–34. [PubMed: 16286931]
- Tokuda, S., Mahaffey, C. L., Monks, B., Faulkner, C. R., Birnbaum, M. J., Danzer, S. C. & Frankel, W. N. 2011. A novel Akt3 mutation associated with enhanced kinase activity and seizure susceptibility in mice. Hum Mol Genet, 20, 988–99. [PMC free article: PMC3033189] [PubMed: 21159799]
- Tsai, V., Parker, W. E., Orlova, K. A., Baybis, M., Chi, A. W., Berg, B. D., Birnbaum, J. F., Estevez, J., Okochi, K., Sarnat, H. B., Flores-sarnat, L., Aronica, E. & Crino, P. B. 2014. Fetal brain mTOR signaling activation in tuberous sclerosis complex. Cereb Cortex, 24, 315–27. [PMC free article: PMC3888364] [PubMed: 23081885]
- Uhlmann, E. J., Li, W., Scheidenhelm, D. K., Gau, C. L., Tamanoi, F. & Gutmann, D. H. 2004. Loss of tuberous sclerosis complex 1 (Tsc1) expression results in increased Rheb/S6K pathway signaling important for astrocyte cell size regulation. Glia, 47, 180–8. [PubMed: 15185396]
- Uhlmann, E. J., Wong, M., Baldwin, R. L., Bajenaru, M. L., Onda, H., Kwiatkowski, D. J., Yamada, K. & Gutmann, D. H. 2002. Astrocyte-specific TSC1 conditional knockout mice exhibit abnormal neuronal organization and seizures. Ann Neurol, 52, 285–96. [PubMed: 12205640]
- Urbanska, M., Gozdz, A., Swiech, L. J. & Jaworski, J. 2012. Mammalian target of rapamycin complex 1 (mTORC1) and 2 (mTORC2) control the dendritic arbor morphology of hippocampal neurons. The Journal of biological chemistry, 287, 30240–30256. [PMC free article: PMC3436277] [PubMed: 22810227]
- Van vliet, E. A., Forte, G., Holtman, L., Den burger, J. C., Sinjewel, A., De vries, H. E., Aronica, E. & Gorter, J. A. 2012. Inhibition of mammalian target of rapamycin reduces epileptogenesis and blood-brain barrier leakage but not microglia activation. Epilepsia, 53, 1254–63. [PubMed: 22612226]
- Van vliet, E. A., Otte, W. M., Wadman, W. J., Aronica, E., Kooij, G., De vries, H. E., Dijkhuizen, R. M. & Gorter, J. A. 2016a. Blood-brain barrier leakage after status epilepticus in rapamycin-treated rats I: Magnetic resonance imaging. Epilepsia, 57, 59–69. [PubMed: 26691904]
- Van vliet, E. A., Otte, W. M., Wadman, W. J., Aronica, E., Kooij, G., De vries, H. E., Dijkhuizen, R. M. & Gorter, J. A. 2016b. Blood-brain barrier leakage after status epilepticus in rapamycin-treated rats II: Potential mechanisms. Epilepsia, 57, 70–8. [PubMed: 26691741]
- Vanderplow, A. M., Eagle, A. L., Kermath, B. A., Bjornson, K. J., Robison, A. J. & Cahill, M. E. 2021. Akt-mTOR hypoactivity in bipolar disorder gives rise to cognitive impairments associated with altered neuronal structure and function. Neuron, 109, 1479–1496.e6. [PMC free article: PMC8105282] [PubMed: 33765445]
- Vézina, C., Kudelski, A. & Sehgal, S. N. 1975. Rapamycin (AY-22,989), a new antifungal antibiotic. I. Taxonomy of the producing streptomycete and isolation of the active principle. J Antibiot (Tokyo), 28, 721–6. [PubMed: 1102508]
- Vezzani, A. 2020. Brain Inflammation and Seizures: Evolving Concepts and New Findings in the Last 2 Decades. Epilepsy Curr, 20, 40s–43s. [PMC free article: PMC7726731] [PubMed: 33012196]
- Vogt, D., Cho, K. K. A., Lee, A. T., Sohal, V. S. & Rubenstein, J. L. R. 2015. The Parvalbumin/Somatostatin Ratio Is Increased in <em>Pten</em> Mutant Mice and by Human <em>PTEN</em> ASD Alleles. Cell Reports, 11, 944–956. [PMC free article: PMC4431948] [PubMed: 25937288]
- Waltereit, R., Welzl, H., Dichgans, J., Lipp, H. P., Schmidt, W. J. & Weller, M. 2006. Enhanced episodic-like memory and kindling epilepsy in a rat model of tuberous sclerosis. J Neurochem, 96, 407–13. [PubMed: 16300636]
- Wang, Y., Cheng, A. & Mattson, M. P. 2006. The PTEN phosphatase is essential for long-term depression of hippocampal synapses. Neuromolecular Med, 8, 329–36. [PubMed: 16775384]
- Way, S. W., Mckenna, J., 3rd, mietzsch, U., Reith, R. M., Wu, H. C. & Gambello, M. J. 2009. Loss of Tsc2 in radial glia models the brain pathology of tuberous sclerosis complex in the mouse. Hum Mol Genet, 18, 1252–65. [PMC free article: PMC2655769] [PubMed: 19150975]
- Way, S. W., Rozas, N. S., Wu, H. C., Mckenna, J., 3rd, reith, R. M., Hashmi, S. S., Dash, P. K. & Gambello, M. J. 2012. The differential effects of prenatal and/or postnatal rapamycin on neurodevelopmental defects and cognition in a neuroglial mouse model of tuberous sclerosis complex. Hum Mol Genet, 21, 3226–36. [PMC free article: PMC3384384] [PubMed: 22532572]
- Wen, S., Stolarov, J., Myers, M. P., Su, J. D., Wigler, M. H., Tonks, N. K. & Durden, D. L. 2001. PTEN controls tumor-induced angiogenesis. Proc Natl Acad Sci U S A, 98, 4622–7. [PMC free article: PMC31884] [PubMed: 11274365]
- Weston, M., Chen, H. & Swann, J. 2014. Loss of mTOR repressors Tsc1 or Pten has divergent effects on excitatory and inhibitory synaptic transmission in single hippocampal neuron cultures. Front Mol Neurosci, 7:1–15. [PMC free article: PMC3922082] [PubMed: 24574959]
- Williams MR, DeSpenza T Jr, Li M, Gulledge AT, Luikart BW. Hyperactivity of newborn Pten knock-out neurons results from increased excitatory synaptic drive. J Neurosci. 2015 Jan 21;35(3):943-59. [PMC free article: PMC4300333] [PubMed: 25609613]
- Winden, K. D., Sundberg, M., Yang, C., Wafa, S. M. A., Dwyer, S., Chen, P. F., Buttermore, E. D. & Sahin, M. 2019. Biallelic Mutations in TSC2 Lead to Abnormalities Associated with Cortical Tubers in Human iPSC-Derived Neurons. J Neurosci, 39, 9294–9305. [PMC free article: PMC6867816] [PubMed: 31591157]
- Wong, M. 2008. Mechanisms of epileptogenesis in tuberous sclerosis complex and related malformations of cortical development with abnormal glioneuronal proliferation. Epilepsia, 49, 8–21. [PMC free article: PMC3934641] [PubMed: 17727667]
- Wong, M. 2019. The role of glia in epilepsy, intellectual disability, and other neurodevelopmental disorders in tuberous sclerosis complex. J Neurodev Disord, 11, 30. [PMC free article: PMC6913020] [PubMed: 31838997]
- Wong, M., Ess, K. C., Uhlmann, E. J., Jansen, L. A., Li, W., Crino, P. B., Mennerick, S., Yamada, K. A. & Gutmann, D. H. 2003. Impaired glial glutamate transport in a mouse tuberous sclerosis epilepsy model. Ann Neurol, 54, 251–6. [PubMed: 12891680]
- Yamawaki, R., Thind, K. & Buckmaster, P. S. 2015. Blockade of excitatory synaptogenesis with proximal dendrites of dentate granule cells following rapamycin treatment in a mouse model of temporal lobe epilepsy. J Comp Neurol, 523, 281–97. [PMC free article: PMC4252507] [PubMed: 25234294]
- Yang, F., Yang, L., Wataya-kaneda, M., Teng, L. & Katayama, I. 2020. Epilepsy in a melanocyte-lineage mTOR hyperactivation mouse model: A novel epilepsy model. PLoS One, 15, e0228204. [PMC free article: PMC6980560] [PubMed: 31978189]
- Yuskaitis, C. J., Jones, B. M., Wolfson, R. L., Super, C. E., Dhamne, S. C., Rotenberg, A., Sabatini, D. M., Sahin, M. & Poduri, A. 2018. A mouse model of DEPDC5-related epilepsy: Neuronal loss of Depdc5 causes dysplastic and ectopic neurons, increased mTOR signaling, and seizure susceptibility. Neurobiol Dis, 111, 91–101. [PMC free article: PMC5803417] [PubMed: 29274432]
- Yuskaitis, C. J., Rossitto, L. A., Gurnani, S., Bainbridge, E., Poduri, A. & Sahin, M. 2019. Chronic mTORC1 inhibition rescues behavioral and biochemical deficits resulting from neuronal Depdc5 loss in mice. Hum Mol Genet, 28, 2952–2964. [PMC free article: PMC6736288] [PubMed: 31174205]
- Zeng, L.-H., Rensing, N. R. & Wong, M. 2009. The Mammalian Target of Rapamycin Signaling Pathway Mediates Epileptogenesis in a Model of Temporal Lobe Epilepsy. The Journal of Neuroscience, 29, 6964–6972. [PMC free article: PMC2727061] [PubMed: 19474323]
- Zeng, L.-H., Xu, L., Gutmann, D. H. & Wong, M. 2008. Rapamycin prevents epilepsy in a mouse model of tuberous sclerosis complex. Annals of Neurology, 63, 444–453. [PMC free article: PMC3937593] [PubMed: 18389497]
- Zeng, L. H., Bero, A. W., Zhang, B., Holtzman, D. M. & Wong, M. 2010. Modulation of astrocyte glutamate transporters decreases seizures in a mouse model of Tuberous Sclerosis Complex. Neurobiol Dis, 37, 764–71. [PMC free article: PMC2823985] [PubMed: 20045054]
- Zeng, L. H., Rensing, N. R., Zhang, B., Gutmann, D. H., Gambello, M. J. & Wong, M. 2011. Tsc2 gene inactivation causes a more severe epilepsy phenotype than Tsc1 inactivation in a mouse model of tuberous sclerosis complex. Hum Mol Genet, 20, 445–54. [PMC free article: PMC3016907] [PubMed: 21062901]
- Zhang, B., Zou, J., Han, L., Beeler, B., Friedman, J. L., Griffin, E., Piao, Y. S., Rensing, N. R. & Wong, M. 2018. The specificity and role of microglia in epileptogenesis in mouse models of tuberous sclerosis complex. Epilepsia, 59, 1796–1806. [PMC free article: PMC6121723] [PubMed: 30079598]
- Zhang, B., Zou, J., Rensing, N. R., Yang, M. & Wong, M. 2015. Inflammatory mechanisms contribute to the neurological manifestations of tuberous sclerosis complex. Neurobiol Dis, 80, 70–9. [PMC free article: PMC4468035] [PubMed: 26003087]
- Zhang, L., Huang, T., Teaw, S. & Bordey, A. 2019. Hypervascularization in mTOR-dependent focal and global cortical malformations displays differential rapamycin sensitivity. Epilepsia, 60, 1255–1265. [PMC free article: PMC6558978] [PubMed: 31125447]
- Zhang, L., Huang, T., Teaw, S., Nguyen, L. H., Hsieh, L. S., Gong, X., Burns, L. H. & Bordey, A. 2020. Filamin A inhibition reduces seizure activity in a mouse model of focal cortical malformations. Sci Transl Med, 12(531):eaay0289. [PubMed: 32075941]
- Zhang, W. & Buckmaster, P. S. 2009. Dysfunction of the dentate basket cell circuit in a rat model of temporal lobe epilepsy. J Neurosci, 29, 7846–56. [PMC free article: PMC2838908] [PubMed: 19535596]
- Zhao, J. P. & Yoshii, A. 2019. Hyperexcitability of the local cortical circuit in mouse models of tuberous sclerosis complex. Mol Brain, 12, 6. [PMC free article: PMC6347813] [PubMed: 30683131]
- Zhao, S., Li, Z., Zhang, M., Zhang, L., Zheng, H., Ning, J., Wang, Y., Wang, F., Zhang, X., Gan, H., Wang, Y., Zhang, X., Luo, H., Bu, G., Xu, H., Yao, Y. & Zhang, Y. W. 2019. A brain somatic RHEB doublet mutation causes focal cortical dysplasia type II. Exp Mol Med, 51, 84. [PMC free article: PMC6802736] [PubMed: 31337748]
- Zhao, X., Liao, Y., Morgan, S., Mathur, R., Feustel, P., Mazurkiewicz, J., Qian, J., Chang, J., Mathern, G. W., Adamo, M. A., Ritaccio, A. L., Gruenthal, M., Zhu, X. & Huang, Y. 2018. Noninflammatory Changes of Microglia Are Sufficient to Cause Epilepsy. Cell Rep, 22, 2080–2093. [PMC free article: PMC5880308] [PubMed: 29466735]
- Zhao, X. F., Liao, Y., Alam, M. M., Mathur, R., Feustel, P., Mazurkiewicz, J. E., Adamo, M. A., Zhu, X. C. & Huang, Y. 2020. Microglial mTOR is Neuronal Protective and Antiepileptogenic in the Pilocarpine Model of Temporal Lobe Epilepsy. J Neurosci, 40, 7593–7608. [PMC free article: PMC7531547] [PubMed: 32868461]
- Zhong, S., Zhao, Z., Xie, W., Cai, Y., Zhang, Y., Ding, J. & Wang, X. 2021. GABAergic Interneuron and Neurotransmission Are mTOR-Dependently Disturbed in Experimental Focal Cortical Dysplasia. Mol Neurobiol, 58, 156–169. [PubMed: 32909150]
- Zhou, J., Blundell, J., Ogawa, S., Kwon, C. H., Zhang, W., Sinton, C., Powell, C. M. & Parada, L. F. 2009. Pharmacological inhibition of mTORC1 suppresses anatomical, cellular, and behavioral abnormalities in neural-specific Pten knock-out mice. J Neurosci, 29, 1773–83. [PMC free article: PMC3904448] [PubMed: 19211884]
- Zimmer, T. S., Broekaart, D. W. M., Gruber, V. E., Van vliet, E. A., Mühlebner, A. & Aronica, E. 2020. Tuberous Sclerosis Complex as Disease Model for Investigating mTOR-Related Gliopathy During Epileptogenesis. Front Neurol, 11, 1028. [PMC free article: PMC7527496] [PubMed: 33041976]
- Zou, J., Zhang, B., Gutmann, D. H. & Wong, M. 2017. Postnatal reduction of tuberous sclerosis complex 1 expression in astrocytes and neurons causes seizures in an age-dependent manner. Epilepsia, 58, 2053–2063. [PMC free article: PMC5716871] [PubMed: 29023667]
- Zou, Z., Tao, T., Li, H. & Zhu, X. 2020. mTOR signaling pathway and mTOR inhibitors in cancer: progress and challenges. Cell Biosci, 10, 31. [PMC free article: PMC7063815] [PubMed: 32175074]
- Abstract
- The mTOR Signaling Pathway
- Clinical Disorders Caused by mTOR Pathway Mutations
- Genetic Models of mTORopathies
- Molecular Mechanisms of Epileptogenesis due to mTOR Hyperactivation
- Morphological and Pathological Effects of mTOR Hyperactivation on Cell Structure
- Genetic Effects: Not All mTORopathy Genes Are Equal
- mTOR Hyperactivation in Acquired Epilepsy
- Challenges and Opportunities
- Acknowledgments
- Disclosure Statement
- References
- mTOR Hyperactivity Levels Influence the Severity of Epilepsy and Associated Neuropathology in an Experimental Model of Tuberous Sclerosis Complex and Focal Cortical Dysplasia.[J Neurosci. 2019]mTOR Hyperactivity Levels Influence the Severity of Epilepsy and Associated Neuropathology in an Experimental Model of Tuberous Sclerosis Complex and Focal Cortical Dysplasia.Nguyen LH, Mahadeo T, Bordey A. J Neurosci. 2019 Apr 3; 39(14):2762-2773. Epub 2019 Jan 30.
- Review Epilepsy in the mTORopathies: opportunities for precision medicine.[Brain Commun. 2021]Review Epilepsy in the mTORopathies: opportunities for precision medicine.Moloney PB, Cavalleri GL, Delanty N. Brain Commun. 2021; 3(4):fcab222. Epub 2021 Sep 25.
- Review mTOR pathway inhibition as a new therapeutic strategy in epilepsy and epileptogenesis.[Pharmacol Res. 2016]Review mTOR pathway inhibition as a new therapeutic strategy in epilepsy and epileptogenesis.Citraro R, Leo A, Constanti A, Russo E, De Sarro G. Pharmacol Res. 2016 May; 107:333-343. Epub 2016 Apr 2.
- Review Mammalian target of rapamycin (mTOR) inhibition as a potential antiepileptogenic therapy: From tuberous sclerosis to common acquired epilepsies.[Epilepsia. 2010]Review Mammalian target of rapamycin (mTOR) inhibition as a potential antiepileptogenic therapy: From tuberous sclerosis to common acquired epilepsies.Wong M. Epilepsia. 2010 Jan; 51(1):27-36. Epub 2009 Oct 8.
- Review mTOR signaling in epilepsy: insights from malformations of cortical development.[Cold Spring Harb Perspect Med....]Review mTOR signaling in epilepsy: insights from malformations of cortical development.Crino PB. Cold Spring Harb Perspect Med. 2015 Apr 1; 5(4). Epub 2015 Apr 1.
- mTOR in Acquired and Genetic Models of Epilepsy - Jasper's Basic Mechanisms of t...mTOR in Acquired and Genetic Models of Epilepsy - Jasper's Basic Mechanisms of the Epilepsies
Your browsing activity is empty.
Activity recording is turned off.
See more...