Next Article in Journal
Alterations in Cellular Iron Metabolism Provide More Therapeutic Opportunities for Cancer
Previous Article in Journal
Frizzled Receptors as Potential Therapeutic Targets in Human Cancers
Previous Article in Special Issue
B-Myb Mediates Proliferation and Migration of Non-Small-Cell Lung Cancer via Suppressing IGFBP3
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of mTOR Complexes in Neurogenesis

Biology Program, School of Natural Sciences and Mathematics, Stockton University, Galloway, NJ 08205, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2018, 19(5), 1544; https://doi.org/10.3390/ijms19051544
Submission received: 30 April 2018 / Revised: 13 May 2018 / Accepted: 16 May 2018 / Published: 22 May 2018
(This article belongs to the Special Issue Cell Growth Regulation)

Abstract

:
Dysregulation of neural stem cells (NSCs) is associated with several neurodevelopmental disorders, including epilepsy and autism spectrum disorder. The mammalian target of rapamycin (mTOR) integrates the intracellular signals to control cell growth, nutrient metabolism, and protein translation. mTOR regulates many functions in the development of the brain, such as proliferation, differentiation, migration, and dendrite formation. In addition, mTOR is important in synaptic formation and plasticity. Abnormalities in mTOR activity is linked with severe deficits in nervous system development, including tumors, autism, and seizures. Dissecting the wide-ranging roles of mTOR activity during critical periods in development will greatly expand our understanding of neurogenesis.

Graphical Abstract

1. Introduction

Neurogenesis from the embryonic brain throughout adulthood requires a well-regulated interaction of inductive cues and secreted signals that guide the development of neural stem cells (NSCs). For a stem cell to transition into a properly functioning neuron it must process complex external and internal signals to divide, differentiate, migrate, and integrate properly. These stages are guided by extrinsic factors that promote the transcription of genes to define the cell through maturation. Many extracellular signals converge on the mammalian target of rapamycin (mTOR) pathway, which is a chief regulator of cell growth, proliferation, and protein translation. Several studies have shown that mTOR is involved in many aspects of neurogenesis. Activation of mTOR alters NSC differentiation, neural progenitor migration, dendrite development, and neuron maturation.
Disruption in mTOR pathway activity leads to a host of neurodevelopmental disorders, including the congenital multisystem disorder tuberous sclerosis (TS) [1,2,3]. Cortical malformations form in most TS patients during fetal and early postnatal development, and they contribute to the development of seizures and cognitive deficits [4,5]. More than 50% of TS patients develop varying degrees of Autism Spectrum Disorder (ASD) [6]. ASD patients exhibit impaired social interactions, repetitive behaviors, and diminished interests. Hyperconnectivity of neuronal circuits by the disruption in protein synthesis is thought to drive these altered behaviors [7]. As ASD is comorbid with TS, hyperactivity of the mTOR pathway is strongly implicated in driving both pathologies. Studies in mouse models of autism that alter the upstream elements of the mTOR pathway recapitulate cortical malformations [8,9,10,11]. In addition, many cortical neurons exhibit enlarged dendrites, altered dendritic spines, and hyperactivity, which can result in seizure episodes [12,13,14]. Additionally, increasing mTOR activity in neurons can result in ASD-like behaviors in mice [15]. Because mutations upstream of the regulators of the mTOR pathway in NSCs lead to hyperactivity in mature neurons, it is important to understand mTOR signaling during nervous system development. In this review, we describe our current understanding of the role of mTOR and related pathways at each stage of neurogenesis, focusing on embryonic cortical and postnatal development.

2. Overview of Cortical and Postnatal Neurogenesis

NSCs arise from the ectoderm with their fate being influenced by the timely activation of transcription factors and development mediated through the Akt/mTOR pathway. As development of the neural tube progresses, the rostral regions enlarge to form three primary vesicles, including the early prosencephalon [16]. The population of NSCs from the anterior neural tube develops into multipotent radial glia progenitors in the telencephalic vesicle. These divide to form radial glial cells (RGCs) at the most apical layer of the ventricular zone.
The majority of corticogenesis occurs in rodents from embryonic day 10.5 to embryonic day 18.5 where RGCs give rise to the majority of excitatory neurons in the cerebral cortex [16]. Primary divisions of RGCs are asymmetric and form a new RGC and intermediate progenitor. With continued development, intermediate progenitors of RGCs migrate from the apical ventricular zone and comprise the subventricular zone (SVZ), which is a second zone of proliferation [17]. Intermediate progenitors migrate along the apical processes of RGCs, delaminate, and begin to form the cortical plate where they continue to differentiate into post-mitotic glutamatergic neurons. Each cortical layer is produced sequentially in an inside-out manner where progenitors that are produced earlier populate the deepest layers of the cortex and neurons that will be in superficial layers are produced later [16]. In addition, NSCs in the ventral ganglionic eminences produce progenitors that migrate tangentially into the developing cortex where they establish the network of inhibitory interneurons [18,19]. Upon reaching their final location in the cortex, both types of neurons begin to extend dendrites, as well as a single axon in order to form synaptic connections necessary for mature neural circuits.
The postnatal SVZ is a transient state in the first week of postnatal life, during which the embryonic SVZ transitions to the adult SVZ [20,21]. RGCs begin to retract their apical processes that contact the pia, and their rates of proliferation slow [22,23]. Neurogenesis in the SVZ is a tightly regulated event where the NSCs expand their population and differentiate into neuroblasts that incorporate into developing circuits [24]. NSCs contact the lateral ventricles and blood vessels in the SVZ exposing them to important guidance molecules, such as IGF-1 and EGF, which can influence the downstream signaling cascades, like Akt and mTOR [25]. NSCs differentiate into daughter cells, called transit amplifying cells (TACs), which are highly proliferative. TACs, in turn, differentiate into neuroblasts that then migrate through the SVZ, into the rostral migratory stream (RMS) and become mature neurons in the olfactory bulb (OB). In addition, postnatal NSCs contribute to both neuron and glia populations, with the majority of newly born neurons migrating to the OB [26,27,28]. Newly born neurons mature and form connections that are stereotypic of their position in the OB. For example, newly born granule neurons extend a single apical dendrite that branch in the external plexiform layer to form dendro-dendritic connections with mitral cells. These cells form extensive dendritic arbors in the first three weeks of their development before pruning dendrites and spines as they mature [29].

3. The mTOR Signaling Network

Increasing protein synthesis in responses to growth and proliferation is essential to maintain homeostasis in all cell types. mTOR is a large serine/threonine protein kinase that is a part of the phosphoinositide 3-kinase (PI3K)-related kinase family with two divergent complexes, mTOR complex 1 (mTORC1) and 2 (mTORC2). mTORC1 integrates highly intracellular ATP, glucose, and certain amino acid signaling to regulate cellular processes (Figure 1). Low ATP levels indirectly inhibit mTORC1 by promoting TSC1/2 complex formation [30]. The TSC1/TSC2 complex normally suppresses the GTPase Rheb [31]. GTP-bound Rheb then activates mTORC1, which is a complex containing mTOR, Raptor, MLST8, and PRAS40 [32,33]. In addition to PI3K/ Akt activation, Ras-MAPK signaling can activate mTORC1 through the phosphorylation of TSC2 and PRAS40 [34,35,36]. Activation of mTORC1 results in increased protein translation by phosphorylating 4E-BP (including 4E-BP1/2/3) and p70 S6 kinases [37,38]. 4E-BPs and S6K1/2 inhibit the formation of 5’ translation initiating complex by binding eIF-4E. mTORC1-mediated phosphorylation of 4E-BPs and S6K1/2 liberates their respective binding partners, facilitating preinitiation complex formation and initiating translation [39,40].
In TS, a loss of function mutation in TSC1 or TSC2 leads to the hyperactivity of mTOR [41]. In animal models of TS, the hyperactivation of mTORC1 by TSC1 or TSC2 ablation increases cell size, cell survival, and the production of slow-growing tumors, which is consistent with unrestrained protein translation [42,43,44]. In addition to protein translation, mTORC1 activation also dampens the levels of autophagy via ULK1 [45]. Autophagy plays a critical role in the degradation and recycling of cellular components. Deletion of TSC2 increases both AMPK and mTORC1 inhibition of autophagy, thus promoting tumorigenesis in TS patients [46,47].
Both mTOR complexes share the binding partners DEPTOR and mLST8 in common; however, the rapamycin-insensitive Rictor is unique to mTORC2. Upstream regulators of mTORC2 are not well defined; however, growth factor signaling is known to activate mTORC2 [48]. Like mTORC1, there are multiple downstream targets of mTORC2, including PKC, SGK1, MST1, and Akt. Through PKC and MST1, mTORC2 can increase cell survival and proliferation [49]. Activation of mTORC2 results in phosphorylation of protein kinase B/Akt at serine 473, which, in turn, positively regulates mTORC1 activity [48,50]. Akt is a key regulator of survival during cellular stress, and dysfunction can promote the development of tumors. The mTORC2 binding partner mSin1 localizes the complex to the plasma membrane near Akt [51]. Deletion of the pleckstrin homology (PH) domain of mSin1 leads to constitutive mTORC2-AKT signaling. Recent studies have shown that the PH domain of mSin1 can inhibit mTORC2 complex formation and reduce Akt signaling [52]. Akt has several downstream targets that can affect both transcription and protein translation. Akt phosphorylates both forkhead box class O (FOXO) transcription factors and MDM2, leading to increased transcriptional activity and cell survival. Additionally, TORC2 activation in yeast, via Ypk1, regulates G2/M cell cycle entry, as well as actin organization and endocytosis [53,54].

4. mTOR Complexes in Cell Cycle Regulation

Proper development requires the maintenance of cell growth and cell cycle progression in stem cells, factors that are mediated by mTOR [55,56,57]. Insulin and growth factor signaling pathways activate PI3K through tyrosine kinase receptor activation. PI3K, in turn, stimulates Akt activation [58]. Akt also leads to GSK-3 phosphorylation, which allows for β-catenin and cyclin D1 activation to promote transcription and cell cycle progression [59,60]. There are three mammalian homologous isoforms of Akt. While the functions of each isoform overlap, it has been reported that Akt1 and Akt3, in particular, contribute to cell survival and nervous system development [61,62,63]. Increased levels of both Akt and mTOR have been observed in tumors; however, driving mTORC1 activity alone is not sufficient for cancer progression in cells due to the possible negative feedback on Akt from downstream mTOR targets [56].
The effects of mTOR activation on cell cycle and cell growth vary during the developmental time points. Blastocyst-stage mouse embryos that are treated with rapamycin display decreased trophoblast proliferation without affecting the pluripotent inner cell mass cells [64]. During the gastrulation stage, rapamycin treatment can reduce proliferation rates in primary germ layers. However, complete deletion of the kinase domain of mTOR shortly after implantation leads to embryonic lethality [64,65]. Additionally, the role of mTOR on cell cycle progression could be mediated through positive neurometabolic-vascular coupling in an increased number of G1–S cycling cells in the SVZ [66]. These data illustrate the importance of mTOR interaction amongst early developmental lineages and timing specificity.
Akt activity can regulate cell cycle progression, a process that is essential to the maintenance or the differentiation of stem cells. Levels of cyclin D1 is required for progression through G1 and could lead to complete mitosis [67]. Akt has been shown to stabilize the cell cycle regulatory protein p21Cip1/WAF1 through phosphorylation, which inhibits proliferating cell nuclear antigen (PCNA) and further promotes the assembly of the cyclin D1: cyclin-dependent kinase 4 (CDK4) complex [68]. Cell growth is reliant on nutritional availability and sensing environmental cues, mediated by insulin signaling to mTORC1 and further downstream signaling to S6K1 and 4E-BP1. The downstream effect of these signaling when nutrient and growth factor availability is high is the promotion of translation initiation [69].
Previous studies have shown that the alteration of upstream elements in mTORC1 signaling can lead to the enhanced proliferation in NSCs [70]. Conditional deletion of PTEN not only increases soma size and aberrant migration, but also increases cortical thickness from overproduction of neurons [71]. Deletion of PTEN in human neural progenitors cultured in brain organoids enhances proliferation and can stimulate cerebral folding [72]. UTX, a histone demethylase, promotes PTEN expression, and deletion of UTX during neural development enhances Akt and mTOR phosphorylation, increasing NSC proliferation [73]. Overexpression of Akt1 can induce proliferation in cortical progenitors [74]. Inhibition of mTORC1 activity in PTEN conditional knockout mice rescues soma hypertrophy and seizure frequencies [75,76]; however, mTORC1 inhibition only partially reverses morphological abnormalities, suggesting that Akt activation by PTEN deletion may exert pathological effects independent of mTORC1 [77]. These results suggest that cell proliferation is enhanced by Akt activation via mTORC1-independent mechanisms.
Although fewer studies have examined mTORC2 in NSC physiology, recent research has shown a role for mTORC2 in promoting NSC proliferation. IGF1 induces PI3K and mTORC2 activation in cortical progenitors, promoting proliferation [78]. Phosphorylation of Akt at S477 and T479 residues by mTORC2 promotes Akt activation and cell survival and cell cycle entry [79]. Notch is a key regulator of NSC self-renewal, proliferation, and differentiation [80]. Previous studies have shown that non-canonical notch signaling activates mTORC2/Akt signaling cascade, promoting both cell proliferation and tumor formation [81,82]. Recent studies have shown that the activation of mTORC1 at targets downstream of Akt do not promote NSC cell cycle entry [43,83,84]. Taken together, these studies suggest that mTORC2 can promote NSC proliferation through Akt activation, but mTORC1 activation alone is insufficient to promote cell cycle entry (Figure 2).

5. Activation of mTOR Complexes Promotes Neural Stem Cell Differentiation

The role of the two mTOR complexes vary in stem cell maintenance due to their different downstream targets. mTORC1/p70S6K is maintained at lower levels than mTORC2 in embryonic stem cells (ESCs) through TSC1/2 activation [85]. ESCs display reduced global translation rates and subsequently lower p70 S6K to maintain their undifferentiated state. Elevations in mTORC1 signaling through knockdown of TSC2 display an increase in protein synthesis and loss of stemness [85]. Complete blocking of global translation resulted in cell death, which suggests that mTORC1 is essential in low signaling environments [85]. The function of the higher levels of mTORC2 is not as clear but thought to be related to cytoskeletal functioning and its interaction with Akt [85]. Additionally, DEP domain-containing mTOR interacting protein (DEPTOR), which is an endogenous mTOR inhibitor, is upregulated in undifferentiated stages. The role of mTORC1 inhibition through DEPTOR has been shown as a critical factor in the maintenance of stemness, and it is downregulated upon differentiation [86].
Disrupting mTOR signaling during cortical development presents pathological defects. mTORC1 is critical for NSC differentiation into daughter cells, as reduced activity reduces neural progenitor cell populations [87]. This regulation is accomplished by the targeting of the translational repressor 4E-BP2 [83]. Additionally, insulin signaling promotes neuronal differentiation in NSCs via mTORC1 activation [88]. Deletion of Raptor in cortical progenitors results in microcephaly presenting around E17.5, indicating a strong role for mTORC1 in maintaining proper cell size and number. In this study, dense cortical layering was observed when compared to controls, but the positioning of neurons was not disturbed. Thus, the loss of Raptor is suggested to be specific to differentiation and migration in corticogenesis [89]. Furthermore, TSC1 deletion and Rheb activation promotes the differentiation and the misplacement of cortical progenitors in the embryonic brain [43,90]. Additionally, the role of mTORC2 in corticogenesis could be perturbed by the deletion of Rictor in development. Rictor deletion resulted in smaller cell size, yet not as profound as Raptor deletion, and altered neuron morphology when limited to purkinje cells [91]. Normalizing protein translation of hyperactive mTORC1 by blocking mTORC1-dependent phosphorylation of 4E-BPs is sufficient to prevent neuronal misplacement and cell enlargement [90].
In the first week of postnatal life, the embryonic ventricular zone (VZ) transitions to the adult SVZ with mTOR activity driving self-renewal or differentiation. The proliferation of RGCs slows and these precursors differentiate into astrocytes, ependymal cells, and adult NSCs in the SVZ. NSCs give rise to TACs that give rise to neuroblasts [66]. Forming a chain-like arrangement, the neuroblasts migrate through the rostral migratory stream to the olfactory bulb [25]. Postnatal and adult NSCs in the SVZ predominantly give rise to olfactory bulb interneurons and also some oligodendrocytes in corpus callosum, fimbria, and striatum [92]. mTORC1 activity in the SVZ is linked to neural lineage expansion, resulting in proliferative cells and newborn neuroblasts. Hyperactivation mTORC1 has been shown to induce symmetric differentiation of NSCs into proliferative daughter cells [84]. Sustained activation of mTORC1 by TSC1 deletion eventually leads to a reduction of proliferative stem cells in the SVZ [43]. This effect appears to be mediated directly by cap-dependent translation as knockdown of 4E-BP2 promotes the differentiation of NSCs into daughter cells [83]. Based on these studies, mTORC1 control of cap-dependent translation modulates the balance between self-renewal and differentiation in NSCs.

6. Hyperactivation of mTORC1 Results in Aberrant Migration

In the VZ, RGCs are responsible for generating the vast majority of excitatory and projections neurons of the cortex [93,94]. Proliferation of RGCs eventually produce intermediate progenitor cells that migrate into the SVZ and typically undergo at least one round of cell division to produce immature neurons [95,96]. The generation of Tbr2+ intermediate progenitors is sustained by mTORC1 activation [97]. In addition to intermediate progenitors in the SVZ, populations outer radial glial cells undergo successive rounds of division and contribute to the expansion of the neocortex in primates [98,99]. As additional neurons arise, they migrate into the preplate to form the cortical plate. Neurons that are generated in the proliferative zones of the cortex abruptly change in shape and direction as they populate the cortical plate [17]. Additional migrating neurons arrive in the cortical plate, bypassing earlier-generated neurons to form the cortical layers in an inside-out sequence [100]. Inhibitory GABAergic interneurons that populate the cortex arise from proliferation zones in the ventral ganglionic eminences, and tangentially migrate into the cortex [101]. These neurons migrate long distances that are guided by chemical cues in the external environment, which are likely secreted from endothelial cells of the developing vasculature [102].
The PI3K-Akt-mTOR pathway plays an essential role in the proper development of cortical layers and differentiation. Hyperactivation of this axis recapitulate pathology present in developmental disorders. Mutations of mTOR result in cortical delamination and dysmorphic neurons [103]. The formation of tubers and micro nodules are the most pronounced symptoms associated with mTOR pathway dysregulation. Enhanced mTOR activity in TS often leads to cortical malformations during development, slow-growing astrocytomas, and cognitive deficits [6]. Mouse models of TS exhibit similar aberrant migration during cortical development, enhancing the cortical thickness and tuber-like lesions [42,43,104]. Rapamycin treatment in animal models of cortical dysplasia rescue these cortical aberrations [61]. The conditional deletion of mTOR in progenitors of the ganglionic eminences led to a reduction in overall interneurons in the cortex [105]. Disrupted-in-Schizophrenia 1 (DISC1) suppresses Akt and mTOR activation, negatively regulating neuron development and migration in the adult hippocampus [106]. In addition, DISC1 knockdown reduces tangentially-migrating interneurons, reaching the embryonic cortex [107,108].
In the postnatal SVZ, NSCs give rise to migrating neuroblasts that navigate long distances to reach the OB. Upon reaching the OB, these newly born neurons tangentially migrate from the RMS to reach their final destinations in the granule and glomerular layers [109]. Deletion of TSC1 in postnatal NSCs leads to aberrant migration and heterotopias, mirroring results that were observed in embryonic mutations [110]. Similarly, the constitutive activation of Rheb enhances neurogenesis, producing greater numbers of neurons in the OB, while also generating ectopic neuronal differentiation in the RMS and olfactory micronodules [111]. Taken together, studies of both embryonic and postnatal neurogenesis reveal a critical role for mTORC1 in regulating the migration of newly born neurons to their proper destinations.

7. Role of mTORC1 in Neurite Development and Synapse Formation

Activation of mTOR complexes exerts profound effects on all the processes of neurogenesis, including dendrite formation. NSCs give rise to diverse populations of neurons, each of which had dendritic trees with disparate morphologies. The dendritic arborization established during maturation determines how the neuron integrates thousands of synaptic inputs during its existence. Although individual neuronal subtypes have specific programs during the formation of the dendrites, there are common features and steps that are shared by all populations of neurons. Upon reaching their terminal position, neurons extend dendrites away from the soma towards their target are guided by external signals. Next, dendrites grow in length and diameter, and they develop the branching characteristic of their subtype. Third, dendritic growths are spatially restricted by repulsive signals [112]. Fourth, dendrites typically start to form specialized structures, such as dendritic spines to enhance synaptic communication. Finally, dendrites and spines are pruned through retraction and elimination to form mature circuits [113].
Upon reaching the cortical plate, all of the excitatory cortical neurons share a similar morphology, which is a single apical dendrite that branches within layer 1 of the cortex. Over time, basal dendrites sprout from the soma and the apical dendrite branches. Dendritic differentiation only occurs after cells completed their migration. PTEN deletion enhances mTOR signaling and increases dendritic aborization in cortical neurons [77]. Similarly, PTEN deletion in hippocampal granule neurons increased dendritic branching and synaptic excitation [114]. In addition, driving mTORC1 via TSC1 knockout or Rheb activation also produces greatly enhanced dendritic branching [42,115]. Inhibition of cap-dependent translation through the overexpression of 4E-BP1 can partially rescue mTORC1-driven dendritic hypertrophy [90].
Granule neurons comprise the largest population of interneurons in the olfactory bulb. Their soma resides in the granule cell layer, and they extend both basal dendrites and a single apical dendrite that branch extensively in the external plexiform layer. Granule neurons are unique in that they are anaxonic, forming dendrodentritic synapses via spiny processes [116]. Several studies have shown that mTORC1 activation increases dendritic complexity in olfactory granule neurons. Postnatal deletion of TSC1 drives dendritic hypertrophy in granule neurons [110,117]. Moreover, Rheb activation increases dendritic branching of granule neurons, particularly basal dendrites [111,118]. Interestingly, the knockdown of global mTOR inhibits apical dendrite branching, but mTORC1-specific inhibition does not reduce dendrite complexity [119]. These results suggest that different mTOR complexes could regulate the apical and basal dendritic trees individually.
Sprouting from newly established dendrites, filopodia begin to emerge during synaptogenesis and form new synapses with nearby neurons. During this period of synaptic development, filopodia can recruit axons, making contacts that drive morphological changes in the young spine. Pruning occurs as synaptic activity matures, thus reducing the overall density of spines. TSC1/2 deletion in neuronal cultures show reduced spine formation with the remaining spines demonstrating an immature filopodia morphology [120,121]. These results were not duplicated when being tested in the postnatal hippocampus [122]. In addition, studies in the animal models of TS have not reached a clear consensus [14,90,123]. However, TSC2 heterozygous mice do display an increase in dendritic spine density at 1 month of age. As removal of unnecessary spines begin around three weeks following birth, this result could reflect a deficit in pruning rather than increased formation of filopodia-like projections. Indeed, rapamycin treatment rescues spine elimination in TSC2 heterozygous mice by enabling autophagy [124]. Reduction in autophagy could disrupt the synaptic transmission. Deletion of Atg7 decreases neurotransmission and ASD model mice have disrupted synaptic plasticity that is associated with glutamate signaling [125,126].
Long-term maintenance and the adaptability of the synapse is crucial for neuron survival and function. Long-term synaptic plasticity requires the expression of new proteins, which are produced both in the soma and locally in the dendrite. Activation of mTORC1 is central to the regulation of translation initiation in the dendritic spine. Glutamate signaling through mGluR and N-methyl-D-aspartate (NMDA) receptors upregulates mTORC1 activation, suggesting an important role for mTOR in synaptic plasticity [127,128]. While mTORC1 activation alone seems to be insufficient to alter dendritic spine shape and function, paired activation of synaptic stimulation, and mTORC1 increases synapse volatility [129]. Increased activation at synapses resulting from mTORC1 dysregulation has been observed in mouse models of TS [42,43]. Additionally, epilepsy is prevalent in patients with TS, and more than 50% of patients are comorbid with ASD. Hyper-connectivity of neuronal circuits by disruption in protein synthesis is thought to drive behavior and cognitive impairment in autism patients [7]. Additionally, a mouse model presented ASD-like behavioral deficits with hyperactivity of mTOR through TSC1/2 mutations. Rapamycin administration was successful in ameliorating the behavioral and the pathological phenotype in the mutant mice [15]. Dysregulation of mTORC1 appears to be integral to the formation and maintenance of synapses, foundational dysfunction that can result in epilepsy and ASD.

8. Conclusions

Throughout development of the nervous system, mTOR plays crucial roles in proliferation, differentiation, and neurite outgrowth and synaptic formation. Many of the recent studies into the role of the mTOR complexes have focused on mTORC1 and control of cap-dependent translation. Proper regulation of protein translation is critical in maintaining self-renewal of stem cells, dendrite formation, and synaptic plasticity. While good evidence exists that mTORC2 phosphorylates Akt, very few studies have demonstrated upstream regulators of mTORC2. Even fewer in vivo studies have examined the role of mTORC2 in neurogenesis. Future studies are needed to elucidate the role of mTORC2.
Regulation of synaptic development and activity is critical for circuit formation and function. To date, the disruption of mTOR signaling has resulted in varying degrees of dendrite alterations. Many studies have demonstrated that activation of mTORC1 leads to increased dendrite length and complexity, but there is still conflicting evidence regarding spine formation and density. These disparities could be due to differences in method of mTOR activation, neuronal subtype, or temporal activation of mTOR. More studies are needed to understand the mechanisms that control mTOR complexes during synapse formation. The mTOR complexes are essential to neurogenesis and the establishment of neural circuits. A multitude of external signals and transduction pathways modulate mTOR activity. Understanding the role of these kinase complexes is necessary to developing future therapies for current neurodevelopmental disorders.

Author Contributions

F.L and N.H. wrote the paper.

Acknowledgments

This work was funded in part by the US National Institutes of Health through a grant to Nathaniel W. Hartman (R15NS092026-01A1).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

4E-BP1Translation initiation factor 4e binding protein
4E-BP24E- protein phosphatase 2
AktProtein Kinase B
ASDAutism Spectrum Disorder
CDK4Cyclin-dependent kinase 4
DEPTOR DEP domain-containing mTOR interacting protein
DISC1Disrupted-in-Schizophrenia 1
EGFEpidermal growth factor
eIF-4eEukaryotic translation initiaton 4e
ErkExtracellular signal related kinase
ESCEmbryonic stem cells
FOXOForkhead box class O
Gsk-3βGlycogen synthase kinase 3 beta
IGFInsulin-like growth factor
MDM2Mouse double minute 2 homolog
MLST8mTOR associated protein LST8
mSin1Mammalian stress-activated protein kinase-interacting protein
mTORmammalian target of rapamycin
mTORC1mammalian target of rapamycin complex 1
mTORC2mammalian target of rapamycin complex 2
NMDAN-methyl-D-aspartate
NSCNeural stem cell
OBOlfactory bulb
P21Cip1/WAF-1Cyclin dependent kinase inhibitor 1
PCNAProliferating cell nuclear antigen
PDK1Phosphoinosidite-dependent kinase 1
PH
PI3K
Pleckstrin homology
Phosphoinositide 3 kinase
Pras40Proline-rich AKT substrate
PTENPhosphatase and tensin homolog
Raptor Regulatory-associated protein of mTOR
RGCRadial glial cell
RhebRas homolog enriched in brain
RictorRapamycin-insensitive companion of mTOR
RMSRostral migratory stream
S6k1S6 Kinase 1
SVZSub ventricular zone
TACTransit amplifying cell
Tbr2T-box brain protein 2
TSTuberous sclerosis
TSC1Tuberous sclerosis complex 1
TSC2Tuberous sclerosis complex 2
UTXUbiquitously transcribed tetratricopeptide repeat, X chromosome
VZVentricular zone
Ypk1Yeast serine/threonine protein kinase 1

References

  1. Crino, P.B.; Nathanson, K.L.; Henske, E.P. The tuberous sclerosis complex. N. Engl. J. Med. 2006, 355, 1345–1356. [Google Scholar] [CrossRef] [PubMed]
  2. Feliciano, D.M.; Lin, T.V.; Hartman, N.W.; Bartley, C.M.; Kubera, C.; Hsieh, L.; Lafourcade, C.; O’Keefe, R.A.; Bordey, A. A circuitry and biochemical basis for tuberous sclerosis symptoms: From epilepsy to neurocognitive deficits. Int. J. Dev. Neurosci. 2013, 31, 667–678. [Google Scholar] [CrossRef] [PubMed]
  3. Tee, A.R.; Sampson, J.R.; Pal, D.K.; Bateman, J.M. The role of mTOR signalling in neurogenesis, insights from tuberous sclerosis complex. Semin. Cell Dev. Biol. 2016, 52, 12–20. [Google Scholar] [CrossRef] [PubMed]
  4. Al Saleem, T.; Wessner, L.L.; Scheithauer, B.W.; Patterson, K.; Roach, E.S.; Dreyer, S.J.; Fujikawa, K.; Bjornsson, J.; Bernstein, J.; Henske, E.P. Malignant tumors of the kidney, brain, and soft tissues in children and young adults with the tuberous sclerosis complex. Cancer 1998, 83, 2208–2216. [Google Scholar] [CrossRef]
  5. Curatolo, P.; Moavero, R.; de Vries, P.J. Neurological and neuropsychiatric aspects of tuberous sclerosis complex. Lancet Neurol. 2015, 14, 733–745. [Google Scholar] [CrossRef]
  6. Ehninger, D.; de Vries, P.J.; Silva, A.J. From mTOR to cognition: molecular and cellular mechanisms of cognitive impairments in tuberous sclerosis. J. Intellect. Disabil. Res. 2009, 53, 838–851. [Google Scholar] [CrossRef] [PubMed]
  7. Supekar, K.; Uddin, L.Q.; Khouzam, A.; Phillips, J.; Gaillard, W.D.; Kenworthy, L.E.; Yerys, B.E.; Vaidya, C.J.; Menon, V. Brain Hyperconnectivity in Children with Autism and its Links to Social Deficits. Cell Rep. 2013, 5, 738–747. [Google Scholar] [CrossRef] [PubMed]
  8. Hardan, A.Y.; Libove, R.A.; Keshavan, M.S.; Melhem, N.M.; Minshew, N.J. Cortical thickness in autism. Biol. Psychiatry 2010, 66, 320–326. [Google Scholar] [CrossRef] [PubMed]
  9. Hardan, A.Y.; Muddasani, S.; Vemulapalli, M.; Keshavan, M.S.; Minshew, N.J. An MRI study of increased cortical thickness in autism. Am. J. Psychiatry 2006, 163, 1290–1292. [Google Scholar] [CrossRef] [PubMed]
  10. Poduri, A.; Evrony, G.D.; Cai, X.; Elhosary, P.C.; Beroukhim, R.; Lehtinen, M.K.; Hills, B.L.; Heinzen, E.L.; Hill, A.; Hill, S.R.; et al. Somatic Activation of AKT3 Causes Hemispheric Developmental Brain Malformations. Neuron 2012, 74, 41–48. [Google Scholar] [CrossRef] [PubMed]
  11. Kwon, C.H.; Luikart, B.W.; Powell, C.M.; Zhou, J.; Matheny, S.A.; Zhang, W.; Li, Y.; Baker, S.J.; Parada, L.F. Pten Regulates Neuronal Arborization and Social Interaction in Mice. Neuron 2006, 50, 377–388. [Google Scholar] [CrossRef] [PubMed]
  12. Backman, S.A.; Stambolic, V.; Suzuki, A.; Haight, J.; Elia, A.; Pretorius, J.; Tsao, M.S.; Shannon, P.; Bolon, B.; Ivy, G.O.; et al. Deletion of Pten in mouse brain causes seizures, ataxia and defects in soma size resembling Lhermitte-Duclos disease. Nat. Genet. 2001, 29, 396–403. [Google Scholar] [CrossRef] [PubMed]
  13. Markus, A.; Zhong, J.; Snider, W.D. Raf and Akt mediate distinct aspects of sensory axon growth. Neuron 2002, 35, 65–76. [Google Scholar] [CrossRef]
  14. Kumar, V.; Zhang, M.-X.; Swank, M.W.; Kunz, J.; Wu, G.-Y. Regulation of dendritic morphogenesis by Ras-PI3K-Akt-mTOR and Ras-MAPK signaling pathways. J. Neurosci. 2005, 25, 11288–11299. [Google Scholar] [CrossRef] [PubMed]
  15. Tsai, P.T.; Hull, C.; Chu, Y.; Greene-Colozzi, E.; Sadowski, A.R.; Leech, J.M.; Steinberg, J.; Crawley, J.N.; Regehr, W.G.; Sahin, M. Autistic-like behaviour and cerebellar dysfunction in Purkinje cell Tsc1 mutant mice. Nature 2012, 488, 647–651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Götz, M.; Huttner, W.B. The cell biology of neurogenesis. Nat. Rev. Mol. Cell Biol. 2005, 6, 777–788. [Google Scholar] [CrossRef] [PubMed]
  17. Noctor, S.C.; Martínez-Cerdeño, V.; Ivic, L.; Kriegstein, A.R. Cortical neurons arise in symmetric and asymmetric division zones and migrate through specific phases. Nat. Neurosci. 2004, 7, 136–144. [Google Scholar] [CrossRef] [PubMed]
  18. Liodis, P.; Denaxa, M.; Grigoriou, M.; Akufo-Addo, C.; Yanagawa, Y.; Pachnis, V. Lhx6 Activity Is Required for the Normal Migration and Specification of Cortical Interneuron Subtypes. J. Neurosci. 2007, 27, 3078–3089. [Google Scholar] [CrossRef] [PubMed]
  19. Silva, C.G.; Peyre, E.; Adhikari, M.H.; Tielens, S.; Tanco, S.; van Damme, P.; Magno, L.; Krusy, N.; Agirman, G.; Magiera, M.M.; et al. Cell-Intrinsic Control of Interneuron Migration Drives Cortical Morphogenesis. Cell 2018, 172, 1063–1067. [Google Scholar] [CrossRef] [PubMed]
  20. Cayre, M.; Canoll, P.; Goldman, J.E. Cell migration in the normal and pathological postnatal mammalian brain. Prog. Neurobiol. 2009, 88, 41–63. [Google Scholar] [CrossRef] [PubMed]
  21. Merkle, F.T.; Tramontin, A.D.; García-Verdugo, J.M.; Alvarez-Buylla, A. Radial glia give rise to adult neural stem cells in the subventricular zone. Proc. Natl. Acad. Sci. USA 2004, 101, 17528–17532. [Google Scholar] [CrossRef] [PubMed]
  22. Lacar, B.; Young, S.Z.; Platel, J.C.; Bordey, A. Imaging and recording subventricular zone progenitor cells in live tissue of postnatal mice. Front. Neurosci. 2010, 4. [Google Scholar] [CrossRef] [PubMed]
  23. Tramontin, A.D.; García-Verdugo, J.M.; Lim, D.A.; Alvarez-Buylla, A. Postnatal development of radial glia and the ventricular zone (VZ): A continuum of the neural stem cell compartment. Cereb. Cortex 2003, 13, 580–587. [Google Scholar] [CrossRef] [PubMed]
  24. Platel, J.C.; Dave, K.A.; Gordon, V.; Lacar, B.; Rubio, M.E.; Bordey, A. NMDA Receptors Activated by Subventricular Zone Astrocytic Glutamate Are Critical for Neuroblast Survival Prior to Entering a Synaptic Network. Neuron 2010, 65, 859–872. [Google Scholar] [CrossRef] [PubMed]
  25. Kokovay, E.; Goderie, S.; Wang, Y.; Lotz, S.; Lin, G.; Sun, Y.; Roysam, B.; Shen, Q.; Temple, S. Adult svz lineage cells home to and leave the vascular niche via differential responses to SDF1/CXCR4 signaling. Cell Stem Cell 2010, 7, 163–173. [Google Scholar] [CrossRef] [PubMed]
  26. Garcia, A.D.R.; Doan, N.B.; Imura, T.; Bush, T.G.; Sofroniew, M.V. GFAP-expressing progenitors are the principal source of constitutive neurogenesis in adult mouse forebrain. Nat. Neurosci. 2004, 7, 1233–1241. [Google Scholar] [CrossRef] [PubMed]
  27. Parras, C.M.; Hunt, C.; Sugimori, M.; Nakafuku, M.; Rowitch, D.; Guillemot, F. The proneural gene Mash1 specifies an early population of telencephalic oligodendrocytes. J. Neurosci. 2007, 27, 4233–4242. [Google Scholar] [CrossRef] [PubMed]
  28. Relucio, J.; Menezes, M.J.; Miyagoe-Suzuki, Y.; Takeda, S.; Colognato, H. Laminin regulates postnatal oligodendrocyte production by promoting oligodendrocyte progenitor survival in the subventricular zone. Glia 2012, 60, 1451–1467. [Google Scholar] [CrossRef] [PubMed]
  29. Whitman, M.C.; Greer, C.A. Synaptic integration of adult-generated olfactory bulb granule cells: Basal axodendritic centrifugal input precedes apical dendrodendritic local circuits. J. Neurosci. 2007, 27, 9951–9961. [Google Scholar] [CrossRef] [PubMed]
  30. Inoki, K.; Zhu, T.; Guan, K.L. TSC2 mediates cellular energy response to control cell growth and survival. Cell 2003, 115, 577–590. [Google Scholar] [CrossRef]
  31. Zoncu, R.; Efeyan, A.; Sabatini, D.M. mTOR: From growth signal integration to cancer, diabetes and ageing. Nat. Rev. Mol. Cell Biol. 2011, 12, 21–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Vander Haar, E.; Lee, S.I.; Bandhakavi, S.; Griffin, T.J.; Kim, D.H. Insulin signalling to mTOR mediated by the Akt/PKB substrate PRAS40. Nat. Cell Biol. 2007, 9, 316–323. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, L.; Harris, T.E.; Roth, R.A.; Lawrence, J.C. PRAS40 regulates mTORC1 kinase activity by functioning as a direct inhibitor of substrate binding. J. Biol. Chem. 2007, 282, 20036–20044. [Google Scholar] [CrossRef] [PubMed]
  34. Ma, L.; Chen, Z.; Erdjument-Bromage, H.; Tempst, P.; Pandolfi, P.P. Phosphorylation and functional inactivation of TSC2 by Erk: Implications for tuberous sclerosis and cancer pathogenesis. Cell 2005, 121, 179–193. [Google Scholar] [CrossRef] [PubMed]
  35. Roux, P.P.; Ballif, B.A.; Anjum, R.; Gygi, S.P.; Blenis, J. Tumor-promoting phorbol esters and activated Ras inactivate the tuberous sclerosis tumor suppressor complex via p90 ribosomal S6 kinase. Proc. Natl. Acad. Sci. USA 2004, 101, 13489–13494. [Google Scholar] [CrossRef] [PubMed]
  36. Carrière, A.; Cargnello, M.; Julien, L.A.; Gao, H.; Bonneil, É.; Thibault, P.; Roux, P.P. Oncogenic MAPK Signaling Stimulates mTORC1 Activity by Promoting RSK-Mediated Raptor Phosphorylation. Curr. Biol. 2008, 18, 1269–1277. [Google Scholar] [CrossRef] [PubMed]
  37. Kim, D.H.; Sarbassov, D.D.; Ali, S.M.; King, J.E.; Latek, R.R.; Erdjument-Bromage, H.; Tempst, P.; Sabatini, D.M. mTOR interacts with raptor to form a nutrient-sensitive complex that signals to the cell growth machinery. Cell 2002, 110, 163–175. [Google Scholar] [CrossRef]
  38. Kim, D.H.; Sarbassov, D.D.; Ali, S.M.; Latek, R.R.; Guntur, K.V.P.; Erdjument-Bromage, H.; Tempst, P.; Sabatini, D.M. GβL, a positive regulator of the rapamycin-sensitive pathway required for the nutrient-sensitive interaction between raptor and mTOR. Mol. Cell 2003, 11, 895–904. [Google Scholar] [CrossRef]
  39. Ma, X.M.; Blenis, J. Molecular mechanisms of mTOR-mediated translational control. Nat. Rev. Mol. Cell Biol. 2009, 10, 307–318. [Google Scholar] [CrossRef] [PubMed]
  40. Holz, M.K.; Ballif, B.A.; Gygi, S.P.; Blenis, J. mTOR and S6K1 mediate assembly of the translation preinitiation complex through dynamic protein interchange and ordered phosphorylation events. Cell 2005, 123, 569–580. [Google Scholar] [CrossRef] [PubMed]
  41. Crino, P.B. MTOR: A pathogenic signaling pathway in developmental brain malformations. Trends Mol. Med. 2011, 17, 734–742. [Google Scholar] [CrossRef] [PubMed]
  42. Feliciano, D.M.; Su, T.; Lopez, J.; Platel, J.C.; Bordey, A. Single-cell Tsc1 knockout during corticogenesis generates tuber-like lesions and reduces seizure threshold in mice. J. Clin. Invest. 2011, 121, 1596–1607. [Google Scholar] [CrossRef] [PubMed]
  43. Magri, L.; Cambiaghi, M.; Cominelli, M.; Alfaro-Cervello, C.; Cursi, M.; Pala, M.; Bulfone, A.; Garca-Verdugo, J.M.; Leocani, L.; Minicucci, F.; et al. Sustained activation of mTOR pathway in embryonic neural stem cells leads to development of tuberous sclerosis complex-associated lesions. Cell Stem Cell 2011, 9, 447–462. [Google Scholar] [CrossRef] [PubMed]
  44. Zordan, P.; Cominelli, M.; Cascino, F.; Tratta, E.; Poliani, P.L.; Galli, R. Tuberous sclerosis complex–associated CNS abnormalities depend on hyperactivation of mTORC1 and Akt. J. Clin. Invest. 2018, 128, 1688–1706. [Google Scholar] [CrossRef] [PubMed]
  45. Hsu, P.P.; Kang, S.A.; Rameseder, J.; Zhang, Y.; Ottina, K.A.; Lim, D.; Peterson, T.R.; Choi, Y.; Gray, N.S.; Yaffe, M.B.; et al. The mTOR-regulated phosphoproteome reveals a mechanism of mTORC1-mediated inhibition of growth factor signaling. Science 2011, 332, 1317–1322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Parkhitko, A.; Myachina, F.; Morrison, T.A.; Hindi, K.M.; Auricchio, N.; Karbowniczek, M.; Wu, J.J.; Finkel, T.; Kwiatkowski, D.J.; Yu, J.J.; et al. Tumorigenesis in tuberous sclerosis complex is autophagy and p62/sequestosome 1 (SQSTM1)-dependent. Proc. Natl. Acad. Sci. USA 2011, 108, 12455–12460. [Google Scholar] [CrossRef] [PubMed]
  47. Di Nardo, A.; Wentz, M.H.; Kwiatkowski, E.; Tsai, P.T.; Leech, J.D.; Greene-Colozzi, E.; Goto, J.; Dilsiz, P.; Talos, D.M.; Clish, C.B.; et al. Neuronal Tsc1/2 complex controls autophagy through AMPK-dependent regulation of ULK1. Hum. Mol. Genet. 2014, 23, 3865–3874. [Google Scholar] [CrossRef] [PubMed]
  48. Sarbassov, D.D.; Guertin, D.A.; Ali, S.M.; Sabatini, D.M. Phosphorylation and regulation of Akt/PKB by the rictor-mTOR complex. Science 2005, 307, 1098–1101. [Google Scholar] [CrossRef] [PubMed]
  49. Sciarretta, S.; Zhai, P.; Maejima, Y.; DelRe, D.P.; Nagarajan, N.; Yee, D.; Liu, T.; Magnuson, M.A.; Volpe, M.; Frati, G.; et al. MTORC2 regulates cardiac response to stress by inhibiting MST1. Cell Rep. 2015, 11, 125–136. [Google Scholar] [CrossRef] [PubMed]
  50. Jacinto, E.; Facchinetti, V.; Liu, D.; Soto, N.; Wei, S.; Jung, S.Y.; Huang, Q.; Qin, J.; Su, B. SIN1/MIP1 Maintains rictor-mTOR Complex Integrity and Regulates Akt Phosphorylation and Substrate Specificity. Cell 2006, 127, 125–137. [Google Scholar] [CrossRef] [PubMed]
  51. Liu, P.; Gan, W.; Chin, Y.R.; Ogura, K.; Guo, J.; Zhang, J.; Wang, B.; Blenis, J.; Cantley, L.C.; Toker, A.; et al. PtdIns(3,4,5)P3-Dependent Activation of the mTORC2 Kinase Complex. Cancer Discov. 2015, 5, 1194–1209. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, J.; Xu, K.; Liu, P.; Geng, Y.; Wang, B.; Gan, W.; Guo, J.; Wu, F.; Chin, Y.R.; Berrios, C.; et al. Inhibition of Rb Phosphorylation Leads to mTORC2-Mediated Activation of Akt. Mol. Cell 2016, 62, 929–942. [Google Scholar] [CrossRef] [PubMed]
  53. Gaubitz, C.; Oliveira, T.M.; Prouteau, M.; Leitner, A.; Karuppasamy, M.; Konstantinidou, G.; Rispal, D.; Eltschinger, S.; Robinson, G.C.; Thore, S.; et al. Molecular Basis of the Rapamycin Insensitivity of Target Of Rapamycin Complex 2. Mol. Cell 2015, 58, 977–988. [Google Scholar] [CrossRef] [PubMed]
  54. Rispal, D.; Eltschinger, S.; Stahl, M.; Vaga, S.; Bodenmiller, B.; Abraham, Y.; Filipuzzi, I.; Movva, N.R.; Aebersold, R.; Helliwell, S.B.; et al. Target of rapamycin complex 2 regulates actin polarization and endocytosis via multiple pathways. J. Biol. Chem. 2015, 290, 14963–14978. [Google Scholar] [CrossRef] [PubMed]
  55. Kwiatkowski, D.J. Tuberous sclerosis: From tubers to mTOR. Ann. Hum. Genet. 2003, 67, 87–96. [Google Scholar] [CrossRef] [PubMed]
  56. Yap, T.A.; Garrett, M.D.; Walton, M.I.; Raynaud, F.; de Bono, J.S.; Workman, P. Targeting the PI3K-AKT-mTOR pathway: Progress, pitfalls, and promises. Curr. Opin. Pharmacol. 2008, 8, 393–412. [Google Scholar] [CrossRef] [PubMed]
  57. O’Neill, C. PI3-kinase/Akt/mTOR signaling: Impaired on/off switches in aging, cognitive decline and Alzheimer’s disease. Exp. Gerontol. 2013, 48, 647–653. [Google Scholar] [CrossRef] [PubMed]
  58. Feng, Z. p53 regulation of the IGF-1/AKT/mTOR pathways and the endosomal compartment. Cold Spring Harb. Perspect. Biol. 2010, 2, a001057. [Google Scholar] [CrossRef] [PubMed]
  59. Diehl, J.A.; Cheng, M.; Roussel, M.F.; Sherr, C.J. Glycogen synthase kinase-3β regulates cyclin D1 proteolysis and subcellular localization. Genes Dev. 1998, 12, 3499–3511. [Google Scholar] [CrossRef] [PubMed]
  60. Fatrai, S.; Elghazi, L.; Balcazar, N.; Cras-Méneur, C.; Krits, I.; Kiyokawa, H.; Bernal-Mizrachi, E. Akt induces β-cell proliferation by regulating cyclin D1, cyclin D2, and p21 levels and cyclin-dependent kinase-4 activity. Diabetes 2006, 55, 318–325. [Google Scholar] [CrossRef] [PubMed]
  61. 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.; et al. An AKT3-FOXG1-reelin network underlies defective migration in human focal malformations of cortical development. Nat. Med. 2015, 21, 1445–1454. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, W.S.; Xu, P.Z.; Gottlob, K.; Chen, M.L.; Sokol, K.; Shiyanova, T.; Roninson, I.; Weng, W.; Suzuki, R.; Tobe, K.; et al. Growth retardation and increased apoptosis in mice with homozygous disruption of the akt1 gene. Genes Dev. 2001, 15, 2203–2208. [Google Scholar] [CrossRef] [PubMed]
  63. Tschopp, O.; Yang, Z.Z.; Brodbeck, D.; Dummler, B.A.; Hemmings-Mieszczak, M.; Watanabe, T.; Michaelis, T.; Frahm, J.; Hemmings, B.A. Essential role of protein kinase B gamma (PKB gamma/Akt3) in postnatal brain development but not in glucose homeostasis. Development 2005, 132, 2943–2954. [Google Scholar] [CrossRef] [PubMed]
  64. Murakami, M.; Ichisaka, T.; Maeda, M.; Oshiro, N.; Hara, K.; Edenhofer, F.; Kiyama, H.; Yonezawa, K.; Yamanaka, S. mTOR is essential for growth and proliferation in early mouse embryos and embryonic stem cells. Mol. Cell. Biol. 2004, 24, 6710–6718. [Google Scholar] [CrossRef] [PubMed]
  65. Ka, M.; Condorelli, G.; Woodgett, J.R.; Kim, W.Y. mTOR regulates brain morphogenesis by mediating GSK3 signaling. Development 2014, 141, 4076–4086. [Google Scholar] [CrossRef] [PubMed]
  66. Lacar, B.; Herman, P.; Hartman, N.W.; Hyder, F.; Bordey, A. S phase entry of neural progenitor cells correlates with increased blood flow in the young subventricular zone. PLoS ONE 2012, 7, e31960. [Google Scholar] [CrossRef] [PubMed]
  67. Pauklin, S.; Vallier, L. The cell-cycle state of stem cells determines cell fate propensity. Cell 2013, 155, 135–147. [Google Scholar] [CrossRef] [PubMed]
  68. Franke, T.F. PI3K/Akt: Getting it right matters. Oncogene 2008, 27, 6473–6488. [Google Scholar] [CrossRef] [PubMed]
  69. Rohde, J.; Heitman, J.; Cardenas, M.E. The TOR kinases link nutrient sensing to cell growth. J. Biol. Chem. 2001, 276, 9583–9586. [Google Scholar] [CrossRef] [PubMed]
  70. Ryskalin, L.; Lazzeri, G.; Flaibani, M.; Biagioni, F.; Gambardella, S.; Frati, A.; Fornai, F. mTOR-Dependent Cell Proliferation in the Brain. Biomed Res. Int. 2017, 2017. [Google Scholar] [CrossRef] [PubMed]
  71. 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.; et al. PI3K/AKT pathway mutations cause a spectrum of brain malformations from megalencephaly to focal cortical dysplasia. Brain 2015, 138, 1613–1628. [Google Scholar] [CrossRef] [PubMed]
  72. Li, Y.; Muffat, J.; Omer, A.; Bosch, I.; Lancaster, M.A.; Sur, M.; Gehrke, L.; Knoblich, J.A.; Jaenisch, R. Induction of Expansion and Folding in Human Cerebral Organoids. Cell Stem Cell 2017, 20, 385–396. [Google Scholar] [CrossRef] [PubMed]
  73. Lei, X.; Jiao, J. UTX Affects Neural Stem Cell Proliferation and Differentiation through PTEN Signaling. Stem Cell Rep. 2018, 10, 1193–1207. [Google Scholar] [CrossRef] [PubMed]
  74. Sinor, A.D. Akt-1 Expression Level Regulates CNS Precursors. J. Neurosci. 2004, 24, 8531–8541. [Google Scholar] [CrossRef] [PubMed]
  75. Kwon, C.H.; Zhu, X.; Zhang, J.; Baker, S.J. mTor is required for hypertrophy of Pten-deficient neuronal soma in vivo. Proc. Natl. Acad. Sci. USA 2003, 100, 12923–12928. [Google Scholar] [CrossRef] [PubMed]
  76. Zhou, J.; Blundell, J.; Ogawa, S.; Kwon, C.H.; Zhang, W.; Sinton, C.; Powell, C.M.; Parada, L.F. Pharmacological inhibition of mTORC1 suppresses anatomical, cellular, and behavioral abnormalities in neural-specific Pten knock-out mice. J. Neurosci. 2009, 29, 1773–1783. [Google Scholar] [CrossRef] [PubMed]
  77. Nikolaeva, I.; Kazdoba, T.M.; Crowell, B.; D’Arcangelo, G. Differential roles for Akt and mTORC1 in the hypertrophy of Pten mutant neurons, a cellular model of brain overgrowth disorders. Neuroscience 2017, 354, 196–207. [Google Scholar] [CrossRef] [PubMed]
  78. Wahane, S.D.; Hellbach, N.; Prentzell, M.T.; Weise, S.C.; Vezzali, R.; Kreutz, C.; Timmer, J.; Krieglstein, K.; Thedieck, K.; Vogel, T. PI3K-p110-alpha-subtype signalling mediates survival, proliferation and neurogenesis of cortical progenitor cells via activation of mTORC2. J. Neurochem. 2014, 130, 255–267. [Google Scholar] [CrossRef] [PubMed]
  79. Liu, P.; Begley, M.; Michowski, W.; Inuzuka, H. Cell-cycle-regulated activation of Akt kinase by phosphorylation at its carboxyl terminus. Nature 2014, 508, 541–545. [Google Scholar] [CrossRef] [PubMed]
  80. Zhou, Z.D.; Kumari, U.; Xiao, Z.C.; Tan, E.K. Notch as a molecular switch in neural stem cells. IUBMB Life 2010, 62, 618–623. [Google Scholar] [CrossRef] [PubMed]
  81. Perumalsamy, L.R.; Nagala, M.; Banerjee, P.; Sarin, A. A hierarchical cascade activated by non-canonical Notch signaling and the mTOR-Rictor complex regulates neglect-induced death in mammalian cells. Cell Death Differ. 2009, 16, 879–889. [Google Scholar] [CrossRef] [PubMed]
  82. Lee, K.S.; Wu, Z.; Song, Y.; Mitra, S.S.; Feroze, A.H.; Cheshier, S.H.; Lu, B. Roles of PINK1, mTORC2, and mitochondria in preserving brain tumor-forming stem cells in a noncanonical Notch signaling pathway. Genes Dev. 2013, 27, 2642–2647. [Google Scholar] [CrossRef] [PubMed]
  83. Hartman, N.W.; Lin, T.V.; Zhang, L.; Paquelet, G.E.; Feliciano, D.M.; Bordey, A. MTORC1 Targets the Translational Repressor 4E-BP2, but Not S6 Kinase 1/2, to Regulate Neural Stem Cell Self-Renewal In Vivo. Cell Rep. 2013, 5, 433–444. [Google Scholar] [CrossRef] [PubMed]
  84. Mahoney, C.; Feliciano, D.M.; Bordey, A.; Hartman, N.W. Switching on mTORC1 induces neurogenesis but not proliferation in neural stem cells of young mice. Neurosci. Lett. 2016, 614, 112–118. [Google Scholar] [CrossRef] [PubMed]
  85. Easley, C.A.; Ben-Yehudah, A.; Redinger, C.J.; Oliver, S.L.; Varum, S.T.; Eisinger, V.M.; Carlisle, D.L.; Donovan, P.J.; Schatten, G.P. mTOR-Mediated Activation of p70 S6K Induces Differentiation of Pluripotent Human Embryonic Stem Cells. Cell. Reprogr. 2010, 12, 263–273. [Google Scholar] [CrossRef] [PubMed]
  86. Agrawal, P.; Reynolds, J.; Chew, S.; Lamba, D.A.; Hughes, R.E. DEPTOR Is a Stemness Factor That Regulates Pluripotency of Embryonic Stem Cells. J. Biol. Chem. 2014, 289, 31818–31826. [Google Scholar] [CrossRef] [PubMed]
  87. Paliouras, G.N.; Hamilton, L.K.; Aumont, A.; Joppe, S.E.; Barnabe-Heider, F.; Fernandes, K.J.L. Mammalian Target of Rapamycin Signaling Is a Key Regulator of the Transit-Amplifying Progenitor Pool in the Adult and Aging Forebrain. J. Neurosci. 2012, 32, 15012–15026. [Google Scholar] [CrossRef] [PubMed]
  88. Han, J.; Wang, B.; Xiao, Z.; Gao, Y.; Zhao, Y.; Zhang, J.; Chen, B.; Wang, X.; Dai, J. Mammalian target of rapamycin (mTOR) is involved in the neuronal differentiation of neural progenitors induced by insulin. Mol. Cell. Neurosci. 2008, 39, 118–124. [Google Scholar] [CrossRef] [PubMed]
  89. Cloetta, D.; Thomanetz, V.; Baranek, C.; Lustenberger, R.M.; Lin, S.; Oliveri, F.; Atanasoski, S.; Ruegg, M. A. Inactivation of mTORC1 in the Developing Brain Causes Microcephaly and Affects Gliogenesis. J. Neurosci. 2013, 33, 7799–7810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Lin, T.V.; Hsieh, L.; Kimura, T.; Malone, T.J.; Bordey, A. Normalizing translation through 4E-BP prevents mTOR-driven cortical mislamination and ameliorates aberrant neuron integration. Proc. Natl. Acad. Sci. USA 2016, 113, 11330–11335. [Google Scholar] [CrossRef] [PubMed]
  91. Thomanetz, V.; Angliker, N.; Cloëtta, D.; Lustenberger, R.M.; Schweighauser, M.; Oliveri, F.; Suzuki, N.; Rüegg, M.A. Ablation of the mTORC2 component rictor in brain or Purkinje cells affects size and neuron morphology. J. Cell Biol. 2013, 201, 293–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Jackson, E.L.; Alvarez-Buylla, A. Characterization of adult neural stem cells and their relation to brain tumors. Cells. Tissues Organs 2008, 188, 212–224. [Google Scholar] [CrossRef] [PubMed]
  93. Kriegstein, A.; Alvarez-Buylla, A. The Glial Nature of Embryonic and Adult Neural Stem Cells. Annu. Rev. Neurosci. 2009, 32, 149–184. [Google Scholar] [CrossRef] [PubMed]
  94. Paridaen, J.T.; Huttner, W. B. Neurogenesis during development of the vertebrate central nervous system. EMBO Rep. 2014, 15, 351–364. [Google Scholar] [CrossRef] [PubMed]
  95. Pontious, A.; Kowalczyk, T.; Englund, C.; Hevner, R.F. Role of intermediate progenitor cells in cerebral cortex development. Dev. Neurosci. 2008, 30, 24–32. [Google Scholar] [CrossRef] [PubMed]
  96. Toyo-oka, K.; Wachi, T.; Hunt, R.F.; Baraban, S.C.; Taya, S.; Ramshaw, H.; Kaibuchi, K.; Schwarz, Q.P.; Lopez, A.F.; Wynshaw-Boris, A. 14-3-3ε and ζ Regulate Neurogenesis and Differentiation of Neuronal Progenitor Cells in the Developing Brain. J. Neurosci. 2014, 34, 12168–12181. [Google Scholar] [CrossRef] [PubMed]
  97. Díaz-Alonso, J.; Aguado, T.; de Salas-Quiroga, A.; Ortega, Z.; Guzmán, M.; Galve-Roperh, I. CB1 Cannabinoid receptor-dependent activation of mTORC1/Pax6 signaling drives Tbr2 expression and basal progenitor expansion in the developing mouse cortex. Cereb. Cortex 2015, 25, 2395–2408. [Google Scholar] [CrossRef] [PubMed]
  98. Hansen, D.V.; Lui, J.H.; Parker, P.R.L.; Kriegstein, A.R. Neurogenic radial glia in the outer subventricular zone of human neocortex. Nature 2010, 464, 554–561. [Google Scholar] [CrossRef] [PubMed]
  99. Nowakowski, T.J.; Pollen, A.A.; Sandoval-Espinosa, C.; Kriegstein, A.R. Transformation of the Radial Glia Scaffold Demarcates Two Stages of Human Cerebral Cortex Development. Neuron 2016, 91, 1219–1227. [Google Scholar] [CrossRef] [PubMed]
  100. Angevine, J.B.; Sidman, R.L. Autoradiographic study of cell migration during histogenesis of cerebral cortex in the mouse. Nature 1961, 192, 766–768. [Google Scholar] [CrossRef] [PubMed]
  101. Marín, O. Cellular and molecular mechanisms controlling the migration of neocortical interneurons. Eur. J. Neurosci. 2013, 38, 2019–2029. [Google Scholar] [CrossRef] [PubMed]
  102. Won, C.; Lin, Z.; Kumar, T.P.; Li, S.; Ding, L.; Elkhal, A.; Szabó, G.; Vasudevan, A. Autonomous vascular networks synchronize GABA neuron migration in the embryonic forebrain. Nat. Commun. 2013, 4, 2149. [Google Scholar] [CrossRef] [PubMed]
  103. 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.; et al. Association of MTOR Mutations With Developmental Brain Disorders, Including Megalencephaly, Focal Cortical Dysplasia, and Pigmentary Mosaicism. JAMA Neurol. 2016, 73, 836. [Google Scholar] [CrossRef] [PubMed]
  104. Magri, L.; Cominelli, M.; Cambiaghi, M.; Cursi, M.; Leocani, L.; Minicucci, F.; Poliani, P.L.; Galli, R. Timing of mTOR activation affects tuberous sclerosis complex neuropathology in mouse models. Dis. Model. Mech. 2013, 6, 1185–1197. [Google Scholar] [CrossRef] [PubMed]
  105. Ka, M.; Smith, A.L.; Kim, W.Y. MTOR controls genesis and autophagy of GABAergic interneurons during brain development. Autophagy 2017, 13, 1348–1363. [Google Scholar] [CrossRef] [PubMed]
  106. Kim, J.Y.; Duan, X.; Liu, C.Y.; Jang, M.H.; Guo, J.U.; Pow-anpongkul, N.; Kang, E.; Song, H.; Ming, G. DISC1 Regulates New Neuron Development in the Adult Brain via Modulation of AKT-mTOR Signaling through KIAA1212. Neuron 2009, 63, 761–773. [Google Scholar] [CrossRef] [PubMed]
  107. Steinecke, A.; Gampe, C.; Valkova, C.; Kaether, C.; Bolz, J. Disrupted-in-Schizophrenia 1 (DISC1) Is Necessary for the Correct Migration of Cortical Interneurons. J. Neurosci. 2012, 32, 738–745. [Google Scholar] [CrossRef] [PubMed]
  108. Steinecke, A.; Gampe, C.; Nitzsche, F.; Bolz, J. DISC1 knockdown impairs the tangential migration of cortical interneurons by affecting the actin cytoskeleton. Front. Cell. Neurosci. 2014, 8, 190. [Google Scholar] [CrossRef] [PubMed]
  109. Alvarez-Buylla, A.; Kohwi, M.; Nguyen, T.M.; Merkle, F.T. The heterogeneity of adult neural stem cells and the emerging complexity of their niche. Cold Spring Harb. Symp. Quant. Biol. 2008, 73, 357–365. [Google Scholar] [CrossRef] [PubMed]
  110. Feliciano, D.M.; Quon, J.L.; Su, T.; Taylor, M.M.; Bordey, A. Postnatal neurogenesis generates heterotopias, olfactory micronodules and cortical infiltration following single-cell Tsc1 deletion. Hum. Mol. Genet. 2012, 21, 799–810. [Google Scholar] [CrossRef] [PubMed]
  111. Lafourcade, C.A.; Lin, T.V.; Feliciano, D.M.; Zhang, L.; Hsieh, L.S.; Bordey, A. Rheb activation in subventricular zone progenitors leads to heterotopia, ectopic neuronal differentiation, and rapamycin-sensitive olfactory micronodules and dendrite hypertrophy of newborn neurons. J. Neurosci. 2013, 33, 2419–2431. [Google Scholar] [CrossRef] [PubMed]
  112. Grueber, W.B.; Sagasti, A. Self-avoidance and tiling: Mechanisms of dendrite and axon spacing. Cold Spring Harb. Perspect. Biol. 2010, 2, a001750. [Google Scholar] [CrossRef] [PubMed]
  113. Koleske, A.J. Molecular mechanisms of dendrite stability. Nat. Rev. Neurosci. 2013, 14, 536–550. [Google Scholar] [CrossRef] [PubMed]
  114. 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. PTEN deletion increases hippocampal granule cell excitability in male and female mice. Neurobiol. Dis. 2017, 108, 339–351. [Google Scholar] [CrossRef] [PubMed]
  115. Sokolov, A.M.; Seluzicki, C.M.; Morton, M.C.; Feliciano, D.M. Dendrite growth and the effect of ectopic Rheb expression on cortical neurons. Neurosci. Lett. 2018, 671, 140–147. [Google Scholar] [CrossRef] [PubMed]
  116. Whitman, M.C.; Greer, C.A. Adult neurogenesis and the olfactory system. Prog. Neurobiol. 2009, 89, 162–175. [Google Scholar] [CrossRef] [PubMed]
  117. Zhang, L.; Huang, T.; Bordey, A. Tsc1 haploinsufficiency is sufficient to increase dendritic patterning and Filamin A levels. Neurosci. Lett. 2016, 629, 15–18. [Google Scholar] [CrossRef] [PubMed]
  118. Zhang, L.; Bartley, C.M.; Gong, X.; Hsieh, L.S.; Lin, T.V.; Feliciano, D.M.; Bordey, A. MEK-ERK1/2-Dependent FLNA Overexpression Promotes Abnormal Dendritic Patterning in Tuberous Sclerosis Independent of mTOR. Neuron 2014, 84, 78–91. [Google Scholar] [CrossRef] [PubMed]
  119. Skalecka, A.; Liszewska, E.; Bilinski, R.; Gkogkas, C.; Khoutorsky, A.; Malik, A.R.; Sonenberg, N.; Jaworski, J. mTOR kinase is needed for the development and stabilization of dendritic arbors in newly born olfactory bulb neurons. Dev. Neurobiol. 2016, 76, 1308–1327. [Google Scholar] [CrossRef] [PubMed]
  120. Tavazoie, S.F.; Alvarez, V.A.; Ridenour, D.A.; Kwiatkowski, D.J.; Sabatini, B.L. Regulation of neuronal morphology and function by the tumor suppressors Tsc1 and Tsc2. Nat. Neurosci. 2005, 8, 1727–1734. [Google Scholar] [CrossRef] [PubMed]
  121. Yasuda, S.; Sugiura, H.; Katsurabayashi, S.; Shimada, T.; Tanaka, H.; Takasaki, K.; Iwasaki, K.; Kobayashi, T.; Hino, O.; Yamagata, K. Activation of Rheb, but not of mTORC1, impairs spine synapse morphogenesis in tuberous sclerosis complex. Sci. Rep. 2014, 4, 5155. [Google Scholar] [CrossRef] [PubMed]
  122. Bateup, H.S.; Takasaki, K.T.; Saulnier, J.L.; Denefrio, C.L.; Sabatini, B.L. Loss of Tsc1 In Vivo Impairs Hippocampal mGluR-LTD and Increases Excitatory Synaptic Function. J. Neurosci. 2011, 31, 8862–8869. [Google Scholar] [CrossRef] [PubMed]
  123. Meikle, L.; Pollizzi, K.; Egnor, A.; Kramvis, I.; Lane, H.; Sahin, M.; Kwiatkowski, D.J. 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. 2008, 28, 5422–5432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Tang, G.; Gudsnuk, K.; Kuo, S.H.; Cotrina, M.L.; Rosoklija, G.; Sosunov, A.; Sonders, M.S.; Kanter, E.; Castagna, C.; Yamamoto, A.; et al. Loss of mTOR-Dependent Macroautophagy Causes Autistic-like Synaptic Pruning Deficits. Neuron 2014, 83, 1131–1143. [Google Scholar] [CrossRef] [PubMed]
  125. Bhattacharya, A.; Kaphzan, H.; Alvarez-Dieppa, A.C.; Murphy, J.P.; Pierre, P.; Klann, E. Genetic Removal of p70 S6 Kinase 1 Corrects Molecular, Synaptic, and Behavioral Phenotypes in Fragile X Syndrome Mice. Neuron 2012, 76, 325–337. [Google Scholar] [CrossRef] [PubMed]
  126. Hernandez, D.; Torres, C.A.; Setlik, W.; Cebrián, C.; Mosharov, E.V.; Tang, G.; Cheng, H.C.; Kholodilov, N.; Yarygina, O.; Burke, R.E.; et al. Regulation of Presynaptic Neurotransmission by Macroautophagy. Neuron 2012, 74, 277–284. [Google Scholar] [CrossRef] [PubMed]
  127. Banko, J.L.; Merhav, M.; Stern, E.; Sonenberg, N.; Rosenblum, K.; Klann, E. Behavioral alterations in mice lacking the translation repressor 4E-BP2. Neurobiol. Learn. Mem. 2007, 87, 248–256. [Google Scholar] [CrossRef] [PubMed]
  128. Hou, L.; Klann, E. Activation of the Phosphoinositide 3-Kinase-Akt-Mammalian Target of Rapamycin Signaling Pathway Is Required for Metabotropic Glutamate Receptor-Dependent Long-Term Depression. J. Neurosci. 2004, 24, 6352–6361. [Google Scholar] [CrossRef] [PubMed]
  129. Henry, F.E.; Hockeimer, W.; Chen, A.; Mysore, S.P.; Sutton, M.A. Mechanistic target of rapamycin is necessary for changes in dendritic spine morphology associated with long-term potentiation. Mol. Brain 2017, 10, 50. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of the mammalian target of rapamycin (mTOR) signaling pathway. mTOR signaling is activated in several ways, including growth factors, amino acids, and increased cellular energy. Growth factors can activate both mTOR complexes or PI3K and ERK signaling. mTORC2 phosphorylates Akt at S473 to enhance activation. In turn, Akt suppress TSC2, which suppresses Rheb. Additionally, ERK can inhibit TSC2 to promote mTORC1 activation. Phosphorylation of 4E-BPs and S6K1/2 by mTORC1 promote cap-dependent translation.
Figure 1. Overview of the mammalian target of rapamycin (mTOR) signaling pathway. mTOR signaling is activated in several ways, including growth factors, amino acids, and increased cellular energy. Growth factors can activate both mTOR complexes or PI3K and ERK signaling. mTORC2 phosphorylates Akt at S473 to enhance activation. In turn, Akt suppress TSC2, which suppresses Rheb. Additionally, ERK can inhibit TSC2 to promote mTORC1 activation. Phosphorylation of 4E-BPs and S6K1/2 by mTORC1 promote cap-dependent translation.
Ijms 19 01544 g001
Figure 2. Effects of mTOR activation during neurogenesis. Neural stem cells (blue) undergo proliferation and either give rise to more stem cells (self-renewal) or daughter cells (green, differentiation). Activation of mTORC2 promotes neural stem cells (NSC) cell cycle entry through Akt. Hyperactivation of mTORC1 results in diminished self-renewal, favoring differentiation and lineage expansion. Daughter cells then migrate (red) from proliferation zones to their terminal positions. Activation of mTORC1 results in aberrant migration of daughter cells. Upon reaching their terminal positions, newly born neurons (gray) extend neurites and properly form dendritic arbors. Cells with high levels of mTORC1 activity can severely alter dendrite formation and synaptic integration. Upward pointing arrows indicate increased activity of designated genes or proteins. Downward pointing arrows indicate decreased activity or knockdown of designated genes or proteins.
Figure 2. Effects of mTOR activation during neurogenesis. Neural stem cells (blue) undergo proliferation and either give rise to more stem cells (self-renewal) or daughter cells (green, differentiation). Activation of mTORC2 promotes neural stem cells (NSC) cell cycle entry through Akt. Hyperactivation of mTORC1 results in diminished self-renewal, favoring differentiation and lineage expansion. Daughter cells then migrate (red) from proliferation zones to their terminal positions. Activation of mTORC1 results in aberrant migration of daughter cells. Upon reaching their terminal positions, newly born neurons (gray) extend neurites and properly form dendritic arbors. Cells with high levels of mTORC1 activity can severely alter dendrite formation and synaptic integration. Upward pointing arrows indicate increased activity of designated genes or proteins. Downward pointing arrows indicate decreased activity or knockdown of designated genes or proteins.
Ijms 19 01544 g002

Share and Cite

MDPI and ACS Style

LiCausi, F.; Hartman, N.W. Role of mTOR Complexes in Neurogenesis. Int. J. Mol. Sci. 2018, 19, 1544. https://doi.org/10.3390/ijms19051544

AMA Style

LiCausi F, Hartman NW. Role of mTOR Complexes in Neurogenesis. International Journal of Molecular Sciences. 2018; 19(5):1544. https://doi.org/10.3390/ijms19051544

Chicago/Turabian Style

LiCausi, Francesca, and Nathaniel W. Hartman. 2018. "Role of mTOR Complexes in Neurogenesis" International Journal of Molecular Sciences 19, no. 5: 1544. https://doi.org/10.3390/ijms19051544

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop