Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter February 17, 2017

Rheb in neuronal degeneration, regeneration, and connectivity

  • Veena Nambiar Potheraveedu , Miriam Schöpel , Raphael Stoll and Rolf Heumann EMAIL logo
From the journal Biological Chemistry

Abstract

The small GTPase Rheb was originally detected as an immediate early response protein whose expression was induced by NMDA-dependent synaptic activity in the brain. Rheb’s activity is highly regulated by its GTPase activating protein (GAP), the tuberous sclerosis complex protein, which stimulates the conversion from the active, GTP-loaded into the inactive, GDP-loaded conformation. Rheb has been established as an evolutionarily conserved molecular switch protein regulating cellular growth, cell volume, cell cycle, autophagy, and amino acid uptake. The subcellular localization of Rheb and its interacting proteins critically regulate its activity and function. In stem cells, constitutive activation of Rheb enhances differentiation at the expense of self-renewal partially explaining the adverse effects of deregulated Rheb in the mammalian brain. In the context of various cellular stress conditions such as oxidative stress, ER-stress, death factor signaling, and cellular aging, Rheb activation surprisingly enhances rather than prevents cellular degeneration. This review addresses cell type- and cell state-specific function(s) of Rheb and mainly focuses on neurons and their surrounding glial cells. Mechanisms will be discussed in the context of therapy that interferes with Rheb’s activity using the antibiotic rapamycin or low molecular weight compounds.

Introduction

Small Ras-related GTP-binding proteins constitute a large superfamily of proteins that play key roles in cell proliferation and differentiation (Borasio et al., 1989), cytoskeletal organization (Luo, 2002), and protein transport (Stenmark and Olkkonen, 2001) by acting as molecular switches in these processes. Ras homolog enriched in brain (Rheb), is a member of the Ras family of small GTP-binding proteins containing Rheb1 (called from here on Rheb) and Rheb2 (Yamagata et al., 1994; Gromov et al., 1995; Tee et al., 2005; Campbell et al., 2009). Rheb and Rheb2 proteins share 51% amino acid identity (Aspuria and Tamanoi, 2004).

Rheb was found to be expressed at high basal levels in hippocampus and cerebral cortex (Yamagata et al., 1994). Other tissues that express high levels of Rheb mRNA are lung, thymus, kidney, and intestine. Rheb is biochemically activated by growth factors such as epithelial growth factors, fibroblast growth factors and brain-derived neurotrophic factors (BDNF) and is associated with cellular growth, protein synthesis, and regeneration (Yamagata et al., 1994; Saucedo et al., 2003; Takei et al., 2004). In addition, Rheb mRNA and protein levels are upregulated as an immediate early response gene after toxic insults (Yamagata et al., 1994; Karassek et al., 2010).

The role of Rheb in neuronal growth, differentiation, aging, axonal regeneration as well as energy homeostasis of the system has gained some popularity (Swiech et al., 2008; Cheng et al., 2011a; Lafourcade et al., 2013; Yang et al., 2014). Interestingly, Rheb over-expression in neurons and in secondary cell lines shows that Rheb changes its function to become an enhancer of apoptosis depending on the cellular state (Wataya-Kaneda et al., 2001; Karassek et al., 2010; Cao et al., 2013; Ehrkamp et al., 2013). Here, we focus on the role of Rheb signaling pathways in neuronal growth, differentiation, regeneration, and connectivity, as opposed to the enhancement of apoptosis and progression of neurodegenerative diseases.

Rheb signaling pathway

Growth factors such as insulin-like growth factor 1 (IGF 1) and insulin activate the receptor tyrosine kinases or G-protein coupled receptors, which trigger the lipid kinase phosphatidylinositol-3 kinase (PI3K) (Gupta and Dey, 2012; Dibble and Cantley, 2015). The serine/threonine kinase Akt is one of the key downstream mediators of the PI3K signaling (Hay and Sonenberg, 2004; Long et al., 2005; Zoncu et al., 2011; Gupta and Dey, 2012; Dibble and Cantley, 2015). Akt links to the mTOR signaling via the tuberous sclerosis complex protein (TSC complex) and Rheb GTPase. The TSC complex consists of TSC1 (hamartin), TSC2 (tuberin) and TBC1D7 (Dibble et al., 2012) and inhibits the activity of the small GTPase Rheb by acting as a GTPase activating protein (GAP) towards Rheb (Gao and Pan, 2001; Potter et al., 2001; Tapon et al., 2001; Inoki et al., 2002; Inoki et al., 2003; Manning and Cantley, 2003; Tee et al., 2003; Saucedo et al., 2003; Zhang et al., 2003; Aspuria and Tamanoi, 2004; Hay and Sonenberg, 2004; Li et al., 2004a,b). In vitro, Akt phosphorylates TSC2 at conserved consensus phosphorylation sequences and down-regulates its GAP activity (Inoki et al., 2002, 2003). Reduced GAP activity of TSC complex allows for the accumulation of GTP-bound relative to GDP-bound Rheb (Inoki et al., 2002; Zhang et al., 2003; Li et al., 2004a,b; Tee et al., 2005). Rheb exhibits a low intrinsic GTPase activity in relation to other Ras-related small G proteins (Manning and Cantley, 2003; Schöpel et al., 2017). Hence, the GTP/GDP loading state of Rheb is highly regulated by its GAP, which is controlled by the presence of growth factors (Figure 1). Rheb binds to its effector ‘mammalian target of rapamycin’ (mTOR) (Long et al., 2005) and GTP-loaded Rheb is required for the activation of mTOR (Inoki et al., 2003; Long et al., 2005; Menon et al., 2014; Dibble and Cantley, 2015). A guanine nucleotide exchange factor (GEF) for Rheb has been identified but the biochemical and cellular relevance is still under discussion (Rehmann et al., 2008; Schöpel et al., 2017).

Figure 1: Rheb-mTORC1 signaling is activated in response to various intracellular and extracellular signals in the nervous system. Rheb promotes growth, regeneration and neuroprotection through the canonical mTORC1-S6K and many other non-canonical pathways.However, depending on the cellular state, cellular stress or under pathological conditions, Rheb become an enhancer of apoptosis. Canonical pathway: translation and cell cycle progression via S6K and 4E-BP1, inhibits autophagy via Ulk and ATG proteins. Non-canonical: mTORC2 regulates cytoskeleton required for axon guidance and synapse size; B-Raf regulates differentiation; Notch regulates cell fate in the external sensory organ (ESO), FKBP38 regulates apoptosis, Nix and LC3 induces mitophagy, RASSF1A enhances autophagy, dynein regulates aggresome formation, syntenin regulates spine morphogenesis, CAD regulates pyramidine synthesis, PERK inhibits protein synthesis, BACE1 in the pathogenesis of AD.
Figure 1:

Rheb-mTORC1 signaling is activated in response to various intracellular and extracellular signals in the nervous system. Rheb promotes growth, regeneration and neuroprotection through the canonical mTORC1-S6K and many other non-canonical pathways.

However, depending on the cellular state, cellular stress or under pathological conditions, Rheb become an enhancer of apoptosis. Canonical pathway: translation and cell cycle progression via S6K and 4E-BP1, inhibits autophagy via Ulk and ATG proteins. Non-canonical: mTORC2 regulates cytoskeleton required for axon guidance and synapse size; B-Raf regulates differentiation; Notch regulates cell fate in the external sensory organ (ESO), FKBP38 regulates apoptosis, Nix and LC3 induces mitophagy, RASSF1A enhances autophagy, dynein regulates aggresome formation, syntenin regulates spine morphogenesis, CAD regulates pyramidine synthesis, PERK inhibits protein synthesis, BACE1 in the pathogenesis of AD.

mTOR (frequently quoted as ‘mechanistic’ target of rapamycin), is an evolutionarily conserved checkpoint protein kinase that has emerged as a major effector of cell growth and proliferation. The mTOR pathway regulates protein synthesis, autophagy and metabolism in response to secreted growth factors, cellular levels of amino acids, glucose, energy levels, and oxygen levels (Manning and Cantley, 2003; Hay and Sonenberg, 2004; Adami et al., 2007; Swiech et al., 2008; Laplante and Sabatini, 2012). The mTOR forms two distinct complexes, mTORC1 and mTORC2, which are functionally different (Loewith et al., 2002). mTORC1 consists of mTOR, regulatory-associated protein of mTOR (Raptor); and mammalian lethal with SEC13 protein 8 (mLST8), along with two endogenous inhibitors of the complex, 40 kDa Proline-rich Akt substrate (PRAS40), and DEP domain-containing mTOR-interacting protein (DEPTOR) (Loewith et al., 2002; Hay and Sonenberg, 2004; Long et al., 2005; Yang et al., 2006; Adami et al., 2007). Downstream of activated mTORC1, protein translation is promoted by phosphorylation of 70 kDa ribosomal S6 kinase (S6K) and eukaryotic translation initiation factor 4E-binding protein 1 (4E-BP1) and autophagy is suppressed via Ulk1/Atg13 (Burnett et al., 1998; Hay and Sonenberg, 2004; Ganley et al., 2009; Hosokawa et al., 2009; Jung et al., 2009; Wang et al., 2009; Sciarretta et al., 2012) (see Figure 1). Rapamycin is an immunosuppressant and anticancer agent that binds to a 12-kDa FK506-binding protein (FKBP12) with high affinity. The FKBP12-rapamycin complex then acts on mTOR to inhibit mTORC1 function and thereby inhibit cell growth (Crespo et al., 2002; Loewith et al., 2002; Hay and Sonenberg, 2004; Adami et al., 2007; Gkogkas et al., 2010). mTORC2, in addition to mTOR and mLST8, contains the essential polypeptides: rapamycin-insensitive companion of mTOR (rictor) and mammalian stress-activated map kinase-interacting protein 1 (mSin1) (Loewith et al., 2002; Sarbassov et al., 2004; Adami et al., 2007; Avruch and Long, 2009). mTORC2 may signal to the actin cytoskeleton through Rho and Rac (Jacinto et al., 2004; Sarbassov et al., 2004).

Regulation of Rheb signaling pathway by subcellular localization

The localization of TSC complex, Rheb and mTORC1 at the lysosomes is critical for the regulation of mTORC1 and downstream pathway activation (Demetriades et al., 2014; Menon et al., 2014). Rheb is localized to multiple endo-membrane compartments including lysosomes due to the farnesylation of the C-terminal CAAX motif (Buerger et al., 2006). This localization of Rheb is not affected by insulin or amino acid levels (Inoki et al., 2002; Zhang et al., 2003; Li et al., 2004a,b; Tee et al., 2005; Menon et al., 2014). Although some studies suggest that Akt mediates the disassembly of the TSC complex, others suggest that Akt-mediated phosphorylation of TSC2 causes its degradation (Inoki et al., 2002; Menon et al., 2014). However, recently it was found that in the absence of insulin or growth factors, the TSC complex resides at the lysosomes in a Rheb-dependent manner and inactivates the Rheb at the lysosomes by acting as a Rheb-GAP. The TSC2 within the intact TSC complex has a strong binding preference for Rheb-GDP. In the presence of insulin, Akt-mediated phosphorylation of TSC complex causes its dissociation from the lysosomes, hence promoting Rheb and mTORC1 activation (Benjamin and Hall, 2014; Menon et al., 2014; Dibble and Cantley, 2015). Interestingly, mTORC1 can also be activated at other endo-membranes including peroxisomes and Golgi complex (Flinn et al., 2010; Zhang et al., 2013).

Cellular levels of amino acids also mediate the regulation of TSC complex dissociation and activation of mTORC1 at the lysosomes (Benjamin and Hall, 2014; Demetriades et al., 2014; Dibble and Cantley, 2015), yet in a different manner involving the Rag-GTPase (Sancak et al., 2008; Groenewoud and Zwartkruis, 2013). Rag-GTPase-mediated translocation of mTORC1 to the lysosomes, brings it into close proximity of Rheb, which stimulates the kinase activity of mTOR (Sancak et al., 2008; Groenewoud and Zwartkruis, 2013; Demetriades et al., 2014). Among the amino acids that control mTORC1 activity, arginine and leucine appear to be important (Carroll et al., 2016). Furthermore, Fawal et al., identified microspherule protein 1 (MCRS1) as an essential protein whose depletion allows the interaction of TSC complex and Rheb, hence inactivating Rheb and mTORC1 (Fawal et al., 2015).

Taken together, the signals from both amino acids and growth factors are integrated at the cytoplasmic face of the lysosomal membrane where mTORC1 is activated. Amino acids and growth factors are necessary and act independently, but neither alone is sufficient to fully activate mTORC1 (Benjamin and Hall, 2014; Menon et al., 2014; Dibble and Cantley, 2015). More recently, it was found that in the absence of amino acids, Rag GTPase can recruit the TSC complex to the lysosomes and maintains Rheb in the GDP-bound form, hence inactivating the Rheb-mTORC1 pathway (Demetriades et al., 2014). Therefore, the spatial control of the TSC complex at the lysosomes has been associated with both amino acid levels and the growth factor stimulation. When either of the two is missing, the TSC complex re-localizes to the lysosomes (Demetriades et al., 2016). In addition, the lysosomal re-localization of TSC complex is also an important stress response to inhibit mTORC1-mediated growth and the presence of any single stress maintains TSC complex at the lysosomes (Demetriades et al., 2016).

Rheb-mTORC1 is also able to sense the cellular redox state (Patel and Tamanoi, 2006; Di Nardo et al., 2009; Yoshida et al., 2011) which occurs independent of the amino acid levels and the Rag-regulator system. In addition to TSC complex, there are other upstream regulators for Rheb. Bcl-2/adenovirus E1B 19-kDa interacting protein 3 (Bnip3) is a hypoxia-induced pro-death protein, that binds with Rheb under hypoxia condition, and inhibits its function either by blocking Rheb interaction to downstream effectors or by interfering with GTP loading of Rheb (Li et al., 2007). Under conditions of low glucose levels, glyceraldehyde-3-phosphate dehydrogenase (GAPDH) interacts with Rheb, thereby preventing the association between Rheb and mTORC1 resulting in its inhibition (Lee et al., 2009).

Tuberous sclerosis complex

Tuberous sclerosis complex (TSC), is a rare genetic disease caused by mutation in the genes coding for TSC complex leading to the hyperactivation of the Rheb-mTORC1 pathway (Potter et al., 2001; Patel et al., 2003; Li et al., 2004a,b; Uhlmann et al., 2004; Crino et al., 2006). TSC is characterized by enhanced cell growth that affects organs such as brain, kidney, lung, heart and it leads to the development of hamartomas or benign focal malformations composed of tissue elements in a disorganized mass (Crino et al., 2006; Curatolo et al., 2008). The neurological symptoms of TSC include mental retardation, intractable epilepsy, and autism (Crino et al., 2006; Curatolo et al., 2008; Lim and Crino, 2013). Mouse models carrying mutations in the TSC1 or TSC2 genes have been extensively investigated to clarify the mechanisms of the TSC pathogenesis (Lim and Crino, 2013). Such studies emphasize that these TSC model mice exhibit impairments in neuronal morphology, dendritic spine formation, synaptic long-term depression and cognitive behaviors (Brown et al., 2012; Sugiura et al., 2015). Interestingly, in mouse models of TSC, there are both, mTORC1 dependent as well as mTORC1-independent Rheb pathways involved in the pathogenesis of TSC-associated abnormalities (Tavazoie et al., 2005; Yasuda et al., 2014; Sugiura et al., 2015)

Non-canonical Rheb pathway

Recently, several signaling pathways of TSC complex and Rheb have been described that function independently of mTORC1 (Jacinto et al., 2004; Karbowniczek et al., 2006; Zhou et al., 2009; Neuman and Henske, 2011). Hence, the activation of these non-canonical pathways cause some of the clinical manifestations of TSC that are resistant to mTORC1 inhibition by rapamycin (Yasuda et al., 2014). In order to cover a broader therapeutic spectrum, small molecules were investigated with Rheb as the primary binding target. Recent studies show that 4,4′-biphenol binds to Rheb and affects cell survival (Schöpel et al., 2013, 2017).

Some of the mTORC1-independent effects of Rheb are the suppression of aggresome formation (Zhou et al., 2009), the regulation of endocytic trafficking pathways (Saito et al., 2005) and determination of the cellular fate via notch signaling (Karbowniczek et al., 2010). The non-canonical mediators of Rheb signaling are as follows: B-Raf kinase (Im et al., 2002; Karbowniczek et al., 2004); FKBP38 (Ma et al., 2010); NIX, and LC3 (Melser et al., 2013); RASSF1A (Nelson and Clark, 2016); carbamoyl-phosphate synthetase 2, aspartate transcarbamoylase, and dihydroorotase (CAD) (Sato et al., 2015); β-site amyloid precursor protein (APP)-cleaving enzyme 1 (BACE1) (Shahani et al., 2014); protein kinase-like endoplasmic reticulum kinase (PERK) (Tyagi et al., 2015); and syntenin (Sugiura et al., 2015). Taken together, Rheb can regulate multiple cellular events both dependent and independent of mTORC1 and hence have roles in development, maintenance and degenerative processes. The functional significance of these interactions is summarized in Table 1.

Table 1:

An overview of various proteins identified to interact with Rheb.

ProteinsFunctional consequences
Proteins involved in canonical signaling pathway
 mTORRheb binding to mTOR is independent of its GTP-bound state but requires GTP-bound state for activation of mTORC1, thus promoting growth, cell cycle progression and inhibition of autophagy.
 TSC complexIn the absence of growth factors or insulin, the TSC complex stimulates the intrinsic GTPase activity of Rheb on the lysosomal surface and, through the conversion to Rheb-GDP, forms a complex with Rheb at the lysosomal membranes.
 Bnip3Under hypoxia condition, Bnip3 binds with Rheb and inhibits its function either by blocking Rheb interaction to downstream effectors or by interfering with Rheb GTP loading (Li et al., 2007).
 PLD1Rheb binds and activates phospholipase D1 (PLD1) in a GTP-dependent manner. Rheb indirectly activates mTORC1 by stimulating PLD1 to produce the lipid, phosphatidic acid (PA) (Sun et al., 2008).
 GAPDHIn the absence of glucose, GAPDH interacts with Rheb independent of the guanyl nucleotide loaded state of Rheb.
Proteins involved in non-canonical signaling pathway
 FKBP38FKBP38 is a member of the family of FK506-binding proteins that acts as an inhibitor of the mTOR (Bai et al., 2007). The inhibitory action of FKBP38 is antagonized by Rheb, which interacts with FKBP38 in a GTP-dependent manner and prevents its association with mTOR. Both Rheb and the anti-apoptotic protein Bcl-2 bind to the same region on FKBP38 and Rheb displaces Bcl-2 from FKBP38 facilitating the binding of Bcl-2 with pro-apoptotic proteins (Ma et al., 2010).
 RASSF1RASSF1 interacts with Rheb to promote autophagy by inhibiting mTORC1 signaling.
 NIXUpon high oxidative phosphorylation activity, Rheb is recruited to the mitochondrial outer membrane. Rheb promotes mitophagy through a physical interaction with the mitochondrial autophagic receptor Nix and the autophagosomal protein LC3-II.
 LC3Rheb physically interact with the mitochondrial autophagic receptor Nix and the autophagosomal protein LC3-II. The recruitment of Rheb to mitochondria leads to the activation of mitophagy.
 SynteninRheb-GDP binds syntenin to regulate spine morphogenesis.
 CADRheb binds with CAD at its C-terminal carbamoyl phosphate synthetase domain in a GTP dependent manner to regulate pyramidine synthesis. This interaction is dependent on the effector domain on Rheb.
 B-RafRheb directly interacts with, and inhibits B-Raf kinase. This direct interaction inhibits C-Raf/B-Raf heterodimerization, and Ras activation.
 PERKUnder ER stress conditions, Rheb associates with PERK in a GTP-dependent manner to inhibit protein synthesis through eIF2α signaling.
 BACE1Rheb binds to BACE1 and this interaction is necessary for BACE1 degradation in a GTP-dependent manner. BACE1 initiates amyloidogenic processing of APP to generate amyloid, which is a hallmark of Alzheimer’s disease (AD) pathology.

Pro-growth role of Rheb

Neuronal growth and differentiation

mTORC1 activity is important in controlling organism size and survival during embryonic development (Patel et al., 2003; Zoncu et al., 2011; Laplante and Sabatini, 2012). Specific deletion of the mTORC1 component Raptor, in the neuronal progenitors, leads to decreased brain size in the developing embryos due to impaired cell cycle progression and increased apoptosis (Cloëtta et al., 2013). Moreover, Rheb knockout mice are embryonically lethal (Goorden et al., 2011, 2015). Conditional Rheb knockout in adult mice leads to limited life span of the animals indicating altogether, that Rheb is essential for embryonic as well as adult animal survival (Goorden et al., 2011, 2015).

Studies using the Drosophila visual system describe that hyperactivation of the insulin receptor pathway in the visual system leads to the precocious acquisition of cell fate markers and premature neuronal differentiation (Bateman and McNeill, 2004). The Drosophila eye is composed of a pattern of eight photoreceptors (R1–R8) called ommatidia and has been used for detailed analysis of spatial aspects of differentiation (Carthew, 2007; Morante et al., 2007). Upon loss of TSC1, the photoreceptors 1, 6, and 7 and non-neuronal cone cells differentiate prematurely. These mutants however adopt normal cell fate, but tissue organization is disrupted (Bateman and McNeill, 2004). In the visual system of Drosophila, Rheb regulates differentiation which occurs independent of the regulation of cell cycle and cell size (Bateman and McNeill, 2004). Such studies indicate that activation of the insulin receptor-mTOR pathway is a critical control element for the timing of neuronal differentiation (Bateman and McNeill, 2004; Gu et al., 2014). The Notch signaling pathway plays a central role in the regulation of neuronal progenitor cell (NPC) differentiation and is required to maintain stem/progenitor cells in an undifferentiated state. It was shown that Rheb regulates cell fate in the external sensory organ (ESO) of Drosophila independent of mTORC1 via notch (Karbowniczek et al., 2010). Interestingly, the activation of Rheb leads to abnormalities in asymmetric cell division in sensory organ precursor cells of Drosophila (Karbowniczek et al., 2010).

Also in the mammalian system, insulin signaling plays an important role in age-dependent neuronal growth and differentiation (Cloëtta et al., 2013; Lafourcade et al., 2013). Constitutive activation of Rheb, in the NPCs and neuroblasts of subventricular zone in mouse may lead to premature neuronal differentiation and disruption in cell migration either though obstruction of the migratory path and/or release of aberrant cues (Lafourcade et al., 2013). In TSC neurons, neuronal migration defects were induced by upregulation of Cul5 expression by the hyperactivated mTORC1, leading to disruption of Reelin-Dab1 signaling (Moon et al., 2015).

Few studies using TSC knockout and Raptor knockout also show that both hyperactivation and inhibition of Rheb-mTORC1 pathway impair differentiation and can lead to premature death (Goorden et al., 2011; Angliker et al., 2015). The reduction of mTORC1 activity in specific regions of adult brain affects learning, memory and social behavior (Angliker et al., 2015; Goorden et al., 2015).

Rheb in synapse size and function

TSC complex, Rheb and mTORC1 are thought to play a key role in neuronal and synaptic plasticity and dentritic morphogenesis through regulation of protein synthesis (Kumar et al., 2005; Tavazoie et al., 2005; Swiech et al., 2008; Costa-Mattioli et al., 2009). mTORC1-mediated translation is required both at the synapse, for synaptic plasticity, and within the cell soma, for maintaining memory (Swiech et al., 2008; Gkogkas et al., 2010).

Drosophila neuromuscular junction (NMJ) was used to determine the role of TSC complex-Rheb-mTORC1 signaling on synapse assembly and function (Knox et al., 2007; Dimitroff et al., 2012; Natarajan et al., 2013). S6K regulated overall synapse growth (synapse bouton size, active zone number, and density) but not synaptic enlargement measured by the number of synaptic boutons per muscle area (Cheng et al., 2011a). However, Rheb over-expression in the motor neuron of the Drosophila larval stage or in NMJ did induce a doubling of the synapse size in terms of number of synaptic boutons per muscle area as well as an increase in synaptic function (Knox et al., 2007). Interestingly, rapamycin treatment reduced overall growth but did not suppress the Rheb-induced synaptic enlargement. This Rheb-induced synaptic enlargement was shown to depend upon bone morphogenetic protein (BMP) signaling mediated by wishful thinking, a BMP type II receptor (Knox et al., 2007). BMP pathways are important for normal NMJ growth in Drosophila. Animals bearing mutations in the wishful thinking gene show a very small NMJ with dramatically compromised synaptic function (Knox et al., 2007). The synaptic enlargement by over-expression of Rheb at the NMJ results in altering the balance between the number of synaptic boutons and the size of the underlying muscle which disrupts the normal growth and development (Dimitroff et al., 2012).

Recent findings show that Rheb over-expression mediated synaptic growth was morphologically and functionally different from that of the TSC mutant, indicating that TSC complex may be able to regulate NMJ synapse independent of Rheb-TORC1 pathway (Natarajan et al., 2013). TSC2 was known to regulate actin cytoskeleton through TORC2, which is critical for NMJ growth and function (Eaton et al., 2002; Luo 2002; Loewith et al., 2002; Coyle et al., 2004; Jacinto et al., 2004; Natarajan et al., 2013). Detailed studies on synaptic growth and enlargement were not yet reported in mice and more investigations in mammals into the role of Rheb and mTORC2 (Yang et al., 2006) would help to understand the regulation of brain connectivity.

Synaptic plasticity is measured as long-term potentiation (LTP) and long-term depression (LTD) in vertebrates and as long-term facilitation (LTF) in invertebrates (Gkogkas et al., 2010; Hoeffer and Klann, 2010). An interesting finding on the role of Rheb-mTORC1 canonical pathway in memory formation of the mollusk, Aplysia californica, was from Weatherill et al., where S6 activated by Rheb act as a critical regulator of 24 h-LTF (Weatherill et al., 2010). S6K- and 4E-BP1-mediated translation is required for synaptic plasticity and for cognitive processing in both vertebrates and invertebrates (Antion et al., 2008; Gkogkas et al., 2010; Weatherill et al., 2010).

Activation of Rheb signaling by loss of phosphatase and tensin homolog (PTEN), an important upstream negative regulator of TSC complex-Rheb-mTORC1 signaling, affects neuronal morphology and socialization behavior in mice (Kwon et al., 2006). In adult rat brains, Rheb can increase the levels of acetylcholine and total choline, which are important for cognitive functions (Jeon et al., 2015). Experiments using conditional mouse TSC1 or TSC2 knockout strains driven by synapsin promoters for neurons, emphasizes the role of these genes in epilepsy and altered cognition, behavioral phenotypes including seizures and altered social interactions (Lim and Crino, 2013).

However, in conditional adult Rheb knockout mice even though a reduction in the downstream effectors like S6K and 4E-BP1 phosphorylation was observed (Weatherill et al., 2010; Goorden et al., 2015), the hippocampal LTP, learning and memory was not effected to a larger extend (Goorden et al., 2015). This indicates that even a mild activation of the mTORC1 signaling can bring about a drastic effect on synaptic plasticity and learning associated with various diseases (Lim and Crino, 2013), while a mild inhibition of mTORC1 appears not to affect memory functions (Goorden et al., 2015). This is in accordance with the need to use high concentrations of rapamycin for impairing learning and plasticity in mice (Gkogkas et al., 2010; Goorden et al., 2015). The relative non-responsiveness in the synaptic plasticity by reduction of Rheb may also be due to the existence of Rheb2, which also activates mTORC1 (Tee et al., 2005).

These data further assess that even though the upregulation of mTORC1 pathway is highly sensitive and leads to cognitive defects as seen in TSC, the sustained suppression of mTORC1 using drugs like rapamycin may not affect the memory and learning significantly (Goorden et al., 2015).

Studies using Rictor knockout in brain show that, mTORC2 affects the size of the soma and dentrites, the number of primary dentrites in neurons as well as the synaptic connectivity of Purkinje cells in the cerebellum (Thomanetz et al., 2013; Angliker et al., 2015). The growth defects observed in Rictor knockout brain are caused by changes in the downstream effectors of mTORC2 and not by affecting mTORC1 signaling. Hence both mTORC1 and mTORC2 control distinct function in terms of social behaviour and motor coordination (Thomanetz et al., 2013).

Rheb in spine morphology and function

TSC complex-Rheb-mTORC1 pathway is known to regulate neuronal structures and function and the neurological symptoms of TSC are, at least in some part, due to cell-autonomous synapse function alterations. More recently, excitatory spine dysmorphogenesis and reduced spine synapse density, were observed in TSC neurons of rodent models (Yasuda et al., 2014; Sugiura et al., 2015). Rheb-syntenin signaling regulates activity dependent learning and memory processes independent of mTORC1 activation. Under normal conditions, Rheb-GDP associates with syntenin, leading to its proteasomal degradation, allowing normal spine formation required for learning and memory formation. However, in TSC mutant neurons, the Rheb is maintained in its GTP-bound form, which cannot bind to syntenin anymore, leading to accumulation of syntenin thereby disrupting the interaction between syndecan-2 and calcium/calmodulin- dependent serine protein kinase. The association of these proteins is crucial for normal spine development. Therefore, in TSC models, Rheb negatively regulates spine synapse formation causing dentritic spine abnormalities (Sugiura et al., 2015).

Another interesting finding was from Tavazoie et al., where under conditions of TSC1 knockout, increased soma and increased size of dendritic spines are observed (Tavazoie et al., 2005). The cytoskeleton of dendritic spines is particularly important in the reception of input from a single axon at the synapse. TSC complex may regulate soma and spine size via a mTOR dependent rapamycin-sensitive pathway and spine length via a mTOR independent rapamycin-insensitive pathway (Tavazoie et al., 2005). TSC complex regulates the cytoskeleton via ERM (ezrin, radixin, and moesin) proteins, neurofilament light chain and activation of Rho family of proteins (Haddad et al., 2002). Whether this regulation also involves mTORC2 and its regulation of cytoskeleton (Jacinto et al., 2004) still needs to be investigated. TSC complex also regulate cofilin and induces alteration in the spine length (Haddad et al., 2002; Luo 2002; Tavazoie et al., 2005).

Rheb in neuronal polarity and axon guidance

The axon and dendrites constitute the cellular basis for directional flow of information within the nervous system. Neuronal polarity is defined by the elaboration of a single axon and multiple dendrites, during brain development (Bradke and Dotti, 1999; Arimura and Kaibuchi, 2007). It has long been known that PI3K-Akt signaling plays a role in neuronal polarization in vitro (Arimura and Kaibuchi, 2007). Previous investigations show that TSC complex-Rheb-mTORC1 components are localized along growth cones of axons of cortical neurons in culture (Haddad et al., 2002; Choi et al., 2008). The presence of TSC complex and Rheb-mTORC1 pathway components as well as the activation of local translation of proteins that are required for axonal growth, together support the axon outgrowth in response to extracellular signals like Ephrins (Nie et al., 2010; Gracias et al., 2014). Rheb and mTORC1 control neuronal polarity by acting as upstream regulators of Rap1B GTPases. The restriction of Rap1B to a single neurite in turn ensures the extension of a single axon (Li et al., 2008).

Axon targeting has been studied extensively using the Drosophila visual system (Tayler and Garrity, 2003; Knox et al., 2007; Dimitroff et al., 2012). A study on the role of Rheb in the mushroom bodies of Drosophila shows that over-expression of Rheb in the mushroom bodies results in enlarged axonal lobes (Brown et al., 2012). Rheb over-expression produced cell autonomous defects in photoreceptor axon guidance in the Drosophila visual system, but did not produce motor-neuron axon misrouting and the motor-neuron axon follows the normal trajectory and synapses at the correct location on muscle (Knox et al., 2007; Dimitroff et al., 2012).

Regulation of actin is essential for axon guidance in the motor neuron and in the visual system and disruption of actin regulation may be the underlying cause of molecular deficits in axon guidance in the Drosophila visual system (Liebl et al., 2000; Luo 2002; Coyle et al., 2004). However, the axon guidance in photoreceptors may not be completely dependent on Rheb-mTORC1 growth signaling but also on TSC complex-mTORC2 regulation of actin cytoskeleton (Knox et al., 2007; Dimitroff et al., 2012).

Altogether, Rheb-mTORC1 and possibly mTORC2 are involved in the establishment of neuronal polarity in both, Drosophila and mammals. However, there may be differences in the detailed mechanisms involved.

Rheb in neuroprotection and axon regeneration

Several studies were conducted in analyzing the neuroprotective roles of Rheb by axon regeneration or increased survival in different neuron types (Jeon and Kim, 2015; Nam et al., 2015; Wu et al., 2016) (Figure 1). Interestingly, the fact that the adult mammalian central nervous system (CNS) is incapable of an axonal regenerative response after injury has been challenged by the recent work by Park et al. showing that constitutive activation of Akt-mTORC1 signaling in adult retinal ganglion neurons prior to optic nerve injury promoted regeneration of their axons (Park et al., 2008; Kim et al., 2012). Irrevocably, Akt and Rheb play an important role in protection and regeneration of axons not only during the period of acute axon injury but also 3 weeks post-lesion, by which time the degenerative process has run its course and ceased (Kim et al., 2011, 2012).

Brain-derived neurotrophic factor (BDNF) and glial cell-derived neurotrophic factor (GDNF) are required for regulating the development, maintenance function and plasticity of mature neurons and hence are important in protection of neurons in Parkinson’s diseases (PD), which is characterized by progressive degeneration of dopaminergic (DA) neurons (Kim et al., 2012; Jeon and Kim, 2015; Nam et al., 2015). Transduction of DA neurons with adeno-associated virus constructs containing a constitutively active form of Rheb (AAV-hRheb S16H), provided protection of DA neurons from 1-methyl-4-phenylpyridine (MPP) induced degeneration in adult brain. This neuroprotective effect is mediated by the induction of BDNF or GDNF expression and signaling through activation of mTORC1-CREB pathway (Nam et al., 2015).

Similar results were also observed in the case of mouse models of Alzheimer’s disease (AD) where hRheb (S16H) expression in the hippocampal neurons induced the sustained production of BDNF via mTORC1-dependent pathway, contributing to neuroprotection in the adult brain (Jeon et al., 2015).

Akt-Rheb-mTORC1 signaling has the ability to enhance many features of axon growth, including not only axon length, but also axon numbers per neuron and number of neurons having multiple axons (Choi et al., 2008). Rheb was found to promote activation of intrinsic sensory axon regrowth in dorsal root ganglia (DRG) across the dorsal root entry zone (DREZ) after inhibitory molecules were digested with chondroitinsulfate-ABC endolyase (ChABC) or in combination with Neurotrophin-3 (Liu et al., 2016; Wu et al., 2016). Both processes depend on the activation of Rheb-mTORC1 pathway supporting that developmental re-growth shares common molecular mechanism with regeneration following injury (Yaniv et al., 2012).

Even though inhibition of autophagy leads to cell death and neurodegeneration (Rubinsztein, 2006; Sarkar and Rubinsztein, 2008; Wang et al., 2009), excessive autophagy by inhibition of Rheb-mTORC1 pathway can result in axon degeneration following acute injury (Cheng et al., 2011b). Increased Akt-mTORC1 dependent macroautophagy induces retrograde axon degeneration in dopaminergic neurons after acute chemotoxic injury (Cheng et al., 2011b; Wang et al., 2012). Therefore, the ability of Akt or Rheb to protect axons from retrograde degeneration is likely to be due to its ability to suppress autophagy through mTORC1 signaling (Cheng et al., 2011b).

Altogether, the studies on the regulation of mTOR signaling on neural and behavioral development show that different inputs to mTOR signaling have distinct activities in the nervous system development, yet non-canonical mechanisms may have to be considered as well (Dimitroff et al., 2012; Yasuda et al., 2014; Sugiura et al., 2015).

Pro-apoptotic role of Rheb

Role in oxidative stress and aging

Cellular stress plays a significant role in the susceptibility, progress, and actual outcome of neuronal damage (Esch et al., 2002). Free radicals and reactive oxygen species (ROS) have been reported for their contribution to neuronal loss in cerebral ischemia, seizure disorders, schizophrenia, PD and AD (Floyd and Carney, 1992; Good et al., 1996; Christen, 2000; Barja, 2004; Hinerfeld et al., 2004). A growing body of evidences implicates oxidative stress as being involved in the propagation of cellular injury that leads to neuropathology in these various conditions (Good et al., 1996; Christen, 2000; Andersen, 2004).

Neurons are considered to be more susceptible to oxidative damage as compared to cells in other tissues (Barja, 2004; Gille et al., 2004; Hinerfeld et al., 2004). ROS can initiate a cascade of reactions that leads to neuronal cell death (Floyd and Carney, 1992; Berlett and Stadtmann, 1997; Uttara et al., 2009). It was found that TSC neurons have increased accumulation of ROS and have oxidative stress (Di Nardo et al., 2009). Over-expressing Rheb was ubiquitously found to sensitize adult Drosophila to oxidative stress (Patel and Tamanoi, 2006). However, selective Rheb over-expression in neurons and fat bodies did not sensitize the whole Drosophila to the oxidative stress response. Nevertheless, the failure of flies with increased Rheb-TOR-S6K signaling to directly elicit a stress response could result in the accumulation of oxidative damage and thus may serve as a model for age-related diseases. Early senescence of locomotor activity was observed in flies that over-express Rheb (Patel and Tamanoi, 2006).

The mTORC1 pathway has been discussed to be involved in cellular aging via its role in growth, stress response and protein metabolism, which affects the susceptibility of the organism to neuronal degeneration and death (Esch et al., 2002; Rubinsztein, 2006; Sarkar and Rubinsztein, 2008; Laplante and Sabatini, 2012; Nixon, 2013; Signer and Morrison, 2013). Hence inhibiting mTORC1-mediated translation results in protection and life span extension (Zoncu et al., 2011; Laplante and Sabatini, 2012; Signer and Morrison, 2013). In Rheb loss-of-function clones, the expression of two important cell fate markers, Prospero and Bar, is delayed, supporting that differentiation and senescence both are controlled by the TSC complex-Rheb-mTORC1 signaling pathway (Bateman and McNeill, 2004; Patel and Tamanoi, 2006).

Activation of TSC2 and repression of Rheb-mTORC1 by peroxisomal ROS induces autophagy which is an important response protecting from stress (Zhang et al., 2013). Hence, activation of Rheb-mTORC1 leads to a defective stress response and degeneration. Pathological protein aggregation and age-dependent neuronal vulnerability may act together to induce neurodegeneration (Yankner et al., 2008). Therefore early senescence caused by oxidative stress could be one of the pre-disposing factors for neurodegeneration (Floyd and Carney, 1992; Raina et al., 2001; Barja, 2004; Yankner et al., 2008; Uttara et al., 2009; Bao and Ohlemiller, 2010).

In summary, the effect of Rheb-mTORC1 pathway inhibition using rapamycin is different depending on the cellular state, which is changing during development and aging (Patel and Tamanoi, 2006; Lamming, 2016).

Response to other stresses

Rheb-mTORC1-S6K signaling also affects the response to various forms of stress such as UV, ER etc. (Patel and Tamanoi, 2006; Karassek et al., 2010). Karassek and colleagues have shown that Rheb plays a role in enhancing the apoptosis induced by UV, tumor necrosis factor α (TNF α) and tunicamycin in neuronal cells via apoptosis signal-regulating kinase 1 (ASK1), which is also activated under oxidative stress or ER stress (Ichijo 1997; Nishitoh et al., 2002; Karassek et al., 2010; Soga et al., 2012).

Prolonged inhibition of autophagy by Rheb-mTORC1 pathway induces protein stress (Zhou et al., 2009; Sciarretta et al., 2012). Studies have also shown that abrogated autophagy can induce apoptosis in response to misfolded protein stress (Zhou et al., 2009; Kang et al., 2011). In addition, Rheb inhibits aggresome formation independent of mTORC1 and plays a role in apoptosis induction (Wataya-Kaneda et al., 2001; Zhou et al., 2009). Furthermore, Ozcan et al. describes that the loss of the TSC1 and TSC2 genes causes ER stress and activates the unfolded protein responses (UPR) dependent on Rheb and mTORC1 (Ozcan et al., 2008). These cells are prone to apoptosis due to ER stress and UPR (Marciniak and Ron, 2006; Ozcan et al., 2008). Evidences generated from TSC brain lesions in vivo, also identified a role for the TSC complex in the neuronal stress response (Floyd and Carney, 1992; Berlett and Stadtmann, 1997; Di Nardo et al., 2009; Uttara et al., 2009). Di Nardo et al. demonstrated that lack of a functional TSC complex and activation of the downstream translational pathway of Rheb-mTORC1, resulted in increased vulnerability to ER stress and to neuronal damage (Di Nardo et al., 2009). In Drosophila, prolonged expression of Rheb in fly photoreceptors results in degeneration of rhabdomeres due to the loss of autophagy (Wang et al., 2009).

Although in general, Rheb-mTORC1 signaling promotes protein synthesis, recently, it was reported that Rheb inhibits protein synthesis by mediating phosphorylation of eukaryotic initiation factor 2α (eiF2α) via PERK independent of mTORC1 signaling (Tyagi et al., 2015). Rheb-PERK signaling is activated under ER stress leading to inhibition of protein synthesis, indicating that Rheb may function as a molecular switch between stimulation and inhibition of protein synthesis under physiological as well as pathological conditions (Tyagi et al., 2015).

Human age-related macular degeneration (AMD) is one of the major sight-threatening diseases where, irreversible neuron death including photoreceptor atrophy and retina ganglion cell (RGC) apoptosis are the major features. Light-induced retina degeneration in rodents has been widely used as models of AMD. Studies using rodent model of AMD supported that Rheb enhances the cell death induced by other stress signals (Shu et al., 2014).

In conclusion, cells failing to elicit a proper stress response are more prone to cell death as compared to ‘healthy cells’. In addition, activation of Rheb-mTORC1 under stress and during aging increase the organism’s susceptibility to neurodegeneration (Patel and Tamanoi, 2006; Ozcan et al., 2008; Di Nardo et al., 2009; Karassek et al., 2010; Kang et al., 2011; Lamming 2016).

Role in neurodegenerative diseases

Many lines of evidences suggest that intracellular protein degradation pathways such as autophagy and the ubiquitin-proteasome system are deregulated in neurodegenerative diseases and may play key roles in the etiology of these pathologies (Rubinsztein, 2006; Sarkar and Rubinsztein, 2008; Wang et al., 2009; Laplante and Sabatini, 2012; Nixon, 2013). The inhibiton of autophagy in TSC mutant or under Rheb over-expression leads to accumulation of toxic aggregate-prone proteins such as ataxin, Tau, and α-synuclein, which are linked to diseases such as spinocerebral ataxia, AD and PD, respectively (reviewed in Rubinsztein, 2006; Sarkar and Rubinsztein, 2008; Swiech et al., 2008). Therefore, inhibition of mTORC1 with rapamycin and small molecules enhances the autophagic degradation of aggregate-prone proteins in vitro and inhibition of protein synthesis in vivo reduces the severity of neurodegeneration (Sarkar and Rubinsztein, 2008).

Huntington’s disease (HD) is a neuro degenerative disease where the protein huntingtin (Htt) has an expanded polyglutamine (Poly-Q) tract and is characterized by early loss of medium spiny neurons in the striatum, which impairs motor and cognitive functions. In mouse models of HD, it was found that Htt might also interact with Rheb and cause abnormally high activation of mTORC1 that may contribute to disease progression through its roles in protein translation and autophagy (Pryor et al., 2014).

However, recently, it was found that mTORC1 activity is reduced in striatum tissues of HD patients and consequently mTORC1 activation by introducing the constitutively active form of Rheb could alleviate metabolic and degenerative phenotypes in HD brain (Lee et al., 2015). This was surprising since mTORC1 inhibits autophagy and it is possible that autophagy is regulated by other pathways, for example by the proteasomal degeradation, independent of mTORC1 (Sarkar and Rubinsztein, 2008). Protective effects may also result from the ability of mTORC1 to promote the lipogenic gene expression via sterol regulatory element-binding proteins (SREBPs) mediated transactivation (Zhang and Manning, 2015). There were also reports that Rhes, a striatum enriched protein, was able to improve neuropathological and motor phenotypes in HD mice (Lee et al., 2015). This effect of Rhes is mediated by promotion of autophagy independent of mTORC1 activation by binding to Beclin-1, a protein critical for the induction of autophagy (Mealer et al., 2014).

Moreover, Rheb levels were found to be downregulated in the AD brain, consistent with increased β-Site APP-cleaving enzyme 1 (BACE1) levels. BACE1 is the rate-limiting and principal enzyme responsible for Aβ generation in neurons of AD brain (Shahani et al., 2014). Rheb interacts with BACE1 in a GTP-dependent manner and promotes its degradation. Such data explain how the activity of Rheb could be linked to AD and HD specific processes.

Role in neurological deficits

TSC patients exhibit a spectrum of neurological deficits including autism, intellectual disability, and intractable epilepsy (Crino et al., 2006). The altered developmental effects in brain of TSC patients is reflective of Rheb over-expression (Lim and Crino, 2013). Rheb activation using constitutively active Rheb, in newborn NPCs and neuroblasts leads to TSC-like lesions, ectopic and premature neuronal differentiation and integration, micronodule formation, and hypertrophic neuronal morphogenesis (Lafourcade et al., 2013). Moreover some of the symptoms of TSC such as seizures are assumed to arise as a consequence of neural cell death and neurodegeneration (Wang et al., 2009). Impaired Rheb signaling may be associated with cognitive deficits seen in TSC, defective learning and memory processes. The altered cognition may be associated with the role of Rheb in both mTORC1-mediated long-lasting synaptic plasticity and memory as well as mTORC1 independent effects on spine and synapse development (Brown et al., 2012; Yasuda et al., 2014; Sugiura et al., 2015). Only a part of TSC complex-mediated changes in dendritic morphology of hippocampal neurons in organotypic cultures were suppressed by rapamycin treatment, indicating that such changes could also occur independent of mTORC1 activation (Knox et al., 2007; Nixon, 2013). Interestingly, increased phosphorylation of the human TSC2 has been found in the frontal cortex of AD and PD patients indicating that damage to neurons may accumulate with persistent Rheb-mTORC1 pathway upregulation (Brown et al., 2012). The activation of mTOR is associated with many diseases other than TSC, collectively termed ‘mTORopathies’ characterized by epilepsy, intractable seizures, a spectrum of developmental delay, altered cortical architecture and evidence of activated mTOR signaling (Lim and Crino, 2013). mTORC1 has also been found to be aberrantly activated in many benign tumors such as Lhermitte-Duclos disease as well as in malignant tumors such as glioblastoma multiforme and primary central nervous system lymphomas (PCNSL) (Nitta et al., 2016). Rheb has been found to be highly expressed in most but not all cases of PCNSL. Nevertheless, the downstream effectors S6K and 4E-BP1 were activated indicating that other upstream signaling pathways such as MAPK, AMPK, WNT, RalB signaling may be involved (Martin et al., 2015; Nitta et al., 2016).

Role of Rheb in neural stem cells

Stem cells persist throughout life in numerous mammalian tissues, allowing the replacement of cells during homeostatic turnover in healthy tissues and cells lost during disease. The embryonic development and differentiation of neural stem cells (NSCs) or NPCs requires mTORC1 activation by Rheb (Bateman and McNeill, 2004; Goorden et al., 2011; Hartman et al., 2013; Lafourcade et al., 2013). The deletion of Rheb in the NSCs leads to impaired postnatal growth and reduced body weight in mice and eventual death (Yang et al., 2014). During development, NPCs differentiate into neurons and glia and mTORC1 is crucial for glial differentiation via its ability to phosphorylate the transcription factor, Stat3 (Cloëtta et al., 2013).

Self-renewal and differentiation of the NSCs are important processes at different stages of brain development. In neonatal NSC, hyperactive mTORC1 induces NSC differentiation at the expense of self-renewal, while genetically decreasing mTORC1 activity prevents their differentiation, leading to reduced neuron production (Hartman et al., 2013).

In addition, constitutive activation of Rheb in the NPCs also affects neuroblast migration, probably due to its role in Reelin-Dab signaling (Moon et al., 2015). Taken together, Rheb is essential for the differentiation of NPCs, although its aberrant activation may cause abnormal differentiation.

Role of Rheb in glial cells

Astrocytes, as the major glial cell population in the brain, provide metabolic and trophic support for neurons, and modulate synaptic transmission and plasticity (Zhao and Flavin, 2000; Sofroniew and Vinters, 2010). Astrocytes are sensitive to infection and injury, and induce and amplify an immune reaction (Saijo et al., 2009; Sofroniew and Vinters, 2010). Astrocytes respond to tissue damage in CNS by becoming reactive (Ridet et al., 1997; Faulkner et al., 2004). This transformation into reactive astrocytes is promoted by upregulation of epidermal growth factor (EGF) receptor expression (Ridet et al., 1997; Faulkner et al., 2004; Codeluppi et al., 2009).

Codeluppi et al. have demonstrated that the Rheb-mTORC1 pathway is upregulated in reactive astrocytes of the injured spinal cord, whereas inhibition of mTORC1 by rapamycin can reverse activation of astrocytes (Codeluppi et al., 2009). Recent findings suggest that the cell size abnormalities seen in the brains from individuals affected with TSC may reflect the Rheb-S6K pathway regulation by the TSC complex in astrocytes (Uhlmann et al., 2004).

Role in glial cell regulated axon regeneration

Recent investigations have shown that the EGF activated Rheb-mTORC1 pathway inhibits the regeneration of damaged axons by enhancing astrocyte proliferation leading to glial scar formation by astrocytes (Codeluppi et al., 2009). This in turn suggest that even though rapamycin is an inhibitor of growth promoted by the Rheb-mTORC1, it could be used to control astrocytic responses in the damaged nervous system to promote a more permissive environment for axon regeneration (Codeluppi et al., 2009). However, the formation of a glial barrier around a lesion site is also an advantage, because it isolates the still intact CNS tissue from secondary lesions (Sofroniew and Vinters, 2010).

Nevertheless, in certain conditions, reactive astrocytes may provide a permissive substratum to support axonal regrowth. This is possible by the expression of several adhesion and synergistic factors by reactive astrocytes (Ridet et al., 1997). Whether the Rheb-mTORC1 pathway is involved in the expression of such molecules that enable astrocyte–neuron interactions in promoting axonal regeneration is yet to be investigated.

Myelin is a fatty substance that wraps around nerve fibers and serves to increase the speed of electrical communication between neurons. A dynamic turnover of myelin membranes is occurring throughout the lifespan and depends on the populations of adult myelinating oligodendrocyte and Schwann cells. Rheb-mediated mTORC1 activation is critical for differentiation of the oligodendrocyte precursor cells (OPCs) to mature oligodendrocytes. mTORC1-mediated activation of several transcription factors regulate the oligodendrocyte differentiation and the differentiation of the OPC is also likely through the control of cell-cycle exit by mTORC1 (Lebrun-Julien et al., 2014; Zou et al., 2014).

While the role of mTORC1 signaling for maintaining integrity of myelinated fibers is still under discussion (Lebrun-Julien et al., 2014; Norrmén et al., 2014; Zou et al., 2014), it has been shown that mTORC1 signaling in the Schwann cells is required for myelin formation by regulating lipid synthesis (Norrmén et al., 2014).

mTORC2-mediated signaling is a minor but important contributor too, as Rictor knockout led to hypomyelination. Therefore, a balanced and strictly regulated activation of mTORC1 and mTORC2 in oligodendrocytes is required for myelination and lipogenesis (Lebrun-Julien et al., 2014).

Role in glial cell regulated neuronal degeneration

Neuroinflammatory mechanisms involving microglial activation, astrogliosis, and lymphocyte infiltration probably also contribute to the cascade of events leading to neuronal degeneration (Carson et al., 2006; Hirsch and Hunot, 2009; More et al., 2013) in several neurodegenerative disorders, such as AD, PD, HD, amyotrophic lateral sclerosis (ALS) and progressive supranuclear palsy (Hirsch and Hunot, 2009; Wright et al., 2013). Exposure to lipopolysaccharides induces neuroinflammation in mice and dopaminergic neuronal cultures, which further leads to dopaminergic neuronal death and accumulation of insoluble cytoplasmic inclusions of α-synuclein in nigral neurons. Activated astrocytes express a broad array of neurotoxic molecules, including pro-inflammatory cytokines, chemokines, ROS and nitrogen species, which were found to play crucial roles in neuron inflammatory process and death (Hirsch and Hunot, 2009).

The Rheb-mTORC1 pathway enhances glial scar formation by astrocytes, and might play a role in regulating neuronal apoptosis in response to injury or stress (Codeluppi et al., 2009). Cao et al. investigated temporal-spatial patterns of Rheb expression after lipopolysaccharide (LPS) lateral ventral injection (Cao et al., 2013). It was found that Rheb was highly expressed and mostly located in neurons and reactive astrocytes in the brain cortex after LPS treatment. This Rheb expression in response to inflammation after LPS treatment induces neuronal apoptosis and therefore, Rheb may be associated with physiological and pathological process in neuroinflammation (Cao et al., 2013). Altogether, astrocytes can play an important role in neural cell damage by either inhibiting axon regeneration or by inducing neuronal apoptosis.

Conclusion

Rheb is emerging as an important component of the signaling pathway associated with various diseases such as TSC, tumors, metabolic diseases, and several neurodegenerative diseases. Studies on the role of Rheb and its associated upstream and downstream effectors in the neurons suggest that Rheb is involved in neuronal plasticity and differentiation, synapse growth and function and in the regeneration of axons as well as in neurodegeneration in response to injury and stress. However, corresponding in vivo experiments are still missing to validate the basic molecular mechanisms involved in the mammalian species. Some of the aspects of neuronal morphology such as synapse growth and axon guidance may be mediated via the mTORC2 pathway. More research on the Rheb-mTORC2 regulation may be needed, to establish the relationship between the two components and their role in actin cytoskeleton mediated regulation of synapse growth and axon guidance. Moreover, the activation of Rheb-mTORC1 pathway in astrocytes plays an important role in both neuroprotection and apoptosis evolving from inflammation as well as in inhibiting axon regeneration from damaged neurons. However, the switch in function of Rheb from a growth promoter to an enhancer of apoptosis in response to cellular and physiological stress indicates that Rheb could have a role in the pathogenesis of several neurodegenerative diseases and aging. In order to design drugs for treating diseases associated with Rheb signaling, future studies should take into consideration the cellular redox state, the cell type specific signaling in brain and in stem cells as well as in cellular senescence and aging.

Acknowledgments

This work was supported by a DFG grant (SFB 642, project A6). V.N.P and M.S thank the Ruhr University Research School Plus funded by Germany’s Excellence Initiative for financial support. R.H. was supported by the Horizon 2020 FET-open, Number – 686841, EU project ‘Magneuron’.

References

Adami, A., García-Alvarez, B., Arias-Palomo, E., Barford, D., and Llorca, O. (2007). Structure of TOR and its complex with KOG1. Mol. Cell 27, 509–516.10.1016/j.molcel.2007.05.040Search in Google Scholar PubMed

Andersen, J.K. (2004). Oxidative stress in neurodegeneration: cause or consequence? Nature Rev. Neurosci. 5, S18-S25.10.1038/nrn1434Search in Google Scholar PubMed

Angliker, N., Burri, M., Zaichuk, M., Fritschy, J.M., and Rüegg, M.A. (2015). mTORC1 and mTORC2 have largely distinct functions in Purkinje cells. Eur. J. Neurosci. 42, 2595–261210.1111/ejn.13051Search in Google Scholar PubMed

Antion, M.D., Hou, L., Wong, H., Hoeffer, C.A., and Klann, E. (2008). mGluR-dependent long-term depression is associated with increased phosphorylation of S6 and synthesis of elongation factor 1A but remains expressed in S6K-deficient mice. Mol. Cell Biol. 28, 2996–3007.10.1128/MCB.00201-08Search in Google Scholar PubMed PubMed Central

Arimura, N. and Kaibuchi, K. (2007). Neuronal polarity: from extracellular signals to intracellular mechanisms. Nat. Rev. Neurosci. 8, 194–205.10.1038/nrn2056Search in Google Scholar PubMed

Aspuria, P.J. and Tamanoi, F. (2004). The Rheb family of GTP-binding proteins. Cell Signal. 16, 1105–1112.10.1016/j.cellsig.2004.03.019Search in Google Scholar PubMed

Avruch, J. and Long, X. (2009). Amino acid regulation of TOR complex 1. Am. J. Physiol. Endocrinol. Metab. 296, E592–E602.10.1152/ajpendo.90645.2008Search in Google Scholar PubMed PubMed Central

Bai, X., Ma, D., Liu, A., Shen, X., Wang, Q.J., Liu, Y., and Jiang, Y. (2007). Rheb activates mTOR by Antagonizing its endogenous inhibitor, FKBP38. Science 318, 977–980.10.1126/science.1147379Search in Google Scholar PubMed

Bao, J. and Ohlemiller, K.K. (2010). Age-related loss of spiral ganglion neurons. Hear. Res. 264, 93–97.10.1016/j.heares.2009.10.009Search in Google Scholar PubMed PubMed Central

Barja, G. (2004). Aging in vertebrates, and the effect of caloric restriction: a mitochondrial free radical production-DNA damage mechanism? Biol. Rev. Camb. Philos. Soc. 79, 235–251.10.1017/S1464793103006213Search in Google Scholar

Bateman, J.M. and McNeill, H. (2004). Temporal control of differentiation by the insulin receptor/tor pathway in Drosophila. Cell 119, 87–96.10.1016/j.cell.2004.08.028Search in Google Scholar

Benjamin, D. and Hall, M.N. (2014). MTORC1: Turning off is just as important as turning on. Cell. 156. 627–628.10.1016/j.cell.2014.01.057Search in Google Scholar

Berlett, B.S. and Stadtman, E.R. (1997). Protein Oxidation in Aging, Disease, and Oxidative Stress. J. Biol. Chem. 272, 20313–20316.10.1074/jbc.272.33.20313Search in Google Scholar

Borasio, G.D., John, J., Wittinghofer, A., Barde, Y. A., Sendtner, M., and Heumann, R. (1989). RAS p21 protein promotes survival outgrowth of cultured embryonic and fiber neurons. Neuron 2, 1087–1096.10.1016/0896-6273(89)90233-XSearch in Google Scholar

Bradke, F. and Dotti, C.G. (1999). The role of local actin instability in axon formation. Science 283, 1931–1934.10.1126/science.283.5409.1931Search in Google Scholar PubMed

Brown, H.L., Kaun, K.R., and Edgar, B.A. (2012). The small GTPase Rheb affects central brain neuronal morphology and memory formation in Drosophila. PLoS One 7, e44888.10.1371/journal.pone.0044888Search in Google Scholar PubMed PubMed Central

Buerger, C., DeVries, B., and Stambolic, V. (2006). Localization of Rheb to the endomembrane is critical for its signaling function. Biochem. Biophys. Res. Commun. 344, 869–880.10.1016/j.bbrc.2006.03.220Search in Google Scholar PubMed

Burnett, P.E., Barrow, R.K., Cohen, N.A., Snyder, S.H., and Sabatini, D.M. (1998). RAFT1 phosphorylation of the translational regulators p70 S6 kinase and 4E-BP1. Proc. Natl. Acad. Sci. USA 95, 1432–1437.10.1073/pnas.95.4.1432Search in Google Scholar PubMed PubMed Central

Campbell, T.B., Basu, S., Hangoc, G., Tao, W., and Broxmeyer, H.E. (2009). Overexpression of Rheb2 enhances mouse hematopoietic progenitor cell growth while impairing stem cell repopulation. Blood 114, 3392–3401.10.1182/blood-2008-12-195214Search in Google Scholar PubMed PubMed Central

Cao, M., Tan, X., Jin, W., Zheng, H., Xu, W., Rui, Y., Li, L., Cao, J., Wu, X., Cui, G., et al. (2013). Upregulation of Ras homolog enriched in the brain (Rheb) in lipopolysaccharide-induced neuroinflammation. Neurochem. Int. 62, 406–417.10.1016/j.neuint.2013.01.025Search in Google Scholar PubMed

Carroll, B., Maetzel, D., Maddocks, O.D.K., Otten, G., Ratcliff, M., Smith, G.R., Dunlop, E.A., Passos, J.F., Davies, O.R., Jaenisch, R., et al. (2016). Control of TSC2-Rheb signaling axis by arginine regulates mTORC1 activity. eLife 5, 1–23.10.7554/eLife.11058Search in Google Scholar PubMed PubMed Central

Carson, M.J., Thrash, J.C., and Walter, B. (2006). The cellular response in neuroinflammation: The role of leukocytes, microglia and astrocytes in neuronal death and survival. Clin. Neurosci. Res. 6, 237–245.10.1016/j.cnr.2006.09.004Search in Google Scholar PubMed PubMed Central

Carthew, R.W. (2007). Pattern formation in the Drosophila eye. Curr. Opin. Genetics Dev. 17, 309–313.10.1016/j.gde.2007.05.001Search in Google Scholar PubMed PubMed Central

Cheng, C., Locke, C., and Davis, G.W. (2011a). S6 kinase localizes to the presynaptic active zone and functions with PDK1 to control synapse development. J. Cell. Biol. 194, 921–935.10.1083/jcb.201101042Search in Google Scholar PubMed PubMed Central

Cheng, H.C., Kim, S.R., Oo, T.F., Kareva, T., Yarygina, O., Rzhetskaya, M., Wang, C., During, M., Talloczy, Z., Tanaka, K., et al. (2011b). Akt suppresses retrograde degeneration of dopaminergic axons by inhibition of macroautophagy. J. Neurosci. 31, 2125–2135.10.1523/JNEUROSCI.5519-10.2011Search in Google Scholar PubMed PubMed Central

Choi, Y.J., Di Nardo, A., Kramvis, I., Meikle, L., Kwiatkowski, D.J., Sahin, M., and He, X. (2008). Tuberous sclerosis complex proteins control axon formation. Genes Dev. 22, 2485–2495.10.1101/gad.1685008Search in Google Scholar PubMed PubMed Central

Christen, Y. (2000). Oxidative stress and Alzheimer disease. Am. J. Clin. Nutr. 71, 621–629.10.1093/ajcn/71.2.621sSearch in Google Scholar PubMed

Cloëtta, D., Thomanetz, V., Baranek, C., Lustenberger, R.M., Lin, S., Oliveri, F., Atanasoski, S., and Rüegg, M.A. (2013). Inactivation of mTORC1 in the developing brain causes microcephaly and affects gliogenesis. J. Neurosci. 33, 7799–7810.10.1523/JNEUROSCI.3294-12.2013Search in Google Scholar PubMed PubMed Central

Codeluppi, S., Svensson, C.I., Hefferan, M.P., Valencia, F., Silldorff, M.D., Oshiro, M., Marsala, M., and Pasquale, E.B. (2009). The Rheb-mTOR pathway is upregulated in reactive astrocytes of the injured spinal cord. J. Neurosci. 29, 1093–1104.10.1523/JNEUROSCI.4103-08.2009Search in Google Scholar PubMed PubMed Central

Costa-Mattioli, M., Sossin, W.S., Klann, E., and Sonenberg, N. (2009). Translational Control of Long-Lasting Synaptic Plasticity and Memory. Neuron 61, 10–26.10.1016/j.neuron.2008.10.055Search in Google Scholar PubMed PubMed Central

Coyle, I.P., Koh, Y.H., Lee, W.C., Slind, J., Fergestad, T., Littleton, J.T., and Ganetzky, B. (2004). Nervous wreck, an SH3 adaptor protein that interacts with Wsp, regulates synaptic growth in Drosophila. Neuron 41, 521–534.10.1016/S0896-6273(04)00016-9Search in Google Scholar

Crespo, J.L., Hall, M.N., and Crespo, L. (2002). Elucidating TOR Signaling and Rapamycin Action: Lessons from Saccharomyces cerevisiae. Microbiol. Mol. Biol. Rev. 66, 579–591.10.1128/MMBR.66.4.579-591.2002Search in Google Scholar

Crino, P.B., Nathanson, K.L., and Henske, E.P. (2006). The tuberous sclerosis complex. New Engl. J. med. 355, 1345–1356.10.1201/9780849358364.ch22Search in Google Scholar

Curatolo, P., Bombardieri, R., and Jozwiak, S. (2008). Tuberous sclerosis. Lancet 37, 657–668.10.1016/S0140-6736(08)61279-9Search in Google Scholar

Demetriades, C., Doumpas, N., and Teleman, A.A. (2014). Regulation of TORC1 in response to amino acid starvation via lysosomal recruitment of TSC2. Cell 156, 786–799.10.1016/j.cell.2014.01.024Search in Google Scholar PubMed PubMed Central

Demetriades, C., Plescher, M., and Teleman, A.A. (2016). Lysosomal recruitment of TSC2 is a universal response to cellular stress. Nat. Commun. 7, 10662.10.1038/ncomms10662Search in Google Scholar PubMed PubMed Central

Dibble, C.C., Elis, W., Menon, S., Qin, W., Klekota, J., Asara, M.J., Finan, P.M., Kwiatkowski, D.J., Murphy, L.O., and Brendan, D. (2012). TBC1D7 is a third subunit of the TSC1-TSC2 complex upstream of mTORC1. Mol. Cell 47, 535–546.10.1016/j.molcel.2012.06.009Search in Google Scholar PubMed PubMed Central

Dibble, C.C. and Cantley, L.C. (2015). Regulation of mTORC1 by PI3K Signaling. Trends Cell Biol. 25, 545–55510.1016/j.tcb.2015.06.002Search in Google Scholar PubMed PubMed Central

Dimitroff, B., Howe, K., Watson, A., Campion, B., Lee, H.G., Zhao, N., O’Connor, M.B., Neufeld, T.P., and Selleck, S.B. (2012). Diet and energy-sensing inputs affect TorC1-mediated axon misrouting but not TorC2-directed synapse growth in a Drosophila model of tuberous sclerosis. PLoS One. 7, e30722.10.1371/journal.pone.0030722Search in Google Scholar PubMed PubMed Central

Di Nardo, A., Kramvis, I., Cho, N., Sadowski, A., Meikle, L., Kwiatkowski, D. J., and Sahin, M. (2009). Tuberous sclerosis complex activity is required to control neuronal stress responses in an mTOR-dependent manner. J. Neurosci. 29, 5926–5937.10.1523/JNEUROSCI.0778-09.2009Search in Google Scholar PubMed PubMed Central

Eaton, B.A., Fetter, R.D., and Davis, G.W. (2002). Dynactin is necessary for synapse stabilization. Neuron 34, 729–741.10.1016/S0896-6273(02)00721-3Search in Google Scholar

Ehrkamp, A., Herrmann, C., Stoll, R., and Heumann, R. (2013). Ras and Rheb Signaling in Survival and Cell Death. Cancers 5, 639–661.10.3390/cancers5020639Search in Google Scholar PubMed PubMed Central

Esch, T., Stefano, G.B., Fricchione, G.L., and Benson, H. (2002). The role of stress in neurodegenerative diseases and mental disorders. Neuroendocrinol. Lett. 23, 199–208.Search in Google Scholar

Faulkner, J.R., Herrmann, J.E., Woo, M.J., Tansey, K.E., Doan, N.B., and Sofroniew, M.V. (2004). Reactive astrocytes protect tissue and preserve function after spinal cord injury. J. Neurosci. 24, 2143–2155.10.1523/JNEUROSCI.3547-03.2004Search in Google Scholar PubMed PubMed Central

Fawal, M.A., Brandt, M., and Djouder, N. (2015). MCRS1 binds and couples rheb to amino acid-dependent mTORC1 activation. Dev. Cell 33, 67–82.10.1016/j.devcel.2015.02.010Search in Google Scholar PubMed

Flinn, R.J., Yan, Y., Goswami, S., Parker, P.J., and Backer, J.M. (2010). The late endosome is essential for mTORC1 signaling. Mol. Biol. Cell 21, 833–841.10.1091/mbc.e09-09-0756Search in Google Scholar PubMed PubMed Central

Floyd, R.A. and Carney, J.M. (1992). Free radical damage to protein and DNA: mechanisms involved and relevant observations on brain undergoing oxidative stress. Ann. Neurol. 32, 22–27.10.1002/ana.410320706Search in Google Scholar PubMed

Ganley, I.G., Lam, D.H., Wang, J., Ding, X., Chen, S., and Jiang, X. (2009). ULK1.ATG13. FIP200 complex mediates mTOR signaling and is essential for autophagy. J. Biol. Chem. 284, 12297–12305.10.1074/jbc.M900573200Search in Google Scholar PubMed PubMed Central

Gao, X. and Pan, D. (2001). TSC1 and TSC2 tumor suppressors antagonize insulin signaling in cell growth. Genes Dev. 15, 1383–1392.10.1101/gad.901101Search in Google Scholar PubMed PubMed Central

Gille, G., Hung, S.T., Reichmann, H., and Rausch, W.D. (2004). Oxidative stress to dopaminergic neurons as models of Parkinson’s disease. Ann. N.Y. Acad. Sci. 1018, 533–540.10.1196/annals.1296.066Search in Google Scholar PubMed

Gkogkas, C., Sonenberg, N., and Costa-Mattioli, M. (2010). Translational control mechanisms in long-lasting synaptic plasticity and memory. J. Biol. Chem. 285, 31913–31917.10.1074/jbc.R110.154476Search in Google Scholar PubMed PubMed Central

Good, P.F., Werner, P., Hsu, A., Olanow, C.W., and Perl, D.P. (1996). Evidence for Neuronal Oxidative Damage in Alzheimer’ s Disease. Am. J. Pathol. 149, 21–28.Search in Google Scholar

Goorden, S.M.I., Hoogeveen-Westerveld, M., Cheng, C., van Woerden, G.M., Mozaffari, M., Post, L., Duckers, H.J., Nellist, M., and Elgersma, Y. (2011). Rheb is essential for murine development. Mol. Cell. Biol. 31, 1672–167810.1128/MCB.00985-10Search in Google Scholar

Goorden, S.M.I., Abs, E., Bruinsma, C.F., Riemslagh, F.W., van Woerden, G.M., and Elgersma, Y. (2015). Intact neuronal function in Rheb1 mutant mice: implications for TORC1-based treatments. Hum. Mol. Gen. 24, 3390–3398.10.1093/hmg/ddv087Search in Google Scholar

Gracias, N.G., Shirkey-Son, N.J., and Hengst, U. (2014). Local translation of TC10 is required for membrane expansion during axon outgrowth. Nat. Commun. 67, 223–230.10.1038/ncomms4506Search in Google Scholar

Groenewoud, M.J. and Zwartkruis, F.J.T. (2013). Rheb and Rags come together at the lysosome to activate mTORC1. Biochem. Soc. Trans. 41, 951–955.10.1042/BST20130037Search in Google Scholar

Gromov, P.S., Madsen, P., Tomerup, N., and Celis, J.E. (1995). A novel approach for expression cloning of small GTPases: identification, tissue distribution and chromosome mapping of the human homolog of rheb. FEBS Lett. 377, 221–226.10.1016/0014-5793(95)01349-0Search in Google Scholar

Gu, T., Zhao, T., and Hewes, R.S. (2014). Insulin signaling regulates neurite growth during metamorphic neuronal remodeling. Biol. Open. 3, 81–93.10.1242/bio.20136437Search in Google Scholar PubMed PubMed Central

Gupta, A. and Dey, C.S. (2012). PTEN, a widely known negative regulator of insulin/PI3K signaling, positively regulates neuronal insulin resistance. Mol. Biol. Cell 23, 3882–3898.10.1091/mbc.e12-05-0337Search in Google Scholar

Haddad, L.A., Smith, N., Bowser, M., Niida, Y., Murthy, V., Gonzalez-Agosti, C., and Ramesh, V. (2002). The TSC1 tumor suppressor hamartin interacts with neurofilament-L and possibly functions as a novel integrator of the neuronal cytoskeleton. J. Biol. Chem. 277, 44180–44186.10.1074/jbc.M207211200Search in Google Scholar PubMed

Hartman, N.W., Lin, T.V., Zhang, L., Paquelet, G.E., Feliciano, D.M., and Bordey, A. (2013). mTORC1 targets the translational repressor 4E-BP2, but not S6 kinase 1/2, to regulate neural stem cell Self-Renewal in vivo. Cell Rep. 5, 433–444.10.1016/j.celrep.2013.09.017Search in Google Scholar PubMed

Hay, N. and Sonenberg, N. (2004). Upstream and downstream of mTOR. Genes Dev. 18, 1926–1945.10.1101/gad.1212704Search in Google Scholar PubMed

Hinerfeld, D., Traini, M.D., Weinberger, R.P., Cochran, B., Doctrow, S.R., Harry, J., and Melov, S. (2004). Endogenous mitochondrial oxidative stress: neurodegeneration, proteomic analysis, specific respiratory chain defects, and efficacious antioxidant therapy in superoxide dismutase 2 null mice. J. Neurochem. 88, 657–667.10.1046/j.1471-4159.2003.02195.xSearch in Google Scholar

Hirsch, E.C. and Hunot, S. (2009). Neuroinflammation in Parkinson’s disease: a target for neuroprotection? Lancet Neurol, 8, 382–397.10.1016/S1474-4422(09)70062-6Search in Google Scholar

Hoeffer, C.A. and Klann, E. (2010). mTOR signaling: At the crossroads of plasticity, memory and disease. Trends Neurosci. 33, 67–7510.1016/j.tins.2009.11.003Search in Google Scholar PubMed PubMed Central

Hosokawa, N., Hara, T., Kaizuka, T., Kishi, C., Takamura, A., Miura, Y., Iemura, S., Natsume, T., Takehana, K., Yamada, N., et al. (2009). Nutrient-dependent mTORC1 Association with the ULK1– Atg13– FIP200 Complex Required for Autophagy. Mol. Biol. Cell 20, 1981–1991.10.1091/mbc.e08-12-1248Search in Google Scholar PubMed PubMed Central

Ichijo, H. (1997). Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science 275, 90–94.10.1126/science.275.5296.90Search in Google Scholar PubMed

Im, E., von Lintig, F.C., Chen, J., Zhuang, S., Qui, W., Chowdhury, S., Worley, P.F., Boss, G.R., and Pilz, R.B. (2002). Rheb is in a high activation state and inhibits B-Raf kinase in mammalian cells. Oncogene 21, 6356–6365.10.1038/sj.onc.1205792Search in Google Scholar PubMed

Inoki, K., Li, Y., Zhu, T., Wu, J., and Guan, K. (2002). TSC2 is phosphorylated and inhibited by Akt and suppresses mTOR signalling. Nat. Cell Biol. 4, 648–657.10.1038/ncb839Search in Google Scholar PubMed

Inoki, K., Li, Y., Xu, T., and Guan, K.L. (2003). Rheb GTPase is a direct target of TSC2 GAP activity and regulates mTOR signaling. Genes Dev. 15, 1829–1834.10.1101/gad.1110003Search in Google Scholar PubMed PubMed Central

Jacinto, E., Loewith, R., Schmidt, A., Lin, S., Rüegg, M.A., Hall, A., and Hall, M.N. (2004). Mammalian TOR complex 2 controls the actin cytoskeleton and is rapamycin insensitive. Nat. Cell Biol. 6, 1122–1128.10.1038/ncb1183Search in Google Scholar PubMed

Jeon, M.T. and Kim, S.R. (2015). Roles of Rheb(S16H) in substantia nigra pars compacta dopaminergic neurons in vivo. Biomed. Rep. 3, 137–140.10.3892/br.2014.397Search in Google Scholar PubMed PubMed Central

Jeon, M.T., Nam, J.H., Shin, W.H., Leem, E., Jeong, K.H., Jung, U.J., Bae, Y.S., Jin, Y.H., Kholodilov, N., Burke, R.E., et al. (2015). In vivo AAV1 transduction with hRheb(S16H) protects hippocampal neurons by BDNF production. Mol. Ther. 23, 445–455.10.1038/mt.2014.241Search in Google Scholar PubMed PubMed Central

Jung, C.H., Jun, C.B., Ro, S., Kim, Y.M., Otto, N.M., Cao, J., Kundu, M., and Kim, D.H. (2009). ULK-Atg13-FIP200 complexes mediate mTOR signaling to the autophagy machinery. Mol. Biol. Cell 20, 1992–2003.10.1091/mbc.e08-12-1249Search in Google Scholar PubMed PubMed Central

Kang, Y.J., Lu, M.K., and Guan, K.L. (2011). The TSC1 and TSC2 tumor suppressors are required for proper ER stress response and protect cells from ER stress-induced apoptosis. Cell Death Differ. 18, 133–144.10.1038/cdd.2010.82Search in Google Scholar PubMed PubMed Central

Karassek, S., Berghaus, C., Schwarten, M., Goemans, C.G., Ohse, N., Kock, G., Jockers, K., Neumann, S., Gottfried, S., Herrmann, C., et al. (2010). Ras homolog enriched in brain (Rheb) enhances apoptotic signaling. J. Biol. Chem. 285, 33979–33991.10.1074/jbc.M109.095968Search in Google Scholar PubMed PubMed Central

Karbowniczek, M., Cash, T., Cheung, M., Robertson, G.P., Astrinidis, A., and Henske, E.P. (2004). Regulation of B-Raf kinase activity by tuberin and Rheb is mammalian target of rapamycin (mTOR)-independent. J. Biol. Chem. 279, 29930–29937.10.1074/jbc.M402591200Search in Google Scholar PubMed

Karbowniczek, M., Robertson, G.P., and Henske, E.P. (2006). Rheb inhibits C-raf activity and B-raf/C-raf heterodimerization. J. Biol. Chem. 281, 25447–25456.10.1074/jbc.M605273200Search in Google Scholar PubMed

Karbowniczek, M., Zitserman, D., Khabibullin, D., Hartman, T., Yu, J., Morrinson, T., Nicolas, E., Squillace, R., Roegiers, F., and Henske, E.P. (2010). The evolutionarily conserved TSC / Rheb pathway activates Notch in tuberous sclerosis complex and Drosophila external sensory organ development. J. Clin. Invest. 120, 93–102.10.1172/JCI40221Search in Google Scholar PubMed PubMed Central

Kim, S.R., Chen, X., Oo, T.F., Kareva, T., Yarygina, O., Wang, C., During, M., Kholodilov, N., and Burke, R.E. (2011). Dopaminergic pathway reconstruction by Akt/Rheb-induced axon regeneration. Ann. Neurol. 70, 110–120.10.1002/ana.22383Search in Google Scholar PubMed PubMed Central

Kim, S.R., Kareva, T., Yarygina, O., Kholodilov, N., and Burke, R.E. (2012). AAV transduction of dopamine neurons with constitutively active Rheb protects from neurodegeneration and mediates axon regrowth. Mol. Ther. 20, 275–286.10.1038/mt.2011.213Search in Google Scholar PubMed PubMed Central

Knox, S., Ge, H., Dimitroff, B.D., Ren, Y., Howe, K.A., Arsham, A.M., Easterday, M.C., Neufeld, T.P., O’Connor, M.B., and Selleck, S.B. (2007). Mechanisms of TSC-mediated control of synapse assembly and axon guidance. PLoS One 2, e375.10.1371/journal.pone.0000375Search in Google Scholar PubMed PubMed Central

Kumar, V., Zhang, M.X., Swank, M.W., Kunz, J., and Wu, G.Y. (2005). Regulation of dendritic morphogenesis by Ras-PI3K-Akt-mTOR and Ras-MAPK signaling pathways. J. Neurosci. 25, 11288–11299.10.1523/JNEUROSCI.2284-05.2005Search in Google Scholar PubMed PubMed Central

Kwon, C.H., Luikart, B.W., Powell, C.M., Zhou, J., Matheny, S.A., Zhang, W., Li, Y., Baker, S.J., and Parada, L.F. (2006). Pten regulates neuronal arborization and social interaction in mice. Neuron 50, 377–388.10.1016/j.neuron.2006.03.023Search in Google Scholar PubMed PubMed Central

Lafourcade, C.A., Lin, T.V., Feliciano, D.M., Zhang, L., Hsieh L.S., and 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. J. Neurosci. 33, 2419–2431.10.1523/JNEUROSCI.1840-12.2013Search in Google Scholar PubMed PubMed Central

Lamming, D.W. (2016). Inhibition of the mechanistic target of rapamycin (mTOR) - Rapamycin and beyond. Cold. Spring. Harb. Perspect. Med. 8, 583–592.10.1101/cshperspect.a025924Search in Google Scholar PubMed PubMed Central

Laplante, M. and Sabatini, D.M. (2012). MTOR signaling in growth control and disease. Cell 149, 274–293.10.1016/j.cell.2012.03.017Search in Google Scholar PubMed PubMed Central

Lebrun-Julien, F., Bachmann, L., Norrmén, C., Trötzmüller, M., Köfeler, H., Rüegg, M.A., Hall, M.N., and Suter, U. (2014). Balanced mTORC1 activity in oligodendrocytes is required for accurate CNS myelination. J. Neurosci. 34, 8432–8448.10.1523/JNEUROSCI.1105-14.2014Search in Google Scholar PubMed PubMed Central

Lee, M.N., Ha, S.H., Kim, J., Koh, A., Lee, C.S., Kim, J.H., Jeon, H., Kim, D.H., Suh, P.G., and Ryu, S.H. (2009). Glycolytic flux signals to mTOR through glyceraldehyde-3-phosphate dehydrogenase-mediated regulation of Rheb. Mol. Cell. Biol. 29, 3991–4001.10.1128/MCB.00165-09Search in Google Scholar PubMed PubMed Central

Lee, J.H., Tecedor, L., Chen, Y.H., Monteys, A.M., Sowada, M.J., Thompson, L.M., and Davidson, B.L. (2015). Reinstating aberrant mTORC1 activity in Huntington's disease mice improves disease phenotypes. Neuron 85, 303–315.10.1016/j.neuron.2014.12.019Search in Google Scholar PubMed PubMed Central

Li, Y., Inoki, K., and Guan, K.L. (2004a). Biochemical and Functional Characterizations of Small GTPase Rheb and TSC2 GAP Activity. Mol. Cell. Biol. 24, 7965–7975.10.1128/MCB.24.18.7965-7975.2004Search in Google Scholar PubMed PubMed Central

Li, Y., Corradetti, M.N., Inoki, K., and Guan, K.L. (2004b). TSC2: filling the GAP in the mTOR signaling pathway. Trends Biochem. Sci. 29, 32–38.10.1016/j.tibs.2003.11.007Search in Google Scholar PubMed

Li, Y., Wang, Y., Kim, E., Beemiller, P., Wang, C.Y., Swanson, J., You, M., and Guan, K.L. (2007). Bnip3 mediates the hypoxia-induced inhibition on mammalian target of rapamycin by interacting with Rheb. J. Biol. Chem. 282, 35803–35813.10.1074/jbc.M705231200Search in Google Scholar

Li, Y.H., Werner, H., and Püschel, A.W. (2008). Rheb and mTOR regulate neuronal polarity through Rap1B. J. Biol. Chem. 283, 33784–33792.10.1074/jbc.M802431200Search in Google Scholar

Liebl, E.C., Forsthoefel, D.J., Franco, L.S., Sample, S.H., Hess, J.E., Cowger, J.A., Chandler, M.P., Shupert, A.M., and Seeger, M.A. (2000). Dosage-sensitive, reciprocal genetic interactions between the Abl tyrosine kinase and the putative GEF trio reveal trio ’ s role in axon pathfinding. Neuron 26, 107–118.10.1016/S0896-6273(00)81142-3Search in Google Scholar

Lim, K.C. and Crino, P.B. (2013). Focal malformations of cortical development: new vistas for molecular pathogenesis. Neuroscience 252, 262–276.10.1016/j.neuroscience.2013.07.037Search in Google Scholar

Liu, Y., Kelamangalath, L., Kim, H., Han, S.B., Tang, X., Zhai, J., Hong, J.W., Lin, S., Son, Y.J., and Smith, G.M. (2016). NT-3 promotes proprioceptive axon regeneration when combined with activation of the mTor intrinsic growth pathway but not with reduction of myelin extrinsic inhibitors. Exp. Neurol. 283, 73–84.10.1016/j.expneurol.2016.05.021Search in Google Scholar

Loewith, R., Jacinto, E., Wullschleger, S., Lorberg, A., Crespo, J.L., Bonenfant, D., Oppliger, W., Jenoe, P., and Hall, M.N. (2002). Two TOR complexes, only one of which is rapamycin sensitive, have distinct roles in cell growth control. Mol. Cell 10, 457–468.10.1016/S1097-2765(02)00636-6Search in Google Scholar

Long, X., Lin, Y., Ortiz-Vega, S., Yonezawa, K., and Avruch, J. (2005). Rheb binds and regulates the mTOR kinase. Curr. Biol. 15, 702–713.10.1016/j.cub.2005.02.053Search in Google Scholar PubMed

Luo, L. (2002). Actin cytoskeleton regulation in neuronal morphogenesis and structural plasticity. Annu. Rev. Cell. Dev. Biol. 18, 601–635.10.1146/annurev.cellbio.18.031802.150501Search in Google Scholar PubMed

Ma, D., Bai, X., Zou, H., Lai, Y., and Jiang, Y. (2010). Rheb GTPase controls apoptosis by regulating interaction of FKBP38 with Bcl-2 and Bcl-XL. J. Biol. Chem. 285, 8621–8627.10.1074/jbc.M109.092353Search in Google Scholar PubMed PubMed Central

Manning, B.D. and Cantley, L.C. (2003). Rheb fills a GAP between TSC and TOR. Trends Biochem. Sci. 28, 573–576.10.1016/j.tibs.2003.09.003Search in Google Scholar PubMed

Marciniak, S.J. and Ron, D. (2006). Endoplasmic reticulum stress signaling in disease. Physiol. Rev. 86, 1133–1149.10.1152/physrev.00015.2006Search in Google Scholar PubMed

Martin, T.D., Chen, X.W., Kaplan, R.E., Saltiel, A.R., Walker, C.L., Reiner, D.J., and Der, C.J. (2015). Ral and Rheb GTPase activating proteins integrate mTOR and GTPase signaling in aging, autophagy, and tumor cell invasion. Mol. Cell 53, 209–22010.1016/j.molcel.2013.12.004Search in Google Scholar PubMed PubMed Central

Mealer, R.G., Murray, A.J., Shahani, N., Subramaniam, S., and Snyder, S.H. (2014). Rhes, a striatal-selective protein implicated in Huntington disease, binds Beclin-1 and activates autophagy. J. Biol. Chem. 289, 3547–3554.10.1074/jbc.M113.536912Search in Google Scholar PubMed PubMed Central

Melser, S., Chatelain, E.H., Lavie, J., Mahfouf, W., Jose, C., Obre, E., Goorden, S., Priault, M., Elgersma, Y., Rezvani, H.R., et al. (2013). Rheb regulates mitophagy induced by mitochondrial energetic status. Cell Metab. 17, 719–730.10.1016/j.cmet.2013.03.014Search in Google Scholar PubMed

Menon, S., Dibble, C.C., Talbott, G., Hoxhaj, G., Valvezan, A.J., Takahashi, H., Cantley, L.C., and Manning, B.D. (2014). Spatial control of the TSC complex integrates insulin and nutrient regulation of mtorc1 at the lysosome. Cell 156, 771–785.10.1016/j.cell.2013.11.049Search in Google Scholar PubMed PubMed Central

Moon, U.Y., Park, J.Y., Park, R, Cho, J.Y., Hughes, L.J., McKenna, J. iiird., Goetzl, L., Cho, S.H., Crino, P.B., Gambello, M.J., et al. (2015). Impaired Reelin-Dab1 signaling contributes to neuronal migration deficits of tuberous sclerosis complex. Cell Rep. 12, 965–978.10.1016/j.celrep.2015.07.013Search in Google Scholar PubMed PubMed Central

Morante, J., Desplan, C., and Celik, A. (2007). Generating patterned arrays of photoreceptors. Curr. Opin. Genet. Dev. 17, 314–319.10.1016/j.gde.2007.05.003Search in Google Scholar PubMed PubMed Central

More, S.V., Kumar, H., Kim, I.S., Song, S.Y., and Choi, D.K. (2013). Cellular and molecular mediators of neuroinflammation in the pathogenesis of Parkinson’s disease. Mediat. Inflamm. 2013, 952375.10.1155/2013/952375Search in Google Scholar PubMed PubMed Central

Nam, J.H., Leem, E., Jeon, M.T., Jeong, K.H., Park, J.W., Jung, U.J., Kholodilov, N., Burke, R.E., Jin, B.K., and Kim, S.R. (2015). Induction of GDNF and BDNF by hRheb(S16H) transduction of SNpc neurons: neuroprotective mechanisms of hRheb(S16H) in a model of Parkinson's disease. Mol. Neurobiol. 51, 487–499.10.1007/s12035-014-8729-2Search in Google Scholar PubMed

Natarajan, R., Trivedi-vyas, D., and Wairkar, Y.P. (2013). Tuberous sclerosis complex regulates drosophila neuromuscular junction growth via the TORC2/Akt pathway. Hum. Mol. Genet. 22, 2010–202310.1093/hmg/ddt053Search in Google Scholar PubMed

Nelson, N. and Clark, G.J. (2016). Rheb may complex with RASSF1A to coordinate Hippo and TOR signaling. Oncotarget 7, 33821–3383110.18632/oncotarget.8447Search in Google Scholar PubMed PubMed Central

Neuman, N.A. and Henske, E.P. (2011). Non-canonical functions of the tuberous sclerosis complex-Rheb signalling axis. EMBO Mol. Med. 3, 189–200.10.1002/emmm.201100131Search in Google Scholar PubMed PubMed Central

Nie, D., Di Nardo, A., Han, J.M., Baharanyi, H., Kramvis, I., Huynh, T., Dabora, S., Codeluppi, S., Pandolfi, P.P., Pasquale, E.B., et al. (2010). Tsc2-Rheb signaling regulates EphA-mediated axon guidance. Nat. Neurosci. 13, 163–172.10.1038/nn.2477Search in Google Scholar PubMed PubMed Central

Nishitoh, H., Matsuzawa, A., Tobiume, K., Saegusa, K., Takeda, K., Inoue, K., Hori, S., Kakizuka, A., and Ichijo, H. (2002). ASK1 is essential for endoplasmic reticulum stress-induced neuronal cell death triggered by expanded polyglutamine repeats. Genes Dev. 16, 1345–1355.10.1101/gad.992302Search in Google Scholar PubMed PubMed Central

Nitta, N., Nakasu, S., Shima, A., and Nozaki, K. (2016). mTORC1 signaling in primary central nervous system lymphoma. Surg. Neurol. Int. 7, S475-S480.10.4103/2152-7806.185781Search in Google Scholar PubMed PubMed Central

Nixon, R. A. (2013). The role of autophagy in neurodegenerative disease. Nat. Med. 19, 983–997.10.1038/nm.3232Search in Google Scholar PubMed

Norrmén, C., Figlia, G., Lebrun-Julien, F., Pereira, J.A., Trötzmüller, M., Köfeler, H.C., Rantanen, V., Wessig, C., van Deijk, A.L., Smit, A.B., et al. (2014). mTORC1 controls PNS myelination along the mTORC1-RXRγ-SREBP-lipid biosynthesis axis in Schwann cells. Cell Rep. 9, 646-66010.1016/j.celrep.2014.09.001Search in Google Scholar PubMed

Ozcan, U., Ozcan, L., Yilmaz, E., Düvel, K., Sahin, M., Manning, B.D., and Hotamisligil, G.S. (2008). Loss of the tuberous sclerosis complex tumor suppressors triggers the unfolded protein response to regulate insulin signaling and apoptosis. Mol. cell. 29, 541–551.10.1016/j.molcel.2007.12.023Search in Google Scholar PubMed PubMed Central

Park, K.K., Liu, K., Hu, Y., Smith, P.D., Wang, C., Cai, B., Xu, B., Connolly, L., Kramvis, I., Sahin, M., et al. (2008). Promoting axon regeneration in the adult CNS by modulation of the PTEN/mTOR pathway. Science 322, 963–966.10.1126/science.1161566Search in Google Scholar PubMed PubMed Central

Patel, P.H., Thapar, N., Guo, L., Martinez, M., Maris, J., Gau, C.L., Lengyel, J.A., and Tamanoi, F. (2003). Drosophila Rheb GTPase is required for cell cycle progression and cell growth. J. Cell Sci. 116, 3601–3610.10.1242/jcs.00661Search in Google Scholar PubMed

Patel, P.H. and Tamanoi, F. (2006). Increased Rheb-TOR signaling enhances sensitivity of the whole organism to oxidative stress. J. Cell Sci. 119, 4285–4292.10.1242/jcs.03199Search in Google Scholar

Potter, C.J., Huang, H., and Xu, T. (2001). Drosophila Tsc1 functions with Tsc2 to antagonize insulin signaling in regulating cell growth, cell proliferation, and organ size. Cell 105, 357–368.10.1016/S0092-8674(01)00333-6Search in Google Scholar

Pryor, W.M., Biagioli, M., Shahani, N., Swarnkar, S., Huang, W.C., Page, D.T., MacDonald, M.E., and Subramaniam, S. (2014). Huntingtin promotes mTORC1 signaling in the pathogenesis of Huntington’s disease. Sci. Signal. 7, ra10310.1126/scisignal.2005633Search in Google Scholar

Raina, A.K., Pardo, P., Rottkamp, C.A., Zhu, X., Pereira-Smith, O.M., and Smith, M.A. (2001). Neurons in Alzheimer disease emerge from senescence. Mech. Ageing Dev. 123, 3–9.10.1016/S0047-6374(01)00333-5Search in Google Scholar

Rehmann, H., Brüning, M., Berghaus, C., Schwarten, M., Köhler, K., Stocker, H., Stoll, R., Zwartkruis, F.J, and Wittinghofer, A. (2008). Biochemical characterisation of TCTP questions its function as a guanine nucleotide exchange factor for Rheb. FEBS Lett. 582, 3005–3010.10.1016/j.febslet.2008.07.057Search in Google Scholar

Ridet, J.L., Malhotra, S.K., Privat, A., and Gage, F.H. (1997). Reactive astrocytes: cellular and molecular cues to biological function. Trends Neurosci. 20, 570–577.10.1016/S0166-2236(97)01139-9Search in Google Scholar

Rubinsztein, D.C. (2006). The roles of intracellular protein-degradation pathways in neurodegeneration. Nature 443, 780–786.10.1038/nature05291Search in Google Scholar PubMed

Saijo, K., Winner, B., Carson, C.T., Collier, J.G., Boyer, L., Rosenfeld, M.G., Gage, F.H., and Glass, C.K. (2009). A Nurr1/CoREST pathway in microglia and astrocytes protects dopaminergic neurons from inflammation-induced death. Cell 137, 47–59.10.1016/j.cell.2009.01.038Search in Google Scholar PubMed PubMed Central

Saito, K., Araki, Y., Kontani, K., Nishina, H., and Katada, T. (2005). Novel role of the small GTPase Rheb: its implication in endocytic pathway independent of the activation of mammalian target of rapamycin. J. Biochem. 137, 423–430.10.1093/jb/mvi046Search in Google Scholar PubMed

Sancak, Y., Peterson, T.R., Shaul, Y.D., Lindquist, R.A., Thoreen, C.C., Bar-Peled, L., and Sabatini, D.M. (2008). The Rag GTPases bind raptor and mediate amino acid signaling to mTORC1. Science 320, 1496–1501.10.1126/science.1157535Search in Google Scholar PubMed PubMed Central

Sarbassov, D.D., Ali, S.M., Kim, D.H., Guertin, D.A., Latek, R.R., Erdjument-Bromage, H., Tempst, P., and Sabatini, D.M. (2004). Rictor, a novel binding partner of mTOR, defines a rapamycin-insensitive and raptor-independent pathway that regulates the cytoskeleton. Curr. Biol. 14, 1296–1302.10.1016/j.cub.2004.06.054Search in Google Scholar PubMed

Sarkar, S. and Rubinsztein, D.C. (2008). Small molecule enhancers of autophagy for neurodegenerative diseases. Mol. Biosyst. 4, 895–901.10.1039/b804606aSearch in Google Scholar PubMed

Sato, T., Akasu, H., Shimono, W., Matsu, C., Fujiwara, Y., Shibagaki, Y., Heard, J.J., Tamanoi, F., and Hattori, S. (2015). Rheb protein binds CAD (carbamoyl-phosphate synthetase 2, aspartate transcarbamoylase, and dihydroorotase) protein in a GTP- and effector domain-dependent manner and influences its cellular localization and carbamoylphosphate synthetase (CPSase) activity. J. Biol. Chem. 290, 1096–110510.1074/jbc.M114.592402Search in Google Scholar PubMed PubMed Central

Saucedo, L.J., Gao, X., Chiarelli, D.A., Li, L., Pan, D., and Edgar, B.A. (2003). Rheb promotes cell growth as a component of the insulin/TOR signalling network. Nat. Cell Biol. 5, 566–571.10.1038/ncb996Search in Google Scholar PubMed

Schöpel, M., Jockers, K.F., Düppe, P.M., Autzen, J., Potheraveedu, V.N., Ince, S., Yip, K.T., Heumann, R., Herrmann, C., Scherkenbeck, J., et al. (2013). Bisphenol a binds to Ras proteins and competes with guanine nucleotide exchange: implications for GTPase-selective antagonists. J. Med. Chem. 56, 9664–9672.10.1021/jm401291qSearch in Google Scholar PubMed

Schöpel, M., Potheraveedu, V.N., Al-Harthy, T., Abdel-Jalil, R., Heumann, R., and Stoll, R. (2017). The small GTPases Ras and Rheb studied by multidimensional NMR spectroscopy: structure and function. Biol Chem, epub ahead of print; DOI: https://doi.org/10.1515/hsz-2016-0276.Search in Google Scholar PubMed

Sciarretta, S., Zhai, P., Shao, D., Maejima, Y., Robbins, J., Volpe, M., Condorelli, G., and Sadoshima, J. (2012). Rheb is a critical regulator of autophagy during myocardial ischemia: pathophysiological implications in obesity and metabolic syndrome. Circulation 125, 1134–1146.10.1161/CIRCULATIONAHA.111.078212Search in Google Scholar PubMed PubMed Central

Shahani, N., Pryor, W., Swarnkar, S., Kholodilov, N., Thinakaran, G., Burke, R.E., and Subramaniam, S. (2014). Rheb GTPase regulates β-secretase levels and amyloid β generation. J. Biol. Chem. 289, 5799–5808.10.1074/jbc.M113.532713Search in Google Scholar PubMed PubMed Central

Shu, Q., Xu, Y., Zhuang, H., Fan, J., Sun, Z., Zhang, M., and Xu, G. (2014). Ras homolog enriched in the brain is linked to retinal ganglion cell apoptosis after light injury in rats. J. Mol. Neurosci. 54, 243–251.10.1007/s12031-014-0281-zSearch in Google Scholar PubMed

Signer, R.A. and Morrison, S.J. (2013). Mechanisms that regulate stem cell aging and life span. Cell Stem Cell 12, 152–165.10.1016/j.stem.2013.01.001Search in Google Scholar PubMed PubMed Central

Sofroniew, M.V. and Vinters, H.V. (2010). Astrocytes: biology and pathology. Acta Neuropathol. 119, 7–35.10.1007/s00401-009-0619-8Search in Google Scholar

Soga, M., Matsuzawa, A., and Ichijo, H. (2012). Oxidative stress-induced diseases via the ASK1 signaling pathway. Int. J. Cell Biol. 2012, 439587.10.1155/2012/439587Search in Google Scholar

Stenmark, H. and Olkkonen, V.M. (2001). The Rab GTPase family. Genome Biol. 2, reviews3007.1–reviews3007.7.10.1186/gb-2001-2-5-reviews3007Search in Google Scholar

Sugiura, H., Yasuda S., Katsurabayashi S., Kawano H., Endo K., Takasaki K., Iwasaki K., Ichikawa M., Kobayashi T., Hino O., et al. (2015). Rheb activation disrupts spine synapse formation through accumulation of syntenin in tuberous sclerosis complex. Nat. Commun. 6, 6842.10.1038/ncomms7842Search in Google Scholar

Sun, Y., Fang, Y., Yoon, M.S., Zhang, C., Roccio, M., Zwartkruis, F.J., Armstrong, M., Brown, H.A., and Chen, J. (2008). Phospholipase D1 is an effector of Rheb in the mTOR pathway. Proc. Natl. Acad. Sci. USA 105, 8286–8291.10.1073/pnas.0712268105Search in Google Scholar

Swiech, L., Perycz, M., Malik, A., and Jaworski, J. (2008). Role of mTOR in physiology and pathology of the nervous system. Biochim. Biophys. Acta 1784, 116–132.10.1016/j.bbapap.2007.08.015Search in Google Scholar

Takei, N., Inamura, N., Kawamura, M., Namba, H. Hara, K., Yonezawa, K., and Nawa1, H. (2004). Brain-derived neurotrophic factor induces mammalian target of rapamycin-dependent local activation of translation machinery and protein synthesis in neuronal dendrites. J. Neurosci. 24, 9760–9769.10.1523/JNEUROSCI.1427-04.2004Search in Google Scholar

Tapon, N., Ito, N., Dickson, B.J., Treisman, J.E., and Hariharan, I.K. (2001). The Drosophila tuberous sclerosis complex gene homologs restrict cell growth and cell proliferation. Cell 105, 345–355.10.1016/S0092-8674(01)00332-4Search in Google Scholar

Tavazoie, S.F., Alvarez, V.A., Ridenour, D.A., Kwiatkowski, D.J., and Sabatini, B.L. (2005). Regulation of neuronal morphology and function by the tumor suppressors Tsc1 and Tsc2. Nat. Neurosci. 8, 1727–1734.10.1038/nn1566Search in Google Scholar

Tayler, T.D. and Garrity, P.A. (2003). Axon targeting in the Drosophila visual system. Curr. Opin. Neurobiol. 13, 90–95.10.1016/S0959-4388(03)00004-7Search in Google Scholar

Tee, A.R., Manning, B.D., Roux, P.P., Cantley, L.C., and Blenis, J. (2003). Tuberous sclerosis complex gene products, tuberin and hamartin, control mTOR signaling by acting as a GTPase-activating protein complex toward Rheb. Curr. Biol. 13, 1259–1268.10.1016/S0960-9822(03)00506-2Search in Google Scholar

Tee, A.R., Blenis, J., and Proud, C.G. (2005). Analysis of mTOR signaling by the small G-proteins, Rheb and RhebL1. FEBS Lett. 579, 4763–4768.10.1016/j.febslet.2005.07.054Search in Google Scholar PubMed

Thomanetz, V., Angliker, N., Cloëtta, D., Lustenberger, R.M., Schweighauser, M., Oliveri, F., Suzuki, N., and Rüegg, M.A. (2013). Ablation of the mTORC2 component rictor in brain or Purkinje cells affects size and neuron morphology. J. Cell Biol. 201, 293–308.10.1083/jcb.201205030Search in Google Scholar PubMed PubMed Central

Tyagi, R., Shahani, N., Gorgen, L., Ferretti, M., Pryor, W., Chen, P.Y., Swarnkar, S., Worley, P.F., Karbstein, K., Snyder, S.H., et al. (2015). Rheb inhibits protein synthesis by activating the PERK-eIF2 a signaling cascade. Cell Rep. 10, 684–693.10.1016/j.celrep.2015.01.014Search in Google Scholar PubMed PubMed Central

Uhlmann, E.J., Li, W., Scheidenhelm, D.K., Gau, C.L., Tamanoi, F., and 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, 180–188.10.1002/glia.20036Search in Google Scholar PubMed

Uttara, B., Singh, A.V., Zamboni, P., and Mahajan, R.T. (2009). Oxidative stress and neurodegenerative diseases: a review of upstream and downstream antioxidant therapeutic options. Curr. Neuropharmacol. 7, 65–74.10.2174/157015909787602823Search in Google Scholar PubMed PubMed Central

Wang, T., Lao, U., and Edgar, B.A. (2009). TOR-mediated autophagy regulates cell death in Drosophila neurodegenerative disease. J. Cell Biol. 186, 703–711.10.1083/jcb.200904090Search in Google Scholar PubMed PubMed Central

Wang, J.T., Medress, Z.A., and Barres, B.A. (2012). Axon degeneration: molecular mechanisms of a self-destruction pathway. J. Cell. Biol. 196, 7–18.10.1083/jcb.201108111Search in Google Scholar PubMed PubMed Central

Wataya-Kaneda, M., Kaneda, Y., Hino, O., Adachi, H., Hirayama, Y., Seyama, K., Satou, T., and Yoshikawa, K. (2001). Cells derived from tuberous sclerosis show a prolonged S phase of the cell cycle and increased apoptosis. Arch. Dermatol. Res. 293, 460–469.10.1007/s004030100259Search in Google Scholar PubMed

Weatherill, D.B., Dyer, J., and Sossin, W.S. (2010). Ribosomal protein S6 kinase is a critical downstream effector of the target of rapamycin complex 1 for long-term facilitation in Aplysia. J. Biol. Chem. 285, 12255–12267.10.1074/jbc.M109.071142Search in Google Scholar PubMed PubMed Central

Wright, A.L., Zinn, R., Hohensinn, B., Konen, L.M., Beynon, S.B., Tan, R.P., Clark, I.A., Abdipranoto, A., and Vissel, B. (2013). Neuroinflammation and neuronal loss precede Aβ plaque deposition in the hAPP-J20 mouse model of Alzheimer’s disease. PLoS One 8, e59586.10.1371/journal.pone.0059586Search in Google Scholar

Wu, D., Klaw, M.C., Kholodilov, N., Burke, R.E., Detloff, M.R., Côté, M.P., and Tom, V.J. (2016). Expressing constitutively active rheb in adult dorsal root ganglion neurons enhances the integration of sensory axons that regenerate across a chondroitinase-treated dorsal root entry zone following dorsal root crush. Front. Mol. Neurosci. 9, 1–18.10.3389/fnmol.2016.00049Search in Google Scholar

Yamagata, K., Sanders, L.K., Kaufmann, W.E., Yee, W., Barnes, C.A., Nathans, D., and Worley, P.F. (1994). rheb, a growth factor- and synaptic activity-regulated gene, encodes a novel Ras-related protein. J. Biol. Chem. 269, 16333–16339.10.1016/S0021-9258(17)34012-7Search in Google Scholar

Yang, Q., Inoki, K., Kim, E., and Guan, K.L. (2006). TSC1/TSC2 and Rheb have different effects on TORC1 and TORC2 activity. Proc. Natl. Acad. Sci. USA 103, 6811–6816.10.1073/pnas.0602282103Search in Google Scholar PubMed PubMed Central

Yang, W., Jiang, W., Luo, L., Bu, J., Pang, D., Wei, J., Du, C., Xia, X., Cui, Y., Liu, S., et al. (2014 ). Genetic deletion of rheb1 in the brain reduces food intake and causes hypoglycemia with altered peripheral metabolism. Int. J. Mol. Sci. 15, 1499–1510.10.3390/ijms15011499Search in Google Scholar PubMed PubMed Central

Yaniv, S.P., Issman-Zecharya, N., Oren-Suissa, M., Podbilewicz, B., and Schuldiner, O. (2012). Axon regrowth during development and regeneration following injury share molecular mechanisms. Curr. Biol. 22, 1774–1782.10.1016/j.cub.2012.07.044Search in Google Scholar PubMed

Yankner, B.A., Lu, T., and Loerch, P. (2008). The aging brain. Annu. Rev. Pathol. 3, 41–66.10.1146/annurev.pathmechdis.2.010506.092044Search in Google Scholar PubMed

Yasuda, S., Sugiura, H., Katsurabayashi, S., Shimada, T., Tanaka, H., Takasaki, K., Iwasaki, K., Kobayashi, T., Hino, O., and Yamagata, K. (2014). Activation of Rheb, but not of mTORC1, impairs spine synapse morphogenesis in tuberous sclerosis complex. Sci. Rep. 4, 5155.10.1038/srep05155Search in Google Scholar PubMed PubMed Central

Yoshida, S., Hong, S., Suzuki, T., Nada, S., Mannan, A.M., Wang, J., Okada, M., Guan, K. L., and Inoki, K. (2011). Redox regulates mammalian target of rapamycin complex 1 (mTORC1) activity by modulating the TSC1/TSC2-Rheb GTPase pathway. J. Biol. Chem. 286, 32651–3266010.1074/jbc.M111.238014Search in Google Scholar PubMed PubMed Central

Zhang, Y. and Manning, B.D. (2015). mTORC1 signaling activates NRF1 to increase cellular proteasome levels. Cell Cycle 14, 2011–2017.10.1080/15384101.2015.1044188Search in Google Scholar PubMed PubMed Central

Zhang, Y., Gao, X., Saucedo, L.J., Ru, B., Edgar, B.A., and Pan, D. (2003). Rheb is a direct target of the tuberous sclerosis tumour suppressor proteins. Nat. Cell Biol. 5, 578–581.10.1038/ncb999Search in Google Scholar

Zhang, J., Kim, J., Alexander, A., Cai, S., Tripathi, D.N., Dere, R., Tee, A.R., Tait-Mulder J., Di Nardo, A., Han, J.M., et al. (2013). A tuberous sclerosis complex signalling node at the peroxisome regulates mTORC1 and autophagy in response to ROS. Nat. Cell Biol. 15, 1186–1196.10.1038/ncb2822Search in Google Scholar

Zhao, G. and Flavin, M.P. (2000). Differential sensitivity of rat hippocampal and cortical astrocytes to oxygen-glucose deprivation injury. Neurosci. Lett. 285, 177–180.10.1016/S0304-3940(00)01056-9Search in Google Scholar

Zhou, X., Ikenoue, T., Chen, X., Li, L., Inoki, K., and Guan, K.L. (2009). Rheb controls misfolded protein metabolism by inhibiting aggresome formation and autophagy. Proc. Natl. Acad. Sci. USA 106, 8923–8928.10.1073/pnas.0903621106Search in Google Scholar PubMed PubMed Central

Zoncu, R., Efeyan, A., and Sabatini, D.M. (2011). mTOR: from growth signal integration to cancer, diabetes and ageing. Nat. Rev. Mol. Cell Biol. 12, 21–35.10.1038/nrm3025Search in Google Scholar PubMed PubMed Central

Zou, Y., Jiang, W., Wang, J., Li, Z., Zhang, J., Bu, J., Zou, J., Zhou, L., Yu, S., Cui, Y., et al. (2014). Oligodendrocyte precursor cell-intrinsic effect of Rheb1 controls differentiation and mediates mTORC1-dependent myelination in brain. J. Neurosci. 34, 15764–15778.10.1523/JNEUROSCI.2267-14.2014Search in Google Scholar PubMed PubMed Central

Received: 2016-10-16
Accepted: 2017-2-2
Published Online: 2017-2-17
Published in Print: 2017-5-1

©2017, Veena Nambiar Potheraveedu et al., published by De Gruyter.

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Downloaded on 19.5.2024 from https://www.degruyter.com/document/doi/10.1515/hsz-2016-0312/html
Scroll to top button