Next Article in Journal
Enhancement of Commercial Antifungal Agents by Kojic Acid
Next Article in Special Issue
Role of Oxidative Stress in Hepatocarcinogenesis Induced by Hepatitis C Virus
Previous Article in Journal
Phorbol Esters Isolated from Jatropha Meal Induced Apoptosis-Mediated Inhibition in Proliferation of Chang and Vero Cell Lines
Previous Article in Special Issue
Effect of Calcium and Potassium on Antioxidant System of Vicia faba L. Under Cadmium Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxidant Stress and Signal Transduction in the Nervous System with the PI 3-K, Akt, and mTOR Cascade

1
Laboratory of Cellular and Molecular Signaling, Newark, NJ 07101, USA
2
Cancer Institute of New Jersey, New Brunswick, NJ 08903, USA
3
New Jersey Health Sciences University, 205 South Orange Avenue, Newark, NJ 07101, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2012, 13(11), 13830-13866; https://doi.org/10.3390/ijms131113830
Submission received: 8 October 2012 / Revised: 19 October 2012 / Accepted: 19 October 2012 / Published: 26 October 2012
(This article belongs to the Special Issue Oxidative Stress and Ageing)

Abstract

:
Oxidative stress impacts multiple systems of the body and can lead to some of the most devastating consequences in the nervous system especially during aging. Both acute and chronic neurodegenerative disorders such as diabetes mellitus, cerebral ischemia, trauma, Alzheimer’s disease, Parkinson’s disease, Huntington’s disease, and tuberous sclerosis through programmed cell death pathways of apoptosis and autophagy can be the result of oxidant stress. Novel therapeutic avenues that focus upon the phosphoinositide 3-kinase (PI 3-K), Akt (protein kinase B), and the mammalian target of rapamycin (mTOR) cascade and related pathways offer exciting prospects to address the onset and potential reversal of neurodegenerative disorders. Effective clinical translation of these pathways into robust therapeutic strategies requires intimate knowledge of the complexity of these pathways and the ability of this cascade to influence biological outcome that can vary among disorders of the nervous system.

Graphical Abstract

1. Introduction

Oxidative stress through the generation of reactive oxygen species (ROS) can severely impact the nervous system and related pathways that control immune function and cellular metabolism during aging. ROS are formed through superoxide free radicals, hydrogen peroxide, singlet oxygen, nitric oxide (NO), and peroxynitrite [13]. ROS can result in mitochondrial dysfunction, the loss of DNA integrity, and the misfolding of proteins that promote cell injury in the nervous system [2,4,5]. Antioxidant systems in the body that include catalase, superoxide dismutase, glutathione peroxidase, and vitamins C, D, E, and K limit the production of ROS to non-toxic levels [613]. However, excessive production of ROS or impairments in the endogenous antioxidant system can lead to oxidative stress and cell death through DNA degradation, mitochondrial dysfunction, and protein misfolding [2,3,1417]. Oxidative stress can lead to cognitive loss [14,1820], Alzheimer’s disease (AD) [1,13,2024], Huntington’s disease (HD) [2531], Parkinson’s disease [1,9,26,3239], epilepsy [25,26,28,4043], and acute injury such as during stroke and trauma [4454]. Oxidative stress also modulates inflammatory cell function [5560] and alters cellular metabolism that leads to complications of diabetes mellitus [5,6171]. Although multiple cellular pathways are impacted during oxidative stress in the nervous system, the kinase cascade that involves phosphoinositide 3–kinase (PI 3-K), protein kinase B (Akt), and the mammalian target of rapamycin (mTOR) offers exciting prospects for the development of new treatment strategies since these pathways are closely integrated and directly affect both acute and chronic neurological disorders [25,55,7274]. Greater understanding of these pathways can foster novel as well as safe clinical treatment avenues for disorders of the nervous system.

2. Oxidant Stress and Cellular Injury

2.1. Cellular Survival and Demise during Neurodegeneration

Oxidative stress pathways that involve the PI 3-K, Akt, and mTOR cascade can lead to cellular injury through the processes of apoptosis and autophagy [39,75] (Figure 1). During neurodegenerative disorders, apoptosis may be a significant contributor of cell dysfunction and loss. Apoptotic DNA fragmentation [76] and caspase activation is present [77] in the brains of patients with Alzheimer’s disease (AD). Alzheimer’s disease models have demonstrated apoptotic proteins in the brain [78]. Apoptotic neuronal nuclei and caspase 3 have been identified in the postmortem nigra of Parkinson’s disease (PD) patients [79]. For immune mediated cells of the brain, loss of microglia can occur through apoptotic cell death [80,81]. In regards to autophagy, this process may play a dual role that can either protect cell survival or potentiate cell injury. During oxidative stress, autophagy can lead to cell death in cerebral astrocytes [82], in cortical neurons [83], in spinal cord motor neurons [84], in purkinje neurons [85], and in sympathetic neurons [86]. Other toxins such as glutamate, potassium deprivation, and staurosporine also lead to autophagy and cell death [85]. During metabolic disease, apoptosis can lead to vascular complications, impair endothelial cell survival, destroy immune mediated cells, inhibit wound repair, and injure neurons [1,87,88]. However, autophagy and apoptosis may have similar outcomes that are intertwined. The agent methamphatamine leads to cell death through both apoptosis and autophagy by inhibiting the disassociation of the “apoptotic-autophagic complex” Bcl-2/Beclin 1 complex [89]. Autophagy and apoptosis also may have opposing roles. Apoptosis may conversely require the inhibition of autophagy [90] or apoptotic neuronal cell death may be independent of the onset of autophagy [91]. Under some conditions, autophagy can offer cytoprotection during neurodegenerative disorders [92,93]. Autophagy may be necessary to protect against neuronal cell loss and α-synuclein toxicity in PD [92]. Mutation of α-synuclein and accumulation of wild-type α-synuclein in dopaminergic neurons have been associated with progression of Parkinson’s disease [94]. Autophagy is necessary for the clearance of aberrant α-synuclein in neurons since inhibition of this autophagic pathway leads to accumulation of high molecular weight and detergent insoluble α-synuclein and neurotoxicity [94]. Mutant α-synuclein, which is poorly internalized into lysosomes, also is degraded by macroautophagy [92]. During metabolic disease such as diabetes mellitus (DM), autophagy may be necessary to remove misfolded proteins and eliminate non-functioning mitochondria in β-cells to prevent β-cell dysfunction and the onset of DM [95]. Exercise in mice has been shown to initiate autophagy and regulate glucose homeostasis [96]. However, the generation of advanced glycation end products (AGEs) during DM also may lead to autophagy that contributes to vascular smooth muscle proliferation, atherosclerosis [97], and potential cardiomyopathy [64]. The elevation of free fatty acids in cell models of DM suggest that fatty acids may be necessary to activate autophagy in beta cells through endoplasmic reticulum stress [98].

2.2. Apoptotic Early and Late Phases

Apoptosis consists of two distinct components that involve genomic DNA degradation and the loss of plasma membrane lipid asymmetry [2]. The loss of asymmetry of membrane phosphatidylserine (PS) distribution is an early component of apoptosis that can be reversible [99]. A later event, the cleavage of genomic DNA into fragments can occur once a cell has been committed to die [100]. Once initiated, genomic DNA degradation usually is not reversible. Enzymes responsible for DNA degradation include the acidic cation independent endonuclease (DNase II), cyclophilins, and the 97 kDa magnesium-dependent endonuclease. Three separate endonuclease activities exist in neurons that include a constitutive acidic cation-independent endonuclease, a constitutive calcium/magnesium-dependent endonuclease, and an inducible magnesium dependent endonuclease [22]. Both membrane PS exposure and genomic DNA degradation are considered to be the outcomes of a series of activation of nucleases and proteases that occurs late during apoptosis [1,88]. Exposure of membrane PS residues during oxidative stress can occur with sepsis, ischemia, vascular clot formation, and β-amyloid deposition [101103]. The early phase of apoptosis can tag cells with membrane PS residues to alert inflammatory cells to engulf and remove injured cells [104,105]. For this to occur such as during periods of oxidative stress, inflammatory cells increase their expression of the membrane phosphatidylserine receptor (PSR) [106108]. To promote cell survival, modulation of inflammatory cell activation is necessary, since removal of temporarily injured cells expressing membrane PS residues can lead to the loss of functional cells [109,110].

2.3. The Sub-Constructs of Autophagy

In contrast to apoptosis, autophagy has three different categories known as microautophagy, macroautophagy, and chaperone-mediated autophagy [99,111,112]. Macroautophagy consists of the degradation of cytoplasmic material and the sequestration of the cytoplasmic protein and organelles into autophagosomes. Autophagosomes fuse with lysosomes for degradation and are then recycled for future cellular processes [113]. Microautophagy represents the sequestration of cytoplasmic components by invagination of the lysosomal membrane. Vesicles subsequently formed are transferred to the lumen of the lysosomes for digestion. In chaperone-mediated autophagy, the cytoplasmic component is transported by cytosolic chaperones to the receptors on the lysosomal membranes for translocation across lysosomal membranes into the lumen.
In yeast, thirty-three autophagic related genes (Atg) have been identified. Of this family of genes, Atg1 and Atg13 are associated with the PI 3-K, Akt, and mTOR cascade [114116]. In yeast, Atg13 (also known as Apg13) is phosphorylated through a Tor dependent mechanism, leading its release from Atg1 and a reduction in Atg1 activity. The Tor kinase was first isolated in Saccharomyces cerevisiae through the analysis of rapamycin toxicity using rapamycin-resistant TOR mutants in yeast that resulted in the identification of the yeast genes TOR1 and TOR2 that encode two isoforms in yeast Tor1 and Tor2 [117]. In mammals, a single gene FRAP1 encodes mTOR [99,118]. Upon starvation or Tor inhibition, Atg13 is dephosphorylated, binds to, and activates Atg1, leading to autophagosome formation [116]. In mammals, a similar regulation of autophagy through mTOR also exists. Two mammalian homologues of Atg1, UNC-51 like kinase 1 (ULK1) and ULK2, have been identified [119121]. Mammalian Atg13 binds to ULK1, ULK2, and FIP200 (FAK-family interacting protein of 200 kDa) to activate ULKs and facilitate the phosphorylation of FIP200 by ULKs [122,123]. Similar to TOR in yeast, mTOR phosphorylates the mammalian homologue Atg13 and the mammalian Atg1 homologues ULK1 and ULK2 to block autophagy [123]. The focal adhesion kinase family interacting protein of 200 kDa (FIP200) has been identified as a ULK binding protein. FIP200 and Atg13 are vital for the stability and activation of ULK1. Mammalian Atg13 binds to ULK1/2 and FIP200 to activate ULKs and facilitate the phosphorylation of FIP200 by ULKs [123]. mTOR activation prevents autophagy in mammalian cells through inhibition of the ULK-Atg13–FIP200 complex by phosphorylating Atg13 and ULKs. In the absence of mTOR activity, dephosphorylation of ULKs and Atg13 ensues leading to the induction of autophagy [122,123].

3. Signal Transduction and Cell Survival with PI 3-K, Akt, and mTOR

3.1. PI 3-K and PDK

Receptor tyrosine kinase (RTK) and the G protein-coupled receptor (CPCR) are required for PI 3-K activation. Following activation, PI 3-K phosphorylates membrane lipids and mediates the transition of Akt from the cytosol to the plasma membrane by promoting the binding of Akt to PI-3,4-P2 and PI-3,4,5-P3 through the plectrin homology (PH) domain (Figure 1). Akt is then phosphorylated on the residues of serine473 and threonine308 by phosphoinositide dependent kinase (PDK) PDK1 and PDK2. PDK1 is responsible for phosphorylating Akt at threonine308[124]. PDK1 contains a C-terminal PH domain that PI-3, 4, 5-P3 can recruit PDK1 to the cell membrane and bind to the C-terminal HM domain of Akt. PDK1 cannot directly phosphorylate Akt on serine473, but phosphorylation of Akt on serine473 is necessary for the full activation of Akt. PDK2 like kinases, such as integrin-linked kinase, DNA dependent protein kinase, PKCβ, and mTORC2, have been identified to promote Akt phosphorylation on serine473[125]. In contrast, the phosphatase and tensin homolog deleted from chromosome 10 (PTEN), which specifically dephosphorylates PI-3,4-P2 and PI-3,4,5-P3 at the D3 position can block PI 3-K signaling and inhibit Akt activation.

3.2. Akt

A number of pathways can influence Akt activity during oxidative stress [126130]. The 90 kDa heat shock protein (Hsp90) that is involved in modulating oxidative stress in cells [131] can increase Akt activity through the inhibition of inhibiting protein phosphatase 2A (PP2A). In addition, the T cell leukemia/lymphoma 1 (TCL1) protein binds to the PH domain of Akt to increase Akt activity (Figure 1). In regards to down-regulation of Akt activity, the carboxyl-terminal modulator protein (CTMP) binds to the carboxyl-terminal regulatory domain of Akt1 at the plasma membrane to prevent Akt1 from phosphorylation. The src homology 2 (SH2) domain-containing inositol phosphatase (SHIP) is an inositol 5′ phosphatase that dephosphorylates inositides and phosphoinositides on the 5′-position [55]. Both SHIP1 and SHIP2 can negatively regulate the activity of Akt. PI-3, 4, 5-P3 are transformed into PI-3, 4-P2 that is less potent than PI-3, 4, 5-P3 to recruit Akt. The SH2 domains containing protein-tyrosine phosphatases SHP1 and SHP2 also modulate the activity of PI 3-K. SHP1 associates with the p85 subunit of PI 3-K to negatively regulate the activation of PI 3-K. SHP2 can be necessary for agents that promote cell differentiation to lead to the activation of PI 3-K and Akt [132].

3.3. mTOR

In relation to mTOR, which also is known as mechanistic target of rapamycin and FK506-binding protein 12-rapamycin complex-associated protein 1 (FRAP1), Akt is a strong stimulator of mTORC1 to lead to the activation of mTORC1 [133]. As a component of the PI 3-K related kinase family that is activated through the PI 3-K and Akt, mTOR is a 289-kDa serine/threonine protein kinase that can control transcription, cytoskeleton organization, cellular survival, and cellular metabolism [25,87,99,133135]. mTOR signaling is dependent upon the protein complexes mTOR Complex 1 (mTORC1) or mTOR Complex 2 (mTORC2) that each contain mTOR (Figure 1). p70 ribosomal S6 kinase (p70S6K) and eukaryotic initiation factor 4E (eIF4E)-binding protein 1 (4EBP1) are downstream targets of mTORC1 [99,136]. Phosphorylation of p70S6K promotes mRNA biogenesis, translation of ribosomal proteins, and cell growth [137,138]. In contrast, phosphorylation of 4EBP1 results in its inactivation. Hypophosphorylated 4EBP1 is active and binds competitively with eukaryotic translation initiation factor 4 gamma (eIF4G) to eukaryotic translation initiation factor 4 epsilon (eIF4E) that regulate translation initiation by interacting with the 5′-mRNA cap structure. The phosphorylation of 4EBP1 by mTORC1 results in its dissociation from eIF4E allowing eIF4G to interact with eIF4E and promotes protein translation [139,140]. Tuberous sclerosis complex (TSC) 1 (hamartin)/TSC2 (tuberin) complex is one of the targets of Akt for the modulation of mTORC1 activity. In the absence of Akt, the TSC1/TSC2 complex is a negative regulator of mTORC1. TSC2 functions as a GTPase-activating protein (GAP), converting a small G protein Ras homologue enriched in brain (Rheb-GTP) to the inactive GDP-bound form (Rheb-GDP) [141]. Once active, Rheb-GTP can directly interact with Raptor to activate mTORC1 and also regulate the binding of 4EBP1 to mTORC1 [142]. Akt phosphorylates TSC2 on multiple sites that leads to the destabilization of TSC2 and disruption of its interaction with TSC1. The phosphorylation of TSC2 on the residues of serine939, serine981, and threonine1462 can increase its binding to the anchor protein 14-3-3 and lead to the cellular sequestration by 14-3-3, disruption of the TSC1/TSC2 complex, and subsequent activation of Rheb and mTORC1 [143].
The proline rich Akt substrate 40 kDa (PRAS40) and I-kappaB kinase (IKK) also are targets of Akt to control the activation of mTORC1. PRAS40 can be phosphorylated on several residues including serine183, serine212, serine221, and threonine246[144,145]. The serine sites are targets of mTOR and the residue of threonine246 is the phosphorylation target of Akt. The phosphorylation of PRAS40 leads to its dissociation with Raptor [146] and promotes the binding of PRAS40 to the cytoplasmic docking protein 14-3-3 [147149]. This removes PRAS40 from interacting with Raptor and facilitates the activation of mTORC1 [150]. Akt also has been shown to promote the activation of mTORC1 through IKKα. Loss of IKKα inhibits mTOR activation in Akt-active cells during inactivation of the negative PI 3-K regulator PTEN [151]. Within IKK, IKKα and IKKβ are catalytic subunits of IKK that possess serine/threonine kinase activity [152]. IKKα regulates mTOR activity by associating with Raptor that is Akt dependent [151]. In addition, IKKβ can phosphorylate TSC1 on serine487 and serine511 leading to the suppression of TSC1, disruption of TSC1/TSC2 complex, and the activation of mTORC1 [153]. Phosphorylation of IKKβ also has been associated with the activation of downstream pathways of mTOR signaling that involve p70S6K [154].

3.4. Apoptosis and Autophagy in the PI 3-K, Akt, and mTOR Cascade

The PI 3-K, Akt, and mTOR cascade closely govern cell survival during apoptosis and autophagy in the nervous system (Figure 1). PI 3-K and Akt activation can foster endothelial survival [128,155161], limit neuronal injury [105,162168], and block inflammatory cell death [59,81,106108,169,170], and block neuronal injury [105,162168]. A number of cellular pathways can be responsible for the activation of PI 3-K and Akt. For example, intracellular calcium release that is controlled by calmodulin activation leads to the association of calcium and calmodulin with the 85 kDa regulatory subunit of PI 3-K to activate Akt and promote neuronal survival [171,172]. Other pathways may be mediated through growth factors and cytokines, such as erythropoietin (EPO) [173]. For example, EPO induces the phosphorylation of Akt at serine473 to lead to its activation. EPO can protect dorsal root ganglion neurons in animal models of diabetes mellitus with streptozotocin through pathways that activate Akt [174]. EPO relies upon Akt activation in pathways that require sirtuins to maintain cerebral vascular cell survival during oxidative stress [159]. During EPO exposure, Akt is activated that results in the post-translational phosphorylation of forkhead transcription factors, such as FoxO proteins. Once phosphorylated, FoxO is sequestered in the cytoplasm by association with 14-3-3 proteins and transcription of “pro-apoptotic” genes is prevented [175]. Mammalian forkhead transcription factors of the O class (FoxO1, FoxO3, FoxO4, and FoxO6), such as FoxO3a, are intimately involved with cellular survival, metabolism, insulin sensitivity, and oxidative stress [176,177]. Akt limits apoptosis through the phosphorylation of FoxO proteins [176,178181]. Akt phosphorylates the residue of serine253 of FoxO3a resulting in its export from the nucleus to the cytoplasm and blocking FoxO3a from activating apoptotic genes. Akt is required for EPO to promote protection such as during glyoxal-advanced glycation end products (AGEs) exposure [167], retinal detachment [182], β-amyloid (Aβ) exposure [80,183185], hypoxia [156,186], and oxidative stress [187189]. In addition, the downstream targets of Akt include BAD, caspase 9, and glycogen synthase kinase-3β (GSK-3β) [190192]. Phosphorylation of GSK-3β, BAD, and caspase 9 by Akt prevents apoptosis and promotes cell longevity pathways [1,3,18,193]. Akt can phosphorylate GSK-3β at serine9 and inactivate this enzyme, this preventing GSK-3β from initiating an apoptotic pathway. Phosphorylation of Bad at serine136 by Akt can result in the inactivation of Bad and prevent neuronal apoptosis [159,194].
Similar to Akt, mTOR can rely upon pathways such as BAD to control apoptosis. mTOR also regulates apoptotic cell death through the mTOR signaling pathways of p70S6K and 4EBP1 [87,133,195]. Phosphorylation of BAD leads to the dissociation of BAD from Bcl-2/Bcl-xL that prevents apoptotic cell death and increases BAD binding to the cytoplasmic docking protein 14-3-3. Activation of p70S6K also promotes the phosphorylation of BAD to limit apoptotic cell injury [195]. The activation of mTOR and p70S6K may also decrease apoptosis in non-neuronal astrocytes through pathways that can increase “anti-apoptotic” Bcl-2/Bcl-xL expression [195]. Growth factors, such as insulin, prevent apoptosis in rat retinal neuronal cells against serum deprivation through activation of mTOR and p70S6K [196]. In addition, over-expression of wild type p70S6K increases cell protection by insulin. In contrast, over-expression of a dominant-negative mutant of p70S6K results in the loss of the ability of insulin to protect neurons [196]. Other growth factors similar to insulin, such as EPO [100,197], also have been reported to be dependent upon mTOR, p70S6K, and 4EBP1 for cytoprotection against apoptosis [145,184,198200]. In regards to 4EBP1, inhibition of mTOR signaling by small interfering RNA (siRNA) inhibits phosphorylation of both p70S6K and 4EBP1 that results in apoptosis [201]. In the absence of mTOR activity, 4EBP1 has increased binding to eIF4E that can lead to the translation of apoptotic promoting proteins and also initiate autophagy [202].
mTOR also works in conjunction with Akt to offer cellular protection in the nervous system. mTOR and Akt are necessary for the protection of endothelial cells against apoptosis [203] and to prevent cell injury from FoxO3a [155,203]. Inflammatory cells succumb to apoptotic injury during oxidative stress if deprived of Akt and mTOR activation [184,204]. Administration of agents that enhance mTOR and Akt activity can prevent apoptotic dopaminergic neuronal death [205].
Akt also functions to modulate apoptosis through the mTOR signaling pathway of PRAS40. Phosphorylation of PRAS40 by Akt can block the activity of this substrate, lead to its dissociation from mTORC1 to allow mTOR activation, and prevent apoptotic cell injury [147,206,207]. PRAS40 is an important regulatory component for protection in the nervous system. Inhibition of PRAS40 activity can prevent cell death against oxidative stress and Aβ suggesting that PRAS40 may be a critical target to prevent neurodegeneration [147,206]. In addition, PRAS40 appears to be closely aligned with the apoptotic injury cascade since PRAS40 can directly control caspase 3 activation [147].
Under some circumstances such as during chronic disorders of neurodegeneration, inhibition of mTOR activity may be beneficial as opposed to activation of mTOR. Activation of mTOR and mTORC1 signaling can promote cell cycle induction that can be detrimental to post-mitotic neurons. Post-mitotic neurons that attempt to enter the cell cycle during AD do not replicate, but can result in apoptotic cell death [208,209]. In studies with amyloid oligomer exposure, neurons can be prevented from entering the cell cycle during the inhibition of mTOR and thus be protected from apoptosis [210]. Furthermore, blockade of the PI 3-K, Akt, and mTOR cascade can lead to the induction of autophagy that may be necessary for cell protection. Under some conditions that involve AD, PD, or HD, induction of autophagy can be cytoprotective. The degree of mTOR activation may be a significant variable in disorders that under some conditions can benefit from precise mTOR activity. During the early phases of autophagy, mTOR activity can be inhibited [211]. However, re-activation of mTOR appears to be required to continue with autophagy as long as elevated levels of mTOR activity do not lead to the eventual blockade of autophagy [212]. This modulation of mTOR with autophagy may be a conserved response in multiple cell systems that is governed by nutrient availability [212].

4. Clinical Disorders Modulated by PI 3-K, Akt, and mTOR

4.1 Diabetes Mellitus in the Nervous System

Throughout the world, more than 165 million individuals are afflicted with DM and it is expected that close to 400 million individuals by the year of 2030 will suffer from DM. The incidence of obesity in the population throughout the world is increasing at an alarming rate that ultimately leads to metabolic disease and diabetes mellitus (DM) [66]. Recent studies have shown that the duration of obese-years rather than body mass index (BMI) translates into a strong risk for developing DM [213]. Increased weight gain also leads to other disorders that may be a result of metabolic disease, such as coronary artery calcifications and the loss of cognition [68,214]. In younger individuals, impaired glucose tolerance is of significant concern, since those with impaired glucose tolerance have a greater than twice the risk for the development of diabetic complications than individuals with normal glucose tolerance [87,129].
Given the progressive development of DM in the population, elucidation of novel pathways that can regulate the onset and progression of diabetic complications in the nervous system and related cardiovascular pathways would be highly desirable. One signaling pathway that is intimately tied to the PI 3-K, Akt, and mTOR pathway in DM involves sirtuins (Figure 2). Sirtuins are class III NAD+-dependent protein histone deacetylases that are the mammalian homologues of Sir2 of the yeast silent information regulator-2 (Sir2) [1]. Of the seven mammalian homologues of Sir2, SIRT1 has been shown to be protective against cellular injury during DM. SIRT1 activation prevents endothelial senescence during hyperglycemia, blocks atherosclerotic lesions during elevated lipid states, modulates adipocyte differentiation, and prevents endothelial cell apoptosis during experimental diabetes [1,128,159,215,216]. SIRT1 also can increase lifespan in higher organisms such as Drosophila and protect cells from oxidative stress [102,217]. Loss of SIRT1 is associated with insulin resistance. Gene deletion or inhibition of SIRT1 impairs insulin signaling by interfering with insulin stimulated insulin receptor phosphorylation and glycogen synthase [218]. In contrast, over-expression of SIRT1 decreases hepatic steatosis and improves insulin sensitivity that leads to improved glucose homeostasis [219]. SIRT1 can increase insulin signaling in insulin-sensitive organs through Akt and PI 3-K [220] and can stimulate glucose-dependent insulin secretion from pancreatic β cells by repressing the uncoupling protein (UCP) gene UCP2 [221]. SIRT1 also controls insulin sensitivity through the inhibition of tyrosine phosphatase1B (PTP1B). SIRT1 over-expression or SIRT1 activation can reduce both PTP1B mRNA and protein levels during insulin-resistance. However, an increase in PTP1B expression prevents SIRT1 mediated glucose uptake and insulin receptor phosphorylation in response to insulin stimulation [218]. SIRT1 also may improve insulin sensitivity through the regulation of fat mobilization, gluconeogenesis, and inflammation [1,3,215].
Agents such as EPO increase endogenous SIRT1 activity and foster nuclear subcellular trafficking of SIRT1 [159]. SIRT1 increases Akt activity that is considered to be a principle pathway for cellular proliferation, survival, and for mediating insulin signaling [6,52,55,128,159,222]. SIRT1 activates Akt1 to also control the phosphorylation and subcellular trafficking of FoxO3a [159]. SIRT1 relies upon the Akt pathway for cytoprotection and the subsequent phosphorylation of target genes such as FoxOs [159,220]. However, SIRT1 may control a fine balance over FoxO activity since acetylation of FoxOs also can limit their transcriptional activity. SIRT1 can lead to the deacetylation of FoxOs to increase the activity of these transcription factors [178,181]. Yet, the deacetylation of FoxOs by SIRT1 may ultimately lead to degradation of these transcription factors [223]. SIRT1 also may rely upon the regulation of FoxO transcription factors to protect against inflammation, to promote β-cell function during DM, and to maintain endothelial survival during oxidative stress [3,181].
Signaling through SIRT1, PI 3-K, and Akt during cell metabolism also is dependent upon mTOR. For example, loss of SIRT1 expression leads to hepatic glucose overproduction, hyperglycemia, products of oxidative stress, and inhibition of the gene encoding Rictor that results in impaired TORC2 and Akt signaling [224]. Under some conditions, SIRT1 and mTOR may have an inverse relationship. SIRT1 can attenuate hepatic steatosis, ameliorate insulin resistance, and restore glucose homeostasis primarily through the inhibition of mTORC1 [219]. SIRT1 also may promote neuronal growth through pathways that inhibit mTOR signaling [225].
Independently, mTOR signaling also may be necessary for maintenance of insulin function. Activation of p70S6K, and inhibition of 4EBP1 in pancreatic β-cells in murine models of DM, results in improved insulin secretion and resistance to β-cell streptozotocin toxicity and obesity [226]. However, loss of p70S6K activity with combined loss of the mTORC2 substrate Akt2 results in defects in both insulin action and β-cell function, suggesting that both mTOR components are required to maintain insulin signaling and prevent DM [227]. Furthermore, mTOR inhibition with rapamycin application leads to insulin resistance, reduces β-cell function and mass, limits insulin secretion, and results in DM [228]. Loss of mTOR signaling with inhibition of p70S6K also can result in hypoinsulinemia, glucose intolerance, insulin insensitivity to glucose secretion, and a decrease in pancreatic β-cell size [25,87]. Although inhibition of mTOR can reduce food intake and prevent fat-diet induced obesity in mice, loss of mTOR activity also attenuates glucose uptake and metabolism in skeletal muscle through the prevention of insulin generated Akt activation and alteration in the translocation of glucose transporters to the plasma membrane [127].
mTOR activity also can regulate lipid metabolism. Activation of mTOR may be essential for adipogenesis and adipocyte differentiation [229,230]. Application of rapamycin has been shown to significantly reduce expression of most adipocyte markers including peroxisome proliferators-activated receptor-γ (PPAR-γ), adipsin, adipocyte protein 2 (aP2), α-adducin (ADD1)/sterol regulatory element-binding protein 1(SREBP1c), and FAS [229]. Rapamycin also prevents intracellular lipid accumulation in 3T3-L1 and 3T3-F442A cells [229]. In addition, activation of mTOR inhibits lipolysis. For example, activation of mTORC1 signaling in 3T3-L1 adipocytes by ectopic expression of Rheb inhibits adipose triglyceride lipase and hormone-sensitive lipase and promotes intracellular accumulation of triglycerides. In contrast, inhibition of mTORC1 signaling by rapamycin or by knockdown of Raptor results in an increase in the expression of lipase and lipolysis [231] and leads to weight loss in gerbil, rats, and humans [228,232]. The ability of mTOR to regulate lipid metabolism has been associated with lipin, a magnesium-dependent phosphatidic acid phosphatase. Lipin plays a role in the regulation of fat cell differentiation and triglyceride synthesis [233]. Mutations in the Lpin1 gene result in hepatic steatosis in fld mice, a genetic model of lipodystrophy [234]. Lipin may function downstream of the peroxisome proliferator-activated receptor γ (PPARγ) coactivator 1alpha (PGC-1α) and selectively activates a PGC-1α mediated fatty acid oxidation and mitochondrial oxidative phosphorylation, suppresses lipogenesis, and lowers circulating lipid levels [87,235]. Activation of mTOR can increase the phosphorylation of lipin and mediate insulin induced lipin phosphorylation, resulting in the release of lipin from intracellular membranes [236].
In the central nervous system, mTOR also has a role in regulating food intake in the hypothalamus. mTOR is expressed ubiquitously in the nervous system. Phosphorylated mTOR and P70S6K has been shown to be expressed in the hippocampus, thalamus, cortex, and paraventricular (PVN) and arcuate (ARC) nuclei of the hypothalamus [237]. The ARC nuclei contain at least two populations of neurons, including orexigenic neurons that express both neuropeptide Y (NPY) and agouti-related peptide (AgRP) and anorexigenic neurons that express proopiomelanocortin (POMC) and cocaine- and amphetamine-regulated transcript (CART). Each of these agents is linked to the regulation of cellular energy homeostasis. mTOR and p70S6K exist in up to 90% of ARC NPY/AgRP neurons and in approximately 45% of ARC POMC/CART neurons. Central administration of leucine increases hypothalamic mTOR signaling and decreases food intake and body weight. The hormone leptin increases hypothalamic mTOR activity and inhibition of mTOR signaling has been associated with anorexia [237]. Leptin significantly results in the phosphorylation of p70S6K and p90 ribosomal S6 kinase 1 (RSK1) in mice placed on a low-fat diet and leads to hyperphagia, weight gain, and leptin resistance during diet-induced obesity [238]. Nutrient availability and insulin regulate leptin expression. Yet, mTORC1 also plays a role in leptin expression in adipose cells, since up-regulation of mTORC1 in 3T3-L1 adipocytes via stable expression of either constitutively active Rheb or dominant-negative AMP activated protein kinase (AMPK) results in a significant increase in leptin expression [239]. AMPK can phosphorylate tuberin (TSC2) and inhibit mTORC1 [240].

4.2. Acute Injury in the Nervous System

The PI 3-K, Akt, and mTOR pathways play a significant role during acute injury in the nervous system (Figure 2). For example, EPO activates the PI 3-K and Akt pathways to protect neurons, vascular cells, and immune cells during oxidative stress [80,155,156,159,167,169,174,175,182,183,187,200,241,242]. Agents that can increase the expression of Raptor are associated with neuroprotection during hypoxia in invertebrate models of stroke [243]. Furthermore, agents that increase activity of Akt, mTOR, and p70S6K also can reduce cerebral infarct size [244]. Activation of mTOR is necessary in primary cerebral microglia [80,184,204,206] and in neurons [25,147] to prevent apoptotic cell death during oxygen-glucose deprivation. Following acute spinal cord injury, an increase in mTOR expression and p70S6K activity also may be required for functional improvement [245]. Studies with bisperoxovanadium that can enhance the activities of Akt and mTOR have been demonstrated to reduce motor neuron death, increase tissue sparing, and reduce cavity formation after spinal cord injury in rats [246]. ATP administration that increases Akt, mTOR, and p70S6K signaling is accompanied by improved locomotor function following spinal cord injury [247]. mTOR in conjunction with other pathways such as signal transducers and activators of transcription (STAT) pathways can foster axonal regeneration [248]. For example, axonal regeneration is increased in adult retinal ganglion cells and in corticospinal neurons following injury paradigms with mTOR activation [249,250].
However, some conditions may require mTOR blockade to promote neuronal protection and autophagy in conjunction with increased activity of the PI 3-K and Akt axis [111]. Inhibition of mTOR and p70S6K activities also improves functional recovery in closed head injury models [251]. Blockade of mTOR has been shown to promote autophagy, inhibit mTOR-mediated inflammation, reduce neural tissue damage, and limit locomotor impairment following spinal cord injury [252]. Inhibition of PTEN (phosphatase and tensin homolog deleted on chromosome 10) that results in enhanced mTOR activity results in increased cerebral infarction [253]. Inhibition of mTOR signaling also prevents cerebral vasospasm and preserves endothelial cell function in animal models of subarachnoid hemorrhage [254].
During disorders of epilepsy that can be a recurrent acute disability, mTOR inhibition may be beneficial. Aberrant or significant mTOR activity is believed to interfere with normal brain function and lead to epilepsy. Inhibition of mTOR activity during kainate-induced epilepsy decreases neuronal cell death, neurogenesis, mossy fiber sprouting, and the development of spontaneous epilepsy [255]. Chronic hippocampal infusion of rapamycin that blocks mTOR signaling also limits mossy fiber sprouting in rat pilocarpine models of temporal lobe epilepsy [256]. mTOR signaling also is considered to be one mechanism for seizure disorders that occur in tuberous sclerosis (TS) [257]. Mutations of TSC1 and TSC2 that lead to hyperactive mTOR result in a high incidence of epilepsy [258]. Early inhibition of mTOR signaling in animal models of TS can prevent astrogliosis and neuronal dysfunction [259].

4.3. Chronic Neurodegeneration

Similar to disorders with acute nervous system injury, the temporal course and level of PI 3-K, Akt, and mTOR activation during chronic neurodegenerative disorders that can progress during aging can influence cellular survival and clinical outcome (Figure 2). During disorders such as AD, a minimum level of the PI 3-K, Akt, and mTOR pathway may be required. Since Aβ is toxic to cells [184,260], activation of the PI 3-K and Akt pathways has been shown to prevent Aβ toxicity [103,168,184,261263]. In regards to mTOR, blockade of mTOR activity may lead to neuronal atrophy in AD. Insufficiency of retinoblastoma tumor suppressor (RB1) inducible Coiled-Coil 1 (RB1CC1) has been observed in the brains of AD patients. In these patients, RB1CC1 appears to be necessary for neurite growth and to maintain mTOR signaling, but the reduced expression of RB1CC1 leads to reduced mTOR activity, neuronal apoptosis, and neuronal atrophy [264]. A decrease in mTOR activity in peripheral lymphocytes also appears to correlate with the progression of AD [265] and inhibition of mTOR activity has been shown to impair memory consolidation [266]. Loss of mTOR signaling also has been shown to impair long-term potentiation and synaptic plasticity in models of AD [267]. In addition, Aβ can block the activation of mTOR and p70S6K in neuroblastoma cells and in lymphocytes of patients with AD [268]. Activation of mTOR and p70S6K has been shown to prevent cell death during Aβ exposure in microglia, cells that are necessary for the removal of Aβ [184].
Additional studies provide further support for the premise that the degree of activity for the PI 3-K, Akt, and mTOR pathways may be an important factor for the treatment of neurodegenerative disorders, such as AD. In fact, some investigations suggest that inhibition of PI 3-K, Akt, and mTOR signaling may be necessary to achieve therapeutic benefit. For example, an increase in the phosphorylated level of Akt substrates, such as mTOR, GSK-3β, and tau protein have been observed in AD, suggesting that these substrates may promote AD progression [269]. Hyper-activation of PI 3-K and Akt associated with decreased calmodulin degradation in lymphoblasts from patients with AD also has been suggested as a potential detriment to cell survival [270]. p70S6K activation also has been associated with hyperphosphorylated tau formation and potential neurofibrillary accumulation in AD patients [271]. In addition, mTOR inhibition that can lead to autophagy in murine models of AD has been shown to improve memory and limit Aβ levels [93].
Inhibition of the PI 3-K, Akt, and mTOR pathway also may be necessary for the treatment of HD, an autosomal dominant disorder characterized by the degeneration of striatal GABAergic projecting neurons that result in involuntary movements and cognitive impairment. Activation of autophagy and the inhibition of mTOR are considered vital for the clearing of aggregate-prone proteins in disorders such as HD [272]. HD is the result of neuronal intracellular aggregates of huntingtin protein mutations that produce abnormally expanded polyglutamine in the N-terminal region of the huntingtin gene and lead to neuronal cell death. As a result, inhibition of mTOR signaling that can promote autophagy may represent a potential therapeutic strategy for HD. Blockade of mTOR activity has been demonstrated to enhance autophagic clearance of proteins with long polyglutamines and a polyalanine-expanded protein [273], attenuate huntingtin accumulation and cell death in cell models of HD, and protect against neurodegeneration in a fly model of HD [274]. Small molecular enhancers of rapamycin also have been shown to promote autophagy with both mTOR dependent and independent mechanisms to increase the clearance of a mutant huntingtin fragment in HD cell models [275]. The rapamycin analog CCI-779 also improves behavioral performance and decreases aggregate formation in a mouse model of HD [274]. Yet, some experimental models of HD suggest that inhibition of only mTORC1 may be insufficient to alter autophagy or huntingtin accumulation. The combined inhibition of mTORC1 and mTORC2 is required for autophagy and reductions in huntingtin accumulation, suggesting that multiple components of the mTOR pathway may modulate cell survival in HD [276]. Other studies support this premise to demonstrate that decreased activity of p70S6K protects against early decline in motor performance with beneficial effects on muscle function, but mutant huntingtin levels in the brain are not affected [277]. Prevention of neuronal demise by mTOR may work in concert with growth arrest and DNA damage protein 34 (GADD34). GADD34 leads to the dephosphorylation of TSC2 and induction of autophagy in cell models of HD with increased cell survival during GADD34 over-expression [27].
In PD, a movement disorder characterized by resting tremor, rigidity, and bradykinesia as a result of the loss of dopaminergic neurons in the substantia nigra, mTOR inactivation and promotion of autophagy may preserve dopaminergic neurons. During inhibition of mTOR and the activation of autophagy, the accumulation of toxic α-synuclein in transgenic mice is reduced and neurodegeneration is decreased in models of PD [278]. Yet, only portions of the PI 3-K, Akt, and mTOR pathway may be necessary for the treatment of PD. Treatment with rapamycin protects neurons that has been shown to be dependent upon Akt activation, but not associated with mTOR signaling [166]. In addition, activation of 4EBP1 during mTOR inhibition can suppress pathologic experimental phenotypes of PD including degeneration of dopaminergic neurons in Drosophila [279].
Other studies indicate that at least some activity of the mTOR pathway is required for neuronal protection in PD. Decreased mTOR activity may lead to neuronal injury in PD through pathways such as autophagy. Inhibition of mTOR signaling during oxidative stress leads to cell death in dopaminergic neurons by enhancing autophagic neuronal death [205]. The stress response protein REDD1 (RTP801), an inhibitor of mTORC1 activity [280], also is up-regulated in dopaminergic neurons in PD patients [281] and can inactivate Akt and mTOR. REDD1 is highly expressed in several cellular models of PD such as treatment with 6–hydroxydopamine (6-OHDA), MPTP, and rotenone [281]. REDD1 is regarded as a potential contributor to neuronal degeneration in PD, since gene silencing of REDD1 is neuroprotective against 6-OHDA [166]. Loss of mTOR activity and the chronic activation of the mTOR pathway 4EBP1 by leucine-rich repeat kinase 2 (LRRK2), a site for dominant mutations PD, is believed to alter protein translation and lead to the loss of dopaminergic neurons [282]. Interestingly, elevated levels of activity in the PI 3-K, Akt, and mTOR pathways can be detrimental in PD. Treatment with derivatives a dopamine, such as L-DOPA, lead to dopamine D1 receptor-mediated activation of mTORC1 resulting in dyskinesia [283].

5. Conclusions and Perspectives

Oxidative stress plays a significant role in the pathology of numerous disorders of the nervous system whether they are of acute or chronic origin during aging. Intimately tied to oxidative stress and the pathways that can determine cell survival are PI 3-K, Akt, and mTOR. The PI 3-K, Akt, and mTOR cascade is involved in the modulation of apoptosis, autophagy, disorders of cellular metabolism, acute nervous system injury, and chronic neurodegeneration. Targeting these pathways offers new inroads for the development of novel therapies that can either prevent or reverse the progression of diabetic complications in the nervous system and nervous system dysfunction. New studies suggest that consideration should be given for the modulation of individual components of the PI 3-K, Akt, and mTOR pathway, since neuronal protection in some cases such as those involving PD may be dependent upon one component such as Akt, but not other downstream components such as mTOR [166]. New work is focusing upon agents that can target different Akt classes with the alkyl-lysophospholipids and small molecule inhibitors of Akt. Alternate strategies that consider PDK1 and eIF4E are also under study. In other scenarios, such as with HD, broader aspects of the cascade may be required that can foster the reduction of toxic cellular aggregates [276]. In preclinical studies, modulation of the PI 3-K, Akt, and mTOR cascade can increase radiosensitivity against tumor cell growth and the vascular supply of tumors [284]. In addition, it appears that mTOR signaling components are required to maintain insulin signaling and prevent DM, since loss of p70S6K activity with the combined loss of the mTORC2 substrate Akt2 results in defects in both insulin action and β-cell function [227]. One must also take into account additional mechanisms that are tied to the PI 3-K, Akt, and mTOR pathways to promote neuroprotection and cellular energy homeostasis, such as GADD34 [27], forkhead transcription factors [110,177,178,285], Raf-mitogen activated protein kinases [286], and SIRT1 [1,3,215].
However, modulation of the PI 3-K, Akt, and mTOR cascade, especially during times of activation is not without its concerns. PI 3-K, Akt, and mTOR are cellular proliferative pathways that can promote aggressive tumor growth. As a result, the United States Food and Drug Administration has approved several rapamycin (sirolimus) and rapamycin derivative compounds (“rapalogs”) for the treatment of nervous system cancers that include subependymal giant cell astrocytoma associated with tuberous sclerosis (everolimus) and neuroendocrine pancreatic tumors (everolimus) [73,99]. Early clinical trials using everolimus for the treatment of advanced neuroendocrine tumors suggest that progression free survival can be improved [287]. Prevention of PI 3-K and Akt activation also can block medulloblastoma growth [288].
In addition to these considerations, parameters related to the temporal administration of agents that can modulate the PI 3-K, Akt, and mTOR signaling and the pathways of cell death that should be targeted appear to be critical. For example, recent phase III trials in relation to AD suggest that attempts to reduce Aβ toxicity in the brains of patients with AD through immunotherapy with Bapineuzumab or with Solanezumab (LY2062430) may have had limited clinical success as a result of not instituting therapy prior to significant clinical disease presentation or progression. In animal models of TS, early inhibition of mTOR signaling rather than late treatment has been shown to prevent astrogliosis, premature death, and seizures [259]. Early rather than late mTOR inhibition also can reduce plaques, tangles, and loss of cognition in murine models of AD [289]. Additional work suggests that the duration of activation of PI 3-K, Akt, and mTOR may be a significant factor for influencing cell survival. Inhibition of mTOR can impair long-term potentiation and synaptic plasticity in models of AD [267] and activation of mTOR signaling can block inflammatory cell death during Aβ exposure [184]. Yet, excessive activation of mTOR can lead to dyskinesia patients with PD [283].
In regards to cell death pathways, apoptosis and autophagy appear to have a complex relationship. During toxin exposure that involves methamphetamine, cell death proceeds through apoptosis and autophagy by inhibiting the disassociation of the Bcl-2/Beclin 1 complex [89], an anti-apoptotic protein that blocks autophagy through its inhibitory interaction with Beclin 1 [290]. Autophagy and apoptosis also can have opposing roles. Some studies report that progression of apoptosis may require the inhibition of autophagy [75,90,291]. In relation to nervous system disorders, it remains unclear when pathways such as autophagy may be beneficial. During oxidative stress, autophagy can lead to cell death in cerebral astrocytes [82], in cortical neurons [83], and in spinal cord motor neurons [84]. Activation of the PI 3-K, Akt, mTOR pathway that can block autophagy may be necessary to protect against AGEs and complications of DM with atherosclerosis [97], prevent spinal cord injury [247], and maintain synaptic plasticity [267]. However, autophagy may be necessary during inhibition of mTOR signaling to improve cognitive function, limit Aβ toxicity [93], and clear mutant huntingtin in HD [275]. Over the next several years, further studies that can continue to unravel the cellular pathways governed by the PI 3-K, Akt, and mTOR cascade and the significant biological role this cascade holds in specific disorders of the nervous system will offer the greatest potential to target these pathways for robust treatments against neurodegenerative disorders.

Acknowledgments

This research was supported by the following grants to Kenneth Maiese: American Diabetes Association, American Heart Association (National), Bugher Foundation Award, Janssen Neuroscience Award, LEARN Foundation Award, NIH NIEHS, NIH NIA, NIH NINDS, and NIH ARRA.
  • Conflict of Interest The authors declare no conflict of interest.

References

  1. Chong, Z.Z.; Shang, Y.C.; Wang, S.; Maiese, K. SIRT1: New avenues of discovery for disorders of oxidative stress. Expert Opin. Ther. Targets 2012, 16, 167–178. [Google Scholar]
  2. Maiese, K.; Chong, Z.Z.; Hou, J.; Shang, Y.C. Oxidative stress: Biomarkers and novel therapeutic pathways. Exp. Gerontol 2010, 45, 217–234. [Google Scholar]
  3. Maiese, K.; Chong, Z.Z.; Shang, Y.C.; Wang, S. Translating cell survival and cell longevity into treatment strategies with SIRT1. Rom. J. Morphol. Embryol 2011, 52, 1173–1185. [Google Scholar]
  4. Jayaram, H.N.; Kusumanchi, P.; Yalowitz, J.A. NMNAT expression and its relation to NAD metabolism. Curr. Med. Chem 2011, 18, 1962–1972. [Google Scholar]
  5. Yang, H.; Jin, X.; Kei Lam, C.W.; Yan, S.K. Oxidative stress and diabetes mellitus. Clin. Chem. Lab. Med 2011, 49, 1773–1782. [Google Scholar]
  6. Maiese, K.; Chong, Z.Z.; Hou, J.; Shang, Y.C. The vitamin nicotinamide: Translating nutrition into clinical care. Molecules 2009, 14, 3446–3485. [Google Scholar]
  7. Maiese, K.; Hou, J.; Chong, Z.Z.; Shang, Y.C. Erythropoietin, forkhead proteins, and oxidative injury: biomarkers and biology. TheScientificWorldJournal 2009, 9, 1072–1104. [Google Scholar]
  8. Muley, M.M.; Thakare, V.N.; Patil, R.R.; Kshirsagar, A.D.; Naik, S.R. Silymarin improves the behavioural, biochemical and histoarchitecture alterations in focal ischemic rats: A comparative evaluation with piracetam and protocatachuic acid. Pharmacol. Biochem. Behav 2012, 102, 286–293. [Google Scholar]
  9. Sun, L.; Gu, L.; Wang, S.; Yuan, J.; Yang, H.; Zhu, J.; Zhang, H. N-acetylcysteine protects against apoptosis through modulation of group I metabotropic glutamate receptor activity. PLoS One 2012, 7, e32503. [Google Scholar]
  10. Suzen, S.; Cihaner, S.S.; Coban, T. Synthesis and comparison of antioxidant properties of indole-based melatonin analogue indole amino acid derivatives. Chem. Biol. Drug Des 2012, 79, 76–83. [Google Scholar]
  11. Tupe, R.S.; Tupe, S.G.; Agte, V.V. Dietary nicotinic acid supplementation improves hepatic zinc uptake and offers hepatoprotection against oxidative damage. Br. J. Nutr 2011, 25, 1–9. [Google Scholar]
  12. Wong, D.Z.; Kadir, H.A.; Lee, C.L.; Goh, B.H. Neuroprotective properties of Loranthus parasiticus aqueous fraction against oxidative stress-induced damage in NG108–15 cells. J. Nat. Med 2012, 66, 544–551. [Google Scholar]
  13. Zhang, G.; Zhao, Z.; Gao, L.; Deng, J.; Wang, B.; Xu, D.; Liu, B.; Qu, Y.; Yu, J.; Li, J.; et al. Gypenoside attenuates white matter lesions induced by chronic cerebral hypoperfusion in rats. Pharmacol. Biochem. Behav 2011, 99, 42–51. [Google Scholar]
  14. Cacciatore, I.; Baldassarre, L.; Fornasari, E.; Mollica, A.; Pinnen, F. Recent advances in the treatment of neurodegenerative diseases based on GSH delivery systems. Oxid. Med. Cell. Longevity 2012, 2012, 240146. [Google Scholar]
  15. Escobar, J.; Pereda, J.; Lopez-Rodas, G.; Sastre, J. Redox signaling and histone acetylation in acute pancreatitis. Free Radic. Biol. Med 2012, 52, 819–837. [Google Scholar]
  16. Lange, C.A.; Bainbridge, J.W. Oxygen sensing in retinal health and disease. Ophthalmologica 2012, 227, 115–131. [Google Scholar]
  17. Wang, J.; Sun, P.; Bao, Y.; Dou, B.; Song, D.; Li, Y. Vitamin E renders protection to PC12 cells against oxidative damage and apoptosis induced by single-walled carbon nanotubes. Toxicol. In Vitro 2012, 26, 32–41. [Google Scholar]
  18. Chong, Z.Z.; Li, F.; Maiese, K. Oxidative stress in the brain: Novel cellular targets that govern survival during neurodegenerative disease. Prog. Neurobiol 2005, 75, 207–246. [Google Scholar]
  19. Li, R.P.; Wang, Z.Z.; Sun, M.X.; Hou, X.L.; Sun, Y.; Deng, Z.F.; Xiao, K. Polydatin protects learning and memory impairments in a rat model of vascular dementia. Phytomedicine 2012, 19, 677–681. [Google Scholar]
  20. Maiese, K.; Chong, Z.Z.; Hou, J.; Shang, Y.C. New strategies for Alzheimer’s disease and cognitive impairment. Oxid. Med. Cell. Longevity 2009, 2, 279–289. [Google Scholar]
  21. Bajda, M.; Guzior, N.; Ignasik, M.; Malawska, B. Multi-target-directed ligands in Alzheimer’s disease treatment. Curr. Med. Chem 2011, 18, 4949–4975. [Google Scholar]
  22. Chong, Z.Z.; Li, F.; Maiese, K. Stress in the brain: Novel cellular mechanisms of injury linked to Alzheimer’s disease. Brain Res. Brain Res. Rev 2005, 49, 1–21. [Google Scholar]
  23. Maiese, K.; Chong, Z.; Li, F. Reducing Oxidative Stress and Enhancing Neurovascular Longevity during Diabetes Mellitus. In Neurovascular Medicine: Pursuing Cellular Longevity for Healthy Aging; Maiese, K., Ed.; Oxford University Press: New York, NY, USA, 2009; pp. 540–564. [Google Scholar]
  24. Srivastava, S.; Haigis, M.C. Role of sirtuins and calorie restriction in neuroprotection: Implications in Alzheimer’s and Parkinson’s diseases. Curr. Pharm. Des 2011, 17, 3418–3433. [Google Scholar]
  25. Chong, Z.Z.; Shang, Y.C.; Zhang, L.; Wang, S.; Maiese, K. Mammalian target of rapamycin: Hitting the bull’s-eye for neurological disorders. Oxid. Med. Cell. Longevity 2010, 3, 374–391. [Google Scholar]
  26. Enz, R. Metabotropic glutamate receptors and interacting proteins: Evolving drug targets. Curr. Drug Targets 2012, 13, 145–156. [Google Scholar]
  27. Hyrskyluoto, A.; Reijonen, S.; Kivinen, J.; Lindholm, D.; Korhonen, L. GADD34 mediates cytoprotective autophagy in mutant huntingtin expressing cells via the mTOR pathway. Exp. Cell Res 2012, 318, 33–42. [Google Scholar]
  28. Maiese, K.; Chong, Z.Z.; Li, F. Driving cellular plasticity and survival through the signal transduction pathways of metabotropic glutamate receptors. Curr. Neurovasc. Res 2005, 2, 425–446. [Google Scholar]
  29. Nagley, P.; Higgins, G.C.; Atkin, J.D.; Beart, P.M. Multifaceted deaths orchestrated by mitochondria in neurones. Biochim. Biophys. Acta 2010, 1802, 167–185. [Google Scholar]
  30. Vidal, R.L.; Figueroa, A.; Court, F.A.; Thielen, P.; Molina, C.; Wirth, C.; Caballero, B.; Kiffin, R.; Segura-Aguilar, J.; Cuervo, A.M.; et al. Targeting the UPR transcription factor XBP1 protects against Huntington’s disease through the regulation of FoxO1 and autophagy. Hum. Mol. Genet 2012, 21, 2245–2262. [Google Scholar]
  31. Yang, D.; Wang, C.E.; Zhao, B.; Li, W.; Ouyang, Z.; Liu, Z.; Yang, H.; Fan, P.; O’Neill, A.; Gu, W.; et al. Expression of Huntington’s disease protein results in apoptotic neurons in the brains of cloned transgenic pigs. Hum. Mol. Genet 2010, 19, 3983–3994. [Google Scholar]
  32. Al Sweidi, S.; Sanchez, M.G.; Bourque, M.; Morissette, M.; Dluzen, D.; Di Paolo, T. Oestrogen receptors and signalling pathways: Implications for neuroprotective effects of sex steroids in Parkinson’s disease. J. Neuroendocrinol 2012, 24, 48–61. [Google Scholar]
  33. Asaithambi, A.; Kanthasamy, A.; Saminathan, H.; Anantharam, V.; Kanthasamy, A.G. Protein kinase D1 (PKD1) activation mediates a compensatory protective response during early stages of oxidative stress-induced neuronal degeneration. Mol. Neurodegener. 2011, 6. [Google Scholar] [CrossRef]
  34. Das, F.; Dey, N.; Venkatesan, B.; Kasinath, B.S.; Ghosh-Choudhury, N.; Choudhury, G.G. High glucose upregulation of early-onset Parkinson’s disease protein DJ-1 integrates the PRAS40/TORC1 axis to mesangial cell hypertrophy. Cell. Signal 2011, 23, 1311–1319. [Google Scholar]
  35. Khan, M.M.; Ahmad, A.; Ishrat, T.; Khan, M.B.; Hoda, M.N.; Khuwaja, G.; Raza, S.S.; Khan, A.; Javed, H.; Vaibhav, K.; et al. Resveratrol attenuates 6-hydroxydopamine-induced oxidative damage and dopamine depletion in rat model of Parkinson’s disease. Brain Res 2010, 1328, 139–151. [Google Scholar]
  36. Kook, Y.H.; Ka, M.; Um, M. Neuroprotective cytokines repress PUMA induction in the 1-methyl-4-phenylpyridinium (MPP(+)) model of Parkinson’s disease. Biochem. Biophys. Res. Commun 2011, 411, 370–374. [Google Scholar]
  37. L’Episcopo, F.; Tirolo, C.; Testa, N.; Caniglia, S.; Morale, M.C.; Deleidi, M.; Serapide, M.F.; Pluchino, S.; Marchetti, B. Plasticity of subventricular zone neuroprogenitors in MPTP (1-Methyl-4-Phenyl-1,2,3,6-Tetrahydropyridine) mouse model of Parkinson’s disease involves cross talk between inflammatory and Wnt/beta-Catenin signaling pathways: Functional consequences for neuroprotection and repair. J. Neurosci 2012, 32, 2062–2085. [Google Scholar]
  38. Maiese, K.; Chong, Z.Z.; Shang, Y.C.; Hou, J. Therapeutic promise and principles: Metabotropic glutamate receptors. Oxid. Med. Cell. Longevity 2008, 1, 1–14. [Google Scholar]
  39. Maiese, K.; Chong, Z.Z.; Shang, Y.C.; Wang, S. Targeting disease through novel pathways of apoptosis and autophagy. Expert Opin. Ther. Targets 2012, in press. [Google Scholar]
  40. Banach, M.; Piskorska, B.; Czuczwar, S.; Borowicz, K. Nitric oxide, epileptic seizures, and action of antiepileptic drugs. CNS Neurol. Disord. Drug Targets 2011, 10, 808–819. [Google Scholar]
  41. O’Dell, C.M.; Das, A.; Wallace, G.t.; Ray, S.K.; Banik, N.L. Understanding the basic mechanisms underlying seizures in mesial temporal lobe epilepsy and possible therapeutic targets: A review. J. Neurosci. Res 2012, 90, 913–924. [Google Scholar]
  42. Sensi, S.L.; Paoletti, P.; Koh, J.Y.; Aizenman, E.; Bush, A.I.; Hershfinkel, M. The neurophysiology and pathology of brain zinc. J. Neurosci 2011, 31, 16076–16085. [Google Scholar]
  43. Talos, D.M.; Sun, H.; Zhou, X.; Fitzgerald, E.C.; Jackson, M.C.; Klein, P.M.; Lan, V.J.; Joseph, A.; Jensen, F.E. The Interaction between early life epilepsy and autistic-like behavioral consequences: A role for the mammalian target of rapamycin (mTOR) pathway. PLoS One 2012, 7, e35885. [Google Scholar]
  44. Beard, R.S., Jr; Reynolds, J.J.; Bearden, S.E. Metabotropic glutamate receptor 5 mediates phosphorylation of vascular endothelial cadherin and nuclear localization of beta-catenin in response to homocysteine. Vascul. Pharmacol 2012, 56, 159–167. [Google Scholar]
  45. Bonilla, C.; Zurita, M.; Otero, L.; Aguayo, C.; Rico, M.A.; Vaquero, J. The severity of brain damage determines bone marrow stromal cell therapy efficacy in a traumatic brain injury model. J. Trauma Acute Care Surg 2012, 72, 1203–1212. [Google Scholar]
  46. Duris, K.; Manaenko, A.; Suzuki, H.; Rolland, W.B.; Krafft, P.R.; Zhang, J.H. α7 Nicotinic acetylcholine receptor agonist PNU-282987 attenuates early brain injury in a perforation model of subarachnoid hemorrhage in rats. Stroke 2011, 42, 3530–3536. [Google Scholar]
  47. Goffus, A.M.; Anderson, G.D.; Hoane, M. Sustained delivery of nicotinamide limits cortical injury and improves functional recovery following traumatic brain injury. Oxid. Med. Cell. Longevity 2010, 3, 145–152. [Google Scholar]
  48. Guo, J.M.; Dong, W.Z.; Liu, A.J.; Cheng, M.H.; Su, D.F. Nicotinamide postpones stroke in stroke-prone spontaneously hypertensive rats. CNS Neurosci. Ther 2012, 18, 267–268. [Google Scholar]
  49. Lanfranconi, S.; Locatelli, F.; Corti, S.; Candelise, L.; Comi, G.P.; Baron, P.L.; Strazzer, S.; Bresolin, N.; Bersano, A. Growth factors in ischemic stroke. J. Cell. Mol. Med 2011, 15, 1645–1687. [Google Scholar]
  50. Ma, Y.; Qu, Y.; Fei, Z. Vascular endothelial growth factor in cerebral ischemia. J. Neurosci. Res 2011, 89, 969–978. [Google Scholar]
  51. Maiese, K.; Chong, Z.Z.; Hou, J.; Shang, Y.C. Erythropoietin and oxidative stress. Curr. Neurovasc. Res 2008, 5, 125–142. [Google Scholar]
  52. Maiese, K.; Chong, Z.Z.; Li, F.; Shang, Y.C. Erythropoietin: Elucidating new cellular targets that broaden therapeutic strategies. Prog. Neurobiol 2008, 85, 194–213. [Google Scholar]
  53. Raza, S.S.; Khan, M.M.; Ashafaq, M.; Ahmad, A.; Khuwaja, G.; Khan, A.; Siddiqui, M.S.; Safhi, M.M.; Islam, F. Silymarin protects neurons from oxidative stress associated damages in focal cerebral ischemia: A behavioral, biochemical and immunohistological study in Wistar rats. J. Neurol. Sci 2011, 309, 45–54. [Google Scholar]
  54. Zhao, L.D.; Wang, J.H.; Jin, G.R.; Zhao, Y.; Zhang, H.J. Neuroprotective effect of Buyang Huanwu decoction against focal cerebral ischemia/reperfusion injury in rats—Time window and mechanism. J. Ethnopharmacol 2012, 140, 339–344. [Google Scholar]
  55. Chong, Z.Z.; Maiese, K. The Src homology 2 domain tyrosine phosphatases SHP-1 and SHP-2: Diversified control of cell growth, inflammation, and injury. Histol. Histopathol 2007, 22, 1251–1267. [Google Scholar]
  56. Kato, S.; Aoyama, M.; Kakita, H.; Hida, H.; Kato, I.; Ito, T.; Goto, T.; Hussein, M.H.; Sawamoto, K.; Togari, H.; et al. Endogenous erythropoietin from astrocyte protects the oligodendrocyte precursor cell against hypoxic and reoxygenation injury. J. Neurosci. Res 2011, 89, 1566–1574. [Google Scholar]
  57. Kigerl, K.A.; Ankeny, D.P.; Garg, S.K.; Wei, P.; Guan, Z.; Lai, W.; McTigue, D.M.; Banerjee, R.; Popovich, P.G. System x(c)(−) regulates microglia and macrophage glutamate excitotoxicity in vivo. Exp. Neurol 2012, 233, 333–341. [Google Scholar]
  58. L’Episcopo, F.; Tirolo, C.; Testa, N.; Caniglia, S.; Morale, M.C.; Cossetti, C.; D’Adamo, P.; Zardini, E.; Andreoni, L.; Ihekwaba, A.E.; et al. Reactive astrocytes and Wnt/beta-catenin signaling link nigrostriatal injury to repair in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine model of Parkinson’s disease. Neurobiol. Dis 2011, 41, 508–527. [Google Scholar]
  59. Shang, Y.C.; Chong, Z.Z.; Hou, J.; Maiese, K. FoxO3a governs early microglial proliferation and employs mitochondrial depolarization with caspase 3, 8, and 9 cleavage during oxidant induced apoptosis. Curr. Neurovasc. Res 2009, 6, 223–238. [Google Scholar]
  60. Shang, Y.C.; Chong, Z.Z.; Hou, J.; Maiese, K. Wnt1, FoxO3a, and NF-kappaB oversee microglial integrity and activation during oxidant stress. Cell. Signal 2010, 22, 1317–1329. [Google Scholar]
  61. Jiang, Y.L.; Ning, Y.; Ma, X.L.; Liu, Y.Y.; Wang, Y.; Zhang, Z.; Shan, C.X.; Xu, Y.D.; Yin, L.M.; Yang, Y.Q. Alteration of the proteome profile of the pancreas in diabetic rats induced by streptozotocin. Int. J. Mol. Med 2011, 28, 153–160. [Google Scholar]
  62. John, C.M.; Ramasamy, R.; Al Naqeeb, G.; Dhiab Al-Nuaimi, A.H.; Adam, A. Enhanced CD4+CD25+ regulatory T cells with splenic proliferation and protection against oxidative stress by nicotinamide in gestational diabetes. Curr. Med. Chem. 2012, in press. [Google Scholar]
  63. Kurban, S.; Mehmetoglu, I.; Yerlikaya, H.F.; Gonen, S.; Erdem, S. Effect of chronic regular exercise on serum ischemia-modified albumin levels and oxidative stress in type 2 diabetes mellitus. Endocr. Res 2011, 36, 116–123. [Google Scholar]
  64. Lee, Y.; Hong, Y.; Lee, S.R.; Chang, K.T. Autophagy contributes to retardation of cardiac growth in diabetic rats. Lab. Anim. Res 2012, 28, 99–107. [Google Scholar]
  65. Liu, Z.; Stanojevic, V.; Brindamour, L.J.; Habener, J.F. GLP1–derived nonapeptide GLP1(28–36)amide protects pancreatic beta-cells from glucolipotoxicity. J. Endocrinol 2012, 213, 143–154. [Google Scholar]
  66. Maiese, K.; Chong, Z.Z.; Shang, Y.C.; Hou, J. Novel avenues of drug discovery and biomarkers for diabetes mellitus. J. Clin. Pharmacol 2011, 51, 128–152. [Google Scholar]
  67. Maiese, K.; Morhan, S.D.; Chong, Z.Z. Oxidative stress biology and cell injury during type 1 and type 2 diabetes mellitus. Curr. Neurovasc. Res 2007, 4, 63–71. [Google Scholar]
  68. Maiese, K.; Shang, Y.C.; Chong, Z.Z.; Hou, J. Diabetes mellitus: Channeling care through cellular discovery. Curr. Neurovasc. Res 2010, 7, 59–64. [Google Scholar]
  69. Pang, J.; Xi, C.; Dai, Y.; Gong, H.; Zhang, T.M. Altered expression of base excision repair genes in response to high glucose-induced oxidative stress in HepG2 hepatocytes. Med. Sci. Monit. 2012, 18, BR281–285. [Google Scholar]
  70. Tang, L.; Zhang, Y.; Jiang, Y.; Willard, L.; Ortiz, E.; Wark, L.; Medeiros, D.; Lin, D. Dietary wolfberry ameliorates retinal structure abnormalities in db/db mice at the early stage of diabetes. Exp. Biol. Med. (Maywood) 2011, 236, 1051–1063. [Google Scholar]
  71. Zengi, A.; Ercan, G.; Caglayan, O.; Tamsel, S.; Karadeniz, M.; Simsir, I.; Harman, E.; Kahraman, C.; Orman, M.; Cetinkalp, S.; et al. Increased oxidative DNA damage in lean normoglycemic offspring of type 2 diabetic patients. Exp. Clin. Endocrinol. Diabetes 2011, 119, 467–471. [Google Scholar]
  72. Acquaah-Mensah, G.K.; Taylor, R.C.; Bhave, S.V. PACAP interactions in the mouse brain: Implications for behavioral and other disorders. Gene 2012, 491, 224–231. [Google Scholar]
  73. Benjamin, D.; Colombi, M.; Moroni, C.; Hall, M.N. Rapamycin passes the torch: A new generation of mTOR inhibitors. Nat. Rev. Drug Discov 2011, 10, 868–880. [Google Scholar]
  74. Hwang, S.K.; Kim, H.H. The functions of mTOR in ischemic diseases. BMB Rep 2011, 44, 506–511. [Google Scholar]
  75. Maiese, K. The many facets of cell injury: Angiogenesis to autophagy. Curr. Neurovasc. Res 2012, 9, 83–84. [Google Scholar]
  76. Broe, M.; Shepherd, C.E.; Milward, E.A.; Halliday, G.M. Relationship between DNA fragmentation, morphological changes and neuronal loss in Alzheimer’s disease and dementia with Lewy bodies. Acta Neuropathol 2001, 101, 616–624. [Google Scholar]
  77. Louneva, N.; Cohen, J.W.; Han, L.Y.; Talbot, K.; Wilson, R.S.; Bennett, D.A.; Trojanowski, J.Q.; Arnold, S.E. Caspase-3 is enriched in postsynaptic densities and increased in Alzheimer’s disease. Am. J. Pathol 2008, 173, 1488–1495. [Google Scholar]
  78. Grammas, P.; Tripathy, D.; Sanchez, A.; Yin, X.; Luo, J. Brain microvasculature and hypoxia-related proteins in Alzheimer’s disease. Int. J. Clin. Exp. Pathol 2011, 4, 616–627. [Google Scholar]
  79. Tatton, N.A. Increased caspase 3 and Bax immunoreactivity accompany nuclear GAPDH translocation and neuronal apoptosis in Parkinson’s disease. Exp. Neurol 2000, 166, 29–43. [Google Scholar]
  80. Shang, Y.C.; Chong, Z.Z.; Wang, S.; Maiese, K. Erythropoietin and Wnt1 govern pathways of mTOR, Apaf-1, and XIAP in inflammatory microglia. Curr. Neurovasc. Res 2011, 8, 270–285. [Google Scholar]
  81. Zhou, X.; Wang, L.; Wang, M.; Xu, L.; Yu, L.; Fang, T.; Wu, M. Emodin-induced microglial apoptosis is associated with TRB3 induction. Immunopharmacol. Immunotoxicol 2011, 33, 594–602. [Google Scholar]
  82. Qin, A.P.; Liu, C.F.; Qin, Y.Y.; Hong, L.Z.; Xu, M.; Yang, L.; Liu, J.; Qin, Z.H.; Zhang, H.L. Autophagy was activated in injured astrocytes and mildly decreased cell survival following glucose and oxygen deprivation and focal cerebral ischemia. Autophagy 2010, 6, 738–753. [Google Scholar]
  83. Wang, J.Y.; Xia, Q.; Chu, K.T.; Pan, J.; Sun, L.N.; Zeng, B.; Zhu, Y.J.; Wang, Q.; Wang, K.; Luo, B.Y. Severe global cerebral ischemia-induced programmed necrosis of hippocampal CA1 neurons in rat is prevented by 3-methyladenine: A widely used inhibitor of autophagy. J. Neuropathol. Exp. Neurol 2011, 70, 314–322. [Google Scholar]
  84. Baba, H.; Sakurai, M.; Abe, K.; Tominaga, R. Autophagy-mediated stress response in motor neuron after transient ischemia in rabbits. J. Vasc. Surg 2009, 50, 381–387. [Google Scholar]
  85. Canu, N.; Tufi, R.; Serafino, A.L.; Amadoro, G.; Ciotti, M.T.; Calissano, P. Role of the autophagic-lysosomal system on low potassium-induced apoptosis in cultured cerebellar granule cells. J. Neurochem 2005, 92, 1228–1242. [Google Scholar] [Green Version]
  86. Xue, L.; Fletcher, G.C.; Tolkovsky, A.M. Autophagy is activated by apoptotic signalling in sympathetic neurons: An alternative mechanism of death execution. Mol. Cell. Neurosci 1999, 14, 180–198. [Google Scholar]
  87. Chong, Z.Z.; Maiese, K. Mammalian target of rapamycin signaling in diabetic cardiovascular disease. Cardiovasc. Diabetol. 2012, 11. [Google Scholar] [CrossRef]
  88. Troy, C.M.; Akpan, N.; Jean, Y.Y. Regulation of caspases in the nervous system implications for functions in health and disease. Prog. Mol. Biol. Transl. Sci 2011, 99, 265–305. [Google Scholar]
  89. Nopparat, C.; Porter, J.E.; Ebadi, M.; Govitrapong, P. The mechanism for the neuroprotective effect of melatonin against methamphetamine-induced autophagy. J. Pineal. Res 2010, 49, 382–389. [Google Scholar]
  90. Luo, S.; Rubinsztein, D.C. Apoptosis blocks Beclin 1–dependent autophagosome synthesis: An effect rescued by Bcl-xL. Cell Death Differ 2010, 17, 268–277. [Google Scholar]
  91. Wang, S.; Chong, Z.Z.; Shang, Y.C.; Maiese, K. WISP1 (CCN4) autoregulates its expression and nuclear trafficking of beta-catenin during oxidant stress with limited effects upon neuronal autophagy. Curr. Neurovasc. Res 2012, 9, 89–99. [Google Scholar]
  92. Spencer, B.; Potkar, R.; Trejo, M.; Rockenstein, E.; Patrick, C.; Gindi, R.; Adame, A.; Wyss-Coray, T.; Masliah, E. Beclin 1 gene transfer activates autophagy and ameliorates the neurodegenerative pathology in alpha-synuclein models of Parkinson’s and Lewy body diseases. J. Neurosci 2009, 29, 13578–13588. [Google Scholar]
  93. Spilman, P.; Podlutskaya, N.; Hart, M.J.; Debnath, J.; Gorostiza, O.; Bredesen, D.; Richardson, A.; Strong, R.; Galvan, V. Inhibition of mTOR by rapamycin abolishes cognitive deficits and reduces amyloid-beta levels in a mouse model of Alzheimer’s disease. PLoS One 2010, 5, e9979. [Google Scholar]
  94. Xilouri, M.; Vogiatzi, T.; Vekrellis, K.; Park, D.; Stefanis, L. Abberant alpha-synuclein confers toxicity to neurons in part through inhibition of chaperone-mediated autophagy. PLoS One 2009, 4, e5515. [Google Scholar]
  95. Liu, Y.; Shi, S.; Gu, Z.; Du, Y.; Liu, M.; Yan, S.; Gao, J.; Li, J.; Shao, Y.; Zhong, W.; et al. Impaired autophagic function in rat islets with aging. Age (Dordr) 2012. [Google Scholar] [CrossRef]
  96. He, C.; Bassik, M.C.; Moresi, V.; Sun, K.; Wei, Y.; Zou, Z.; An, Z.; Loh, J.; Fisher, J.; Sun, Q.; et al. Exercise-induced BCL2-regulated autophagy is required for muscle glucose homeostasis. Nature 2012, 481, 511–515. [Google Scholar]
  97. Hu, P.; Lai, D.; Lu, P.; Gao, J.; He, H. ERK and Akt signaling pathways are involved in advanced glycation end product-induced autophagy in rat vascular smooth muscle cells. Int. J. Mol. Med 2012, 29, 613–618. [Google Scholar]
  98. Martino, L.; Masini, M.; Novelli, M.; Beffy, P.; Bugliani, M.; Marselli, L.; Masiello, P.; Marchetti, P.; De Tata, V. Palmitate activates autophagy in INS-1E beta-cells and in isolated rat and human pancreatic islets. PLoS One 2012, 7, e36188. [Google Scholar]
  99. Chong, Z.Z.; Shang, Y.C.; Wang, S.; Maiese, K. Shedding new light on neurodegenerative diseases through the mammalian target of rapamycin. Prog. Neurobiol. 2012, in press. [Google Scholar]
  100. Maiese, K.; Chong, Z.Z.; Shang, Y.C.; Wang, S. Erythropoietin: New directions for the nervous system. Int. J. Mol. Sci 2012, 13, 11102–11129. [Google Scholar]
  101. Aksu, U.; Demirci, C.; Ince, C. The pathogenesis of acute kidney injury and the toxic triangle of oxygen, reactive oxygen species and nitric oxide. Contrib. Nephrol 2011, 174, 119–128. [Google Scholar]
  102. Balan, V.; Miller, G.S.; Kaplun, L.; Balan, K.; Chong, Z.Z.; Li, F.; Kaplun, A.; VanBerkum, M.F.; Arking, R.; Freeman, D.C.; et al. Life span extension and neuronal cell protection by Drosophila nicotinamidase. J. Biol. Chem 2008, 283, 27810–27819. [Google Scholar]
  103. Chong, Z.Z.; Li, F.; Maiese, K. Cellular demise and inflammatory microglial activation during beta-amyloid toxicity are governed by Wnt1 and canonical signaling pathways. Cell. Signal 2007, 19, 1150–1162. [Google Scholar]
  104. Bailey, T.J.; Fossum, S.L.; Fimbel, S.M.; Montgomery, J.E.; Hyde, D.R. The inhibitor of phagocytosis, O-phospho-l-serine, suppresses Muller glia proliferation and cone cell regeneration in the light-damaged zebrafish retina. Exp. Eye Res 2010, 91, 601–612. [Google Scholar]
  105. Chong, Z.Z.; Kang, J.; Li, F.; Maiese, K. mGluRI targets microglial activation and selectively prevents neuronal cell engulfment through Akt and Caspase dependent pathways. Curr. Neurovasc. Res 2005, 2, 197–211. [Google Scholar]
  106. Kang, J.Q.; Chong, Z.Z.; Maiese, K. Critical role for Akt1 in the modulation of apoptotic phosphatidylserine exposure and microglial activation. Mol. Pharmacol 2003, 64, 557–569. [Google Scholar]
  107. Kang, J.Q.; Chong, Z.Z.; Maiese, K. Akt1 protects against inflammatory microglial activation through maintenance of membrane asymmetry and modulation of cysteine protease activity. J. Neurosci. Res 2003, 74, 37–51. [Google Scholar]
  108. Shang, Y.C.; Chong, Z.Z.; Hou, J.; Maiese, K. The forkhead transcription factor FoxO3a controls microglial inflammatory activation and eventual apoptotic injury through caspase 3. Curr. Neurovasc. Res 2009, 6, 20–31. [Google Scholar]
  109. Koh, P.O. Nicotinamide attenuates the decrease of astrocytic phosphoprotein PEA-15 in focal cerebral ischemic injury. J. Vet. Med. Sci 2012, 74, 377–380. [Google Scholar]
  110. Maiese, K.; Chong, Z.Z.; Shang, Y.C. “Sly as a FOXO”: New paths with Forkhead signaling in the brain. Curr. Neurovasc. Res 2007, 4, 295–302. [Google Scholar]
  111. Balduini, W.; Carloni, S.; Buonocore, G. Autophagy in hypoxia-ischemia induced brain injury. J. Matern. Fetal Neonatal Med 2012, 25, 30–34. [Google Scholar]
  112. Deretic, V.; Jiang, S.; Dupont, N. Autophagy intersections with conventional and unconventional secretion in tissue development, remodeling and inflammation. Trends Cell Biol 2012, 22, 397–406. [Google Scholar]
  113. Silva, D.F.; Esteves, A.R.; Oliveira, C.R.; Cardoso, S.M. Mitochondria: The common upstream driver of amyloid-beta and tau pathology in Alzheimer’s disease. Curr. Alzheimer Res 2011, 8, 563–572. [Google Scholar]
  114. Scott, R.C.; Juhasz, G.; Neufeld, T.P. Direct induction of autophagy by Atg1 inhibits cell growth and induces apoptotic cell death. Curr. Biol 2007, 17, 1–11. [Google Scholar]
  115. Kabeya, Y.; Kamada, Y.; Baba, M.; Takikawa, H.; Sasaki, M.; Ohsumi, Y. Atg17 functions in cooperation with Atg1 and Atg13 in yeast autophagy. Mol. Biol. Cell 2005, 16, 2544–2553. [Google Scholar]
  116. Kamada, Y.; Funakoshi, T.; Shintani, T.; Nagano, K.; Ohsumi, M.; Ohsumi, Y. Tor-mediated induction of autophagy via an Apg1 protein kinase complex. J Cell Biol 2000, 150, 1507–1513. [Google Scholar]
  117. Heitman, J.; Movva, N.R.; Hall, M.N. Targets for cell cycle arrest by the immunosuppressant rapamycin in yeast. Science 1991, 253, 905–909. [Google Scholar]
  118. Brown, E.J.; Albers, M.W.; Shin, T.B.; Ichikawa, K.; Keith, C.T.; Lane, W.S.; Schreiber, S.L. A mammalian protein targeted by G1–arresting rapamycin-receptor complex. Nature 1994, 369, 756–758. [Google Scholar]
  119. Kuroyanagi, H.; Yan, J.; Seki, N.; Yamanouchi, Y.; Suzuki, Y.; Takano, T.; Muramatsu, M.; Shirasawa, T. Human ULK1, a novel serine/threonine kinase related to UNC-51 kinase of Caenorhabditis elegans: cDNA cloning, expression, and chromosomal assignment. Genomics 1998, 51, 76–85. [Google Scholar]
  120. Yan, J.; Kuroyanagi, H.; Kuroiwa, A.; Matsuda, Y.; Tokumitsu, H.; Tomoda, T.; Shirasawa, T.; Muramatsu, M. Identification of mouse ULK1, a novel protein kinase structurally related to C. elegans UNC-51. Biochem. Biophys. Res. Commun 1998, 246, 222–227. [Google Scholar]
  121. Yan, J.; Kuroyanagi, H.; Tomemori, T.; Okazaki, N.; Asato, K.; Matsuda, Y.; Suzuki, Y.; Ohshima, Y.; Mitani, S.; Masuho, Y.; et al. Mouse ULK2, a novel member of the UNC-51-like protein kinases: unique features of functional domains. Oncogene 1999, 18, 5850–5859. [Google Scholar]
  122. Hosokawa, N.; Sasaki, T.; Iemura, S.; Natsume, T.; Hara, T.; Mizushima, N. Atg101, a novel mammalian autophagy protein interacting with Atg13. Autophagy 2009, 5, 973–979. [Google Scholar]
  123. Jung, C.H.; Jun, C.B.; Ro, S.H.; Kim, Y.M.; Otto, N.M.; Cao, J.; Kundu, M.; Kim, D.H. ULK-Atg13-FIP200 complexes mediate mTOR signaling to the autophagy machinery. Mol. Biol. Cell 2009, 20, 1992–2003. [Google Scholar]
  124. Ding, Z.; Liang, J.; Li, J.; Lu, Y.; Ariyaratna, V.; Lu, Z.; Davies, M.A.; Westwick, J.K.; Mills, G.B. Physical association of PDK1 with AKT1 is sufficient for pathway activation independent of membrane localization and phosphatidylinositol 3 kinase. PLoS One 2010, 5, e9910. [Google Scholar]
  125. Glidden, E.J.; Gray, L.G.; Vemuru, S.; Li, D.; Harris, T.E.; Mayo, M.W. Multiple site acetylation of rictor stimulates mammalian target of rapamycin complex 2 (mTORC2)-dependent phosphorylation of Akt protein. J. Biol. Chem 2012, 287, 581–588. [Google Scholar]
  126. Chen, J.X.; Tuo, Q.; Liao, D.F.; Zeng, H. Inhibition of protein tyrosine phosphatase improves angiogenesis via enhancing Ang-1/Tie-2 signaling in diabetes. Exp. Diabetes Res 2012, 2012, 836759. [Google Scholar]
  127. Deblon, N.; Bourgoin, L.; Veyrat-Durebex, C.; Peyrou, M.; Vinciguerra, M.; Caillon, A.; Maeder, C.; Fournier, M.; Montet, X.; Rohner-Jeanrenaud, F.; et al. Chronic mTOR inhibition by rapamycin induces muscle insulin resistance despite weight loss in rats. Br. J. Pharmacol 2012, 165, 2325–2340. [Google Scholar]
  128. Hou, J.; Chong, Z.Z.; Shang, Y.C.; Maiese, K. Early apoptotic vascular signaling is determined by Sirt1 through nuclear shuttling, forkhead trafficking, bad, and mitochondrial caspase activation. Curr. Neurovasc. Res 2010, 7, 95–112. [Google Scholar]
  129. Maiese, K.; Chong, Z.Z.; Shang, Y.C. Mechanistic insights into diabetes mellitus and oxidative stress. Curr. Med. Chem 2007, 14, 1729–1738. [Google Scholar]
  130. Saha, A.K.; Xu, X.J.; Lawson, E.; Deoliveira, R.; Brandon, A.E.; Kraegen, E.W.; Ruderman, N.B. Downregulation of AMPK accompanies leucine- and glucose-induced increases in protein synthesis and insulin resistance in rat skeletal muscle. Diabetes 2010, 59, 2426–2434. [Google Scholar]
  131. Lu, M.J.; Chen, Y.S.; Huang, H.S.; Ma, M.C. Erythropoietin alleviates post-ischemic injury of rat hearts by attenuating nitrosative stress. Life Sci 2012, 90, 776–784. [Google Scholar]
  132. Koshimizu, T.; Kawai, M.; Kondou, H.; Tachikawa, K.; Sakai, N.; Ozono, K.; Michigami, T. Vinculin functions as regulator of chondrogenesis. J. Biol. Chem 2012, 287, 15760–15775. [Google Scholar]
  133. Chong, Z.Z.; Shang, Y.C.; Maiese, K. Cardiovascular disease and mTOR signaling. Trends Cardiovasc. Med 2011, 21, 151–155. [Google Scholar]
  134. James, M.F.; Stivison, E.; Beauchamp, R.; Han, S.; Li, H.; Wallace, M.R.; Gusella, J.F.; Stemmer-Rachamimov, A.O.; Ramesh, V. Regulation of mTOR complex 2 signaling in neurofibromatosis 2-deficient target cell types. Mol. Cancer Res 2012, 10, 649–659. [Google Scholar]
  135. Weber, J.D.; Gutmann, D.H. Deconvoluting mTOR biology. Cell Cycle 2012, 11, 236–248. [Google Scholar]
  136. Chong, Z.Z.; Wang, S.; Shang, Y.C.; Maiese, K. Targeting cardiovascular disease with novel SIRT1 pathways. Future Cardiol 2012, 8, 89–100. [Google Scholar]
  137. Fingar, D.C.; Richardson, C.J.; Tee, A.R.; Cheatham, L.; Tsou, C.; Blenis, J. mTOR controls cell cycle progression through its cell growth effectors S6K1 and 4E-BP1/eukaryotic translation initiation factor 4E. Mol. Cell. Biol 2004, 24, 200–216. [Google Scholar]
  138. Jastrzebski, K.; Hannan, K.M.; Tchoubrieva, E.B.; Hannan, R.D.; Pearson, R.B. Coordinate regulation of ribosome biogenesis and function by the ribosomal protein S6 kinase, a key mediator of mTOR function. Growth Factors 2007, 25, 209–226. [Google Scholar]
  139. Gingras, A.C.; Kennedy, S.G.; O’Leary, M.A.; Sonenberg, N.; Hay, N. 4E-BP1, a repressor of mRNA translation, is phosphorylated and inactivated by the Akt(PKB) signaling pathway. Genes Dev 1998, 12, 502–513. [Google Scholar]
  140. Bhandari, B.K.; Feliers, D.; Duraisamy, S.; Stewart, J.L.; Gingras, A.C.; Abboud, H.E.; Choudhury, G.G.; Sonenberg, N.; Kasinath, B.S. Insulin regulation of protein translation repressor 4E-BP1, an eIF4E-binding protein, in renal epithelial cells. Kidney Int 2001, 59, 866–875. [Google Scholar]
  141. Inoki, K.; Li, Y.; Zhu, T.; Wu, J.; Guan, K.L. TSC2 is phosphorylated and inhibited by Akt and suppresses mTOR signalling. Nat. Cell Biol 2002, 4, 648–657. [Google Scholar]
  142. Sato, T.; Nakashima, A.; Guo, L.; Tamanoi, F. Specific activation of mTORC1 by Rheb G-protein in vitro involves enhanced recruitment of its substrate protein. J. Biol. Chem 2009, 284, 12783–12791. [Google Scholar]
  143. Cai, S.L.; Tee, A.R.; Short, J.D.; Bergeron, J.M.; Kim, J.; Shen, J.; Guo, R.; Johnson, C.L.; Kiguchi, K.; Walker, C.L. Activity of TSC2 is inhibited by AKT-mediated phosphorylation and membrane partitioning. J. Cell Biol 2006, 173, 279–289. [Google Scholar]
  144. Oshiro, N.; Takahashi, R.; Yoshino, K.; Tanimura, K.; Nakashima, A.; Eguchi, S.; Miyamoto, T.; Hara, K.; Takehana, K.; Avruch, J.; et al. The proline-rich Akt substrate of 40 kDa (PRAS40) is a physiological substrate of mammalian target of rapamycin complex 1. J. Biol. Chem 2007, 282, 20329–20339. [Google Scholar]
  145. Wang, L.; Harris, T.E.; Lawrence, J.C., Jr. Regulation of proline-rich Akt substrate of 40 kDa (PRAS40) function by mammalian target of rapamycin complex 1 (mTORC1)-mediated phosphorylation. J. Biol. Chem 2008, 283, 15619–15627. [Google Scholar]
  146. Sancak, Y.; Thoreen, C.C.; Peterson, T.R.; Lindquist, R.A.; Kang, S.A.; Spooner, E.; Carr, S.A.; Sabatini, D.M. PRAS40 is an insulin-regulated inhibitor of the mTORC1 protein kinase. Mol. Cell 2007, 25, 903–915. [Google Scholar]
  147. Chong, Z.Z.; Shang, Y.C.; Wang, S.; Maiese, K. PRAS40 is an integral regulatory component of erythropoietin mTOR signaling and cytoprotection. PLoS One 2012, 7, e45456. [Google Scholar]
  148. Kovacina, K.S.; Park, G.Y.; Bae, S.S.; Guzzetta, A.W.; Schaefer, E.; Birnbaum, M.J.; Roth, R.A. Identification of a proline-rich Akt substrate as a 14-3-3 binding partner. J. Biol. Chem 2003, 278, 10189–10194. [Google Scholar]
  149. Vander Haar, E.; Lee, S.I.; Bandhakavi, S.; Griffin, T.J.; Kim, D.H. Insulin signalling to mTOR mediated by the Akt/PKB substrate PRAS40. Nat. Cell Biol 2007, 9, 316–323. [Google Scholar]
  150. Wang, L.; Harris, T.E.; Roth, R.A.; Lawrence, J.C., Jr. PRAS40 regulates mTORC1 kinase activity by functioning as a direct inhibitor of substrate binding. J. Biol. Chem 2007, 282, 20036–20044. [Google Scholar]
  151. Dan, H.C.; Adli, M.; Baldwin, A.S. Regulation of mammalian target of rapamycin activity in PTEN-inactive prostate cancer cells by I kappa B kinase alpha. Cancer Res 2007, 67, 6263–6269. [Google Scholar]
  152. Zandi, E.; Rothwarf, D.M.; Delhase, M.; Hayakawa, M.; Karin, M. The IkappaB kinase complex (IKK) contains two kinase subunits, IKKα and IKKβ, necessary for IκB phosphorylation and NF-κB activation. Cell 1997, 91, 243–252. [Google Scholar]
  153. Lee, D.F.; Kuo, H.P.; Chen, C.T.; Hsu, J.M.; Chou, C.K.; Wei, Y.; Sun, H.L.; Li, L.Y.; Ping, B.; Huang, W.C.; et al. IKK beta suppression of TSC1 links inflammation and tumor angiogenesis via the mTOR pathway. Cell 2007, 130, 440–455. [Google Scholar]
  154. Yen, C.J.; Izzo, J.G.; Lee, D.F.; Guha, S.; Wei, Y.; Wu, T.T.; Chen, C.T.; Kuo, H.P.; Hsu, J.M.; Sun, H.L.; et al. Bile acid exposure up-regulates tuberous sclerosis complex 1/mammalian target of rapamycin pathway in Barrett’s-associated esophageal adenocarcinoma. Cancer Res 2008, 68, 2632–2640. [Google Scholar]
  155. Chong, Z.Z.; Hou, J.; Shang, Y.C.; Wang, S.; Maiese, K. EPO relies upon novel signaling of Wnt1 that requires Akt1, FoxO3a, GSK-3beta, and beta-Catenin to foster vascular integrity during experimental diabetes. Curr. Neurovasc. Res 2011, 8, 103–120. [Google Scholar]
  156. Chong, Z.Z.; Kang, J.Q.; Maiese, K. Erythropoietin is a novel vascular protectant through activation of Akt1 and mitochondrial modulation of cysteine proteases. Circulation 2002, 106, 2973–2979. [Google Scholar]
  157. Chong, Z.Z.; Kang, J.Q.; Maiese, K. AKT1 drives endothelial cell membrane asymmetry and microglial activation through Bcl-xL and caspase 1, 3, and 9. Exp. Cell Res 2004, 296, 196–207. [Google Scholar]
  158. Hou, J.; Chong, Z.Z.; Shang, Y.C.; Maiese, K. FoxO3a governs early and late apoptotic endothelial programs during elevated glucose through mitochondrial and caspase signaling. Mol. Cell. Endocrinol 2010, 321, 194–206. [Google Scholar]
  159. Hou, J.; Wang, S.; Shang, Y.C.; Chong, Z.Z.; Maiese, K. Erythropoietin employs cell longevity pathways of SIRT1 to foster endothelial vascular integrity during oxidant stress. Curr. Neurovasc. Res 2011, 8, 220–235. [Google Scholar]
  160. Mannell, H.K.; Pircher, J.; Chaudhry, D.I.; Alig, S.K.; Koch, E.G.; Mettler, R.; Pohl, U.; Krotz, F. ARNO regulates VEGF-dependent tissue responses by stabilizing endothelial VEGFR-2 surface expression. Cardiovasc. Res 2012, 93, 111–119. [Google Scholar]
  161. Su, K.H.; Shyue, S.K.; Kou, Y.R.; Ching, L.C.; Chiang, A.N.; Yu, Y.B.; Chen, C.Y.; Pan, C.C.; Lee, T.S. β Common receptor integrates the erythropoietin signaling in activation of endothelial nitric oxide synthase. J. Cell. Physiol 2011, 226, 3330–3339. [Google Scholar]
  162. Chen, T.; Zhang, L.; Qu, Y.; Huo, K.; Jiang, X.; Fei, Z. The selective mGluR5 agonist CHPG protects against traumatic brain injury in vitro and in vivo via ERK and Akt pathway. Int. J. Mol. Med 2012, 29, 630–636. [Google Scholar]
  163. Chong, Z.Z.; Lin, S.H.; Maiese, K. The NAD+ precursor nicotinamide governs neuronal survival during oxidative stress through protein kinase B coupled to FOXO3a and mitochondrial membrane potential. J. Cereb. Blood Flow Metab 2004, 24, 728–743. [Google Scholar]
  164. Chong, Z.Z.; Shang, Y.C.; Hou, J.; Maiese, K. Wnt1 neuroprotection translates into improved neurological function during oxidant stress and cerebral ischemia through AKT1 and mitochondrial apoptotic pathways. Oxid. Med. Cell. Longevity 2010, 3, 153–165. [Google Scholar]
  165. Komandirov, M.A.; Knyazeva, E.A.; Fedorenko, Y.P.; Rudkovskii, M.V.; Stetsurin, D.A.; Uzdensky, A.B. On the role of phosphatidylinositol 3-kinase, protein kinase b/Akt, and glycogen synthase kinase-3β in photodynamic injury of crayfish neurons and glial cells. J. Mol. Neurosci 2011, 45, 229–235. [Google Scholar]
  166. Malagelada, C.; Jin, Z.H.; Jackson-Lewis, V.; Przedborski, S.; Greene, L.A. Rapamycin protects against neuron death in in vitro and in vivo models of Parkinson’s disease. J. Neurosci 2010, 30, 1166–1175. [Google Scholar]
  167. Shen, J.; Wu, Y.; Xu, J.Y.; Zhang, J.; Sinclair, S.H.; Yanoff, M.; Xu, G.; Li, W.; Xu, G.T. ERKand Akt-dependent neuroprotection by erythropoietin (EPO) against glyoxal-AGEs via modulation of Bcl-xL, Bax, and BAD. Invest. Ophthalmol. Vis. Sci 2010, 51, 35–46. [Google Scholar]
  168. Zeng, K.W.; Wang, X.M.; Ko, H.; Kwon, H.C.; Cha, J.W.; Yang, H.O. Hyperoside protects primary rat cortical neurons from neurotoxicity induced by amyloid beta-protein via the PI3K/Akt/Bad/Bcl(XL)-regulated mitochondrial apoptotic pathway. Eur. J. Pharmacol 2011, 672, 45–55. [Google Scholar]
  169. Li, F.; Chong, Z.Z.; Maiese, K. Microglial integrity is maintained by erythropoietin through integration of Akt and its substrates of glycogen synthase kinase-3beta, beta-catenin, and nuclear factor-kappaB. Curr. Neurovasc. Res 2006, 3, 187–201. [Google Scholar]
  170. Pineda, D.; Ampurdanes, C.; Medina, M.G.; Serratosa, J.; Tusell, J.M.; Saura, J.; Planas, A.M.; Navarro, P. Tissue plasminogen activator induces microglial inflammation via a noncatalytic molecular mechanism involving activation of mitogen-activated protein kinases and Akt signaling pathways and AnnexinA2 and Galectin-1 receptors. Glia 2012, 60, 526–540. [Google Scholar]
  171. Perez-Garcia, M.J.; Cena, V.; de Pablo, Y.; Llovera, M.; Comella, J.X.; Soler, R.M. Glial cell line-derived neurotrophic factor increases intracellular calcium concentration. Role of calcium/calmodulin in the activation of the phosphatidylinositol 3–kinase pathway. J. Biol. Chem 2004, 279, 6132–6142. [Google Scholar]
  172. Xu, J.; Zhang, Q.G.; Li, C.; Zhang, G.Y. Subtoxic N-methyl-d-aspartate delayed neuronal death in ischemic brain injury through TrkB receptor- and calmodulin-mediated PI-3K/Akt pathway activation. Hippocampus 2007, 17, 525–537. [Google Scholar]
  173. Maiese, K. The challenges for drug development: Cytokines, genes, and stem cells. Curr. Neurovasc. Res 2012, 9, 231–232. [Google Scholar]
  174. Chattopadhyay, M.; Walter, C.; Mata, M.; Fink, D.J. Neuroprotective effect of herpes simplex virus-mediated gene transfer of erythropoietin in hyperglycemic dorsal root ganglion neurons. Brain 2009, 132, 879–888. [Google Scholar]
  175. Chong, Z.Z.; Maiese, K. Erythropoietin involves the phosphatidylinositol 3-kinase pathway, 14-3-3 protein and FOXO3a nuclear trafficking to preserve endothelial cell integrity. Br. J. Pharmacol 2007, 150, 839–850. [Google Scholar]
  176. Maiese, K.; Chong, Z.Z.; Shang, Y.C.; Hou, J. A “FOXO” in sight: Targeting FoxO proteins from conception to cancer. Med. Res. Rev 2009, 29, 395–418. [Google Scholar]
  177. Storz, P. Forkhead homeobox type O transcription factors in the responses to oxidative stress. Antioxid. Redox Signaling 2011, 14, 593–605. [Google Scholar]
  178. Kousteni, S. FoxO1, the transcriptional chief of staff of energy metabolism. Bone 2012, 50, 437–443. [Google Scholar]
  179. Lam, E.W.; Shah, K.; Brosens, J.J. The diversity of sex steroid action: The role of micro-RNAs and FOXO transcription factors in cycling endometrium and cancer. J. Endocrinol 2012, 212, 13–25. [Google Scholar]
  180. Lappas, M.; Permezel, M. The anti-inflammatory and antioxidative effects of nicotinamide, a vitamin B(3) derivative, are elicited by FoxO3 in human gestational tissues: Implications for preterm birth. J. Nutr. Biochem 2011, 22, 1195–1201. [Google Scholar]
  181. Maiese, K.; Chong, Z.Z.; Shang, Y.C. OutFOXOing disease and disability: The therapeutic potential of targeting FoxO proteins. Trends Mol. Med 2008, 14, 219–227. [Google Scholar]
  182. Xie, Z.; Chen, F.; Wu, X.; Zhuang, C.; Zhu, J.; Wang, J.; Ji, H.; Wang, Y.; Hua, X. Effects of supplemental erythropoietin on its receptor expression and signal transduction pathways in rat model of retinal detachment. Curr. Eye Res 2012, 37, 138–144. [Google Scholar]
  183. Ma, R.; Xiong, N.; Huang, C.; Tang, Q.; Hu, B.; Xiang, J.; Li, G. Erythropoietin protects PC12 cells from beta-amyloid(25–35)-induced apoptosis via PI3K/Akt signaling pathway. Neuropharmacology 2009, 56, 1027–1034. [Google Scholar]
  184. Shang, Y.C.; Chong, Z.Z.; Wang, S.; Maiese, K. Prevention of beta-amyloid degeneration of microglia by erythropoietin depends on Wnt1, the PI 3-K/mTOR pathway, Bad, and Bcl-xL. Aging (Albany NY) 2012, 4, 187–201. [Google Scholar]
  185. Sun, Z.K.; Yang, H.Q.; Pan, J.; Zhen, H.; Wang, Z.Q.; Chen, S.D.; Ding, J.Q. Protective effects of erythropoietin on tau phosphorylation induced by beta-amyloid. J. Neurosci. Res 2008, 86, 3018–3027. [Google Scholar]
  186. Kilic, E.; Kilic, U.; Soliz, J.; Bassetti, C.L.; Gassmann, M.; Hermann, D.M. Brain-derived erythropoietin protects from focal cerebral ischemia by dual activation of ERK-1/-2 and Akt pathways. FASEB J 2005, 19, 2026–2028. [Google Scholar]
  187. Chong, Z.Z.; Kang, J.Q.; Maiese, K. Erythropoietin fosters both intrinsic and extrinsic neuronal protection through modulation of microglia, Akt1, Bad, and caspase-mediated pathways. Br. J. Pharmacol 2003, 138, 1107–1118. [Google Scholar]
  188. Dzietko, M.; Felderhoff-Mueser, U.; Sifringer, M.; Krutz, B.; Bittigau, P.; Thor, F.; Heumann, R.; Buhrer, C.; Ikonomidou, C.; Hansen, H.H. Erythropoietin protects the developing brain against N-methyl-d-aspartate receptor antagonist neurotoxicity. Neurobiol. Dis 2004, 15, 177–187. [Google Scholar]
  189. Um, M.; Lodish, H.F. Antiapoptotic effects of erythropoietin in differentiated neuroblastoma SH-SY5Y cells require activation of both the STAT5 and AKT signaling pathways. J. Biol. Chem 2006, 281, 5648–5656. [Google Scholar]
  190. Chong, Z.Z.; Lin, S.H.; Li, F.; Maiese, K. The sirtuin inhibitor nicotinamide enhances neuronal cell survival during acute anoxic injury through Akt, Bad, PARP, and mitochondrial associated “anti-apoptotic” pathways. Curr. Neurovasc. Res 2005, 2, 271–285. [Google Scholar]
  191. Li, F.; Chong, Z.Z.; Maiese, K. Cell life versus cell longevity: The mysteries surrounding the NAD(+) precursor nicotinamide. Curr. Med. Chem 2006, 13, 883–895. [Google Scholar]
  192. Maiese, K.; Li, F.; Chong, Z.Z. Erythropoietin in the brain: Can the promise to protect be fulfilled? Trends Pharmacol. Sci 2004, 25, 577–583. [Google Scholar]
  193. Wang, S.; Chong, Z.Z.; Shang, Y.C.; Maiese, K. Wnt1 inducible signaling pathway protein 1 (WISP1) blocks neurodegeneration through phosphoinositide 3 kinase/Akt1 and apoptotic mitochondrial signaling involving Bad, Bax, Bim, and Bcl-xL. Curr. Neurovasc. Res 2012, 9, 20–31. [Google Scholar]
  194. Koh, P.O. Nicotinamide attenuates the ischemic brain injury-induced decrease of Akt activation and Bad phosphorylation. Neurosci. Lett 2011, 498, 105–109. [Google Scholar]
  195. Pastor, M.D.; Garcia-Yebenes, I.; Fradejas, N.; Perez-Ortiz, J.M.; Mora-Lee, S.; Tranque, P.; Moro, M.A.; Pende, M.; Calvo, S. mTOR/S6 kinase pathway contributes to astrocyte survival during ischemia. J. Biol. Chem 2009, 284, 22067–22078. [Google Scholar]
  196. Wu, X.; Reiter, C.E.; Antonetti, D.A.; Kimball, S.R.; Jefferson, L.S.; Gardner, T.W. Insulin promotes rat retinal neuronal cell survival in a p70S6K-dependent manner. J. Biol. Chem 2004, 279, 9167–9175. [Google Scholar]
  197. Maiese, K.; Li, F.; Chong, Z.Z. New avenues of exploration for erythropoietin. JAMA 2005, 293, 90–95. [Google Scholar]
  198. Galan-Moya, E.M.; Le Guelte, A.; Lima Fernandes, E.; Thirant, C.; Dwyer, J.; Bidere, N.; Couraud, P.O.; Scott, M.G.; Junier, M.P.; Chneiweiss, H.; et al. Secreted factors from brain endothelial cells maintain glioblastoma stem-like cell expansion through the mTOR pathway. EMBO Rep 2011, 12, 470–476. [Google Scholar]
  199. Marfia, G.; Madaschi, L.; Marra, F.; Menarini, M.; Bottai, D.; Formenti, A.; Bellardita, C.; Di Giulio, A.M.; Carelli, S.; Gorio, A. Adult neural precursors isolated from post mortem brain yield mostly neurons: an erythropoietin-dependent process. Neurobiol. Dis 2011, 43, 86–98. [Google Scholar]
  200. Sanghera, K.P.; Mathalone, N.; Baigi, R.; Panov, E.; Wang, D.; Zhao, X.; Hsu, H.; Wang, H.; Tropepe, V.; Ward, M.; et al. The PI3K/Akt/mTOR pathway mediates retinal progenitor cell survival under hypoxic and superoxide stress. Mol. Cell. Neurosci 2011, 47, 145–153. [Google Scholar]
  201. Hou, G.; Xue, L.; Lu, Z.; Fan, T.; Tian, F.; Xue, Y. An activated mTOR/p70S6K signaling pathway in esophageal squamous cell carcinoma cell lines and inhibition of the pathway by rapamycin and siRNA against mTOR. Cancer Lett 2007, 253, 236–248. [Google Scholar]
  202. Zhang, D.; Contu, R.; Latronico, M.V.; Zhang, J.; Rizzi, R.; Catalucci, D.; Miyamoto, S.; Huang, K.; Ceci, M.; Gu, Y.; et al. MTORC1 regulates cardiac function and myocyte survival through 4E-BP1 inhibition in mice. J. Clin. Invest 2010, 120, 2805–2816. [Google Scholar]
  203. Dormond, O.; Madsen, J.C.; Briscoe, D.M. The effects of mTOR-Akt interactions on anti-apoptotic signaling in vascular endothelial cells. J. Biol. Chem 2007, 282, 23679–23686. [Google Scholar]
  204. Chong, Z.Z.; Li, F.; Maiese, K. The pro-survival pathways of mTOR and protein kinase B target glycogen synthase kinase-3beta and nuclear factor-κB to foster endogenous microglial cell protection. Int. J. Mol. Med 2007, 19, 263–272. [Google Scholar]
  205. Choi, K.C.; Kim, S.H.; Ha, J.Y.; Kim, S.T.; Son, J.H. A novel mTOR activating protein protects dopamine neurons against oxidative stress by repressing autophagy related cell death. J. Neurochem 2010, 112, 366–376. [Google Scholar]
  206. Shang, Y.C.; Chong, Z.Z.; Wang, S.; Maiese, K. WNT1 inducible signaling pathway protein 1 (WISP1) targets PRAS40 to govern β-amyloid apoptotic injury of microglia. Curr. Neurovasc. Res 2012, 9, 239–249. [Google Scholar]
  207. Thedieck, K.; Polak, P.; Kim, M.L.; Molle, K.D.; Cohen, A.; Jeno, P.; Arrieumerlou, C.; Hall, M.N. PRAS40 and PRR5-like protein are new mTOR interactors that regulate apoptosis. PLoS One 2007, 2, e1217. [Google Scholar]
  208. Chong, Z.Z.; Li, F.; Maiese, K. Attempted cell cycle induction in post-mitotic neurons occurs in early and late apoptotic programs through Rb, E2F1, and Caspase 3. Curr. Neurovasc. Res 2006, 3, 25–39. [Google Scholar]
  209. Yu, Y.; Ren, Q.G.; Zhang, Z.H.; Zhou, K.; Yu, Z.Y.; Luo, X.; Wang, W. Phospho-Rb mediating cell cycle reentry induces early apoptosis following oxygen-glucose deprivation in rat cortical neurons. Neurochem. Res 2012, 37, 503–511. [Google Scholar]
  210. Bhaskar, K.; Miller, M.; Chludzinski, A.; Herrup, K.; Zagorski, M.; Lamb, B.T. The PI3K-Akt-mTOR pathway regulates Abeta oligomer induced neuronal cell cycle events. Mol. Neurodegener 2009, 4, 14. [Google Scholar]
  211. Yu, L.; McPhee, C.K.; Zheng, L.; Mardones, G.A.; Rong, Y.; Peng, J.; Mi, N.; Zhao, Y.; Liu, Z.; Wan, F.; et al. Termination of autophagy and reformation of lysosomes regulated by mTOR. Nature 2010, 465, 942–946. [Google Scholar]
  212. Rong, Y.; McPhee, C.K.; Deng, S.; Huang, L.; Chen, L.; Liu, M.; Tracy, K.; Baehrecke, E.H.; Yu, L.; Lenardo, M.J. Spinster is required for autophagic lysosome reformation and mTOR reactivation following starvation. Proc. Natl. Acad. Sci. USA 2011, 108, 7826–7831. [Google Scholar]
  213. Abdullah, A.; Wolfe, R.; Mannan, H.; Stoelwinder, J.U.; Stevenson, C.; Peeters, A. Epidemiologic merit of obese-years, the combination of degree and duration of obesity. Am. J. Epidemiol 2012, 176, 99–107. [Google Scholar]
  214. Reagan, L.P. Diabetes as a chronic metabolic stressor: Causes, consequences and clinical complications. Exp. Neurol 2012, 233, 68–78. [Google Scholar]
  215. Kelly, G.S. A review of the sirtuin system, its clinical implications, and the potential role of dietary activators like resveratrol: Part 2. Altern. Med. Rev 2010, 15, 313–328. [Google Scholar]
  216. Lai, C.S.; Tsai, M.L.; Badmaev, V.; Jimenez, M.; Ho, C.T.; Pan, M.H. Xanthigen suppresses preadipocyte differentiation and adipogenesis through down-regulation of PPARgamma and C/EBPs and modulation of SIRT-1, AMPK, and FoxO pathways. J. Agric. Food Chem 2012, 60, 1094–1101. [Google Scholar]
  217. Chong, Z.Z.; Maiese, K. Enhanced tolerance against early and late apoptotic oxidative stress in mammalian neurons through nicotinamidase and sirtuin mediated pathways. Curr. Neurovasc. Res 2008, 5, 159–170. [Google Scholar]
  218. Sun, C.; Zhang, F.; Ge, X.; Yan, T.; Chen, X.; Shi, X.; Zhai, Q. SIRT1 improves insulin sensitivity under insulin-resistant conditions by repressing PTP1B. Cell Metab 2007, 6, 307–319. [Google Scholar]
  219. Li, Y.; Xu, S.; Giles, A.; Nakamura, K.; Lee, J.W.; Hou, X.; Donmez, G.; Li, J.; Luo, Z.; Walsh, K.; et al. Hepatic overexpression of SIRT1 in mice attenuates endoplasmic reticulum stress and insulin resistance in the liver. FASEB J 2011, 25, 1664–1679. [Google Scholar]
  220. Frojdo, S.; Durand, C.; Molin, L.; Carey, A.L.; El-Osta, A.; Kingwell, B.A.; Febbraio, M.A.; Solari, F.; Vidal, H.; Pirola, L. Phosphoinositide 3-kinase as a novel functional target for the regulation of the insulin signaling pathway by SIRT1. Mol. Cell. Endocrinol 2011, 335, 166–176. [Google Scholar]
  221. Bordone, L.; Motta, M.C.; Picard, F.; Robinson, A.; Jhala, U.S.; Apfeld, J.; McDonagh, T.; Lemieux, M.; McBurney, M.; Szilvasi, A.; et al. Sirt1 regulates insulin secretion by repressing UCP2 in pancreatic beta cells. PLoS Biol 2006, 4, e31. [Google Scholar]
  222. Sundaresan, N.R.; Pillai, V.B.; Wolfgeher, D.; Samant, S.; Vasudevan, P.; Parekh, V.; Raghuraman, H.; Cunningham, J.M.; Gupta, M.; Gupta, M.P. The deacetylase SIRT1 promotes membrane localization and activation of Akt and PDK1 during tumorigenesis and cardiac hypertrophy. Sci. Signal. 2011, 4. [Google Scholar] [CrossRef]
  223. Wang, F.; Chan, C.H.; Chen, K.; Guan, X.; Lin, H.K.; Tong, Q. Deacetylation of FOXO3 by SIRT1 or SIRT2 leads to Skp2-mediated FOXO3 ubiquitination and degradation. Oncogene 2012, 31, 1546–1557. [Google Scholar]
  224. Wang, R.H.; Kim, H.S.; Xiao, C.; Xu, X.; Gavrilova, O.; Deng, C.X. Hepatic Sirt1 deficiency in mice impairs mTorc2/Akt signaling and results in hyperglycemia, oxidative damage, and insulin resistance. J. Clin. Invest 2011, 121, 4477–4490. [Google Scholar]
  225. Guo, W.; Qian, L.; Zhang, J.; Zhang, W.; Morrison, A.; Hayes, P.; Wilson, S.; Chen, T.; Zhao, J. Sirt1 overexpression in neurons promotes neurite outgrowth and cell survival through inhibition of the mTOR signaling. J. Neurosci. Res 2011, 89, 1723–1736. [Google Scholar]
  226. Hamada, S.; Hara, K.; Hamada, T.; Yasuda, H.; Moriyama, H.; Nakayama, R.; Nagata, M.; Yokono, K. Upregulation of the mammalian target of rapamycin complex 1 pathway by Ras homolog enriched in brain in pancreatic beta-cells leads to increased beta-cell mass and prevention of hyperglycemia. Diabetes 2009, 58, 1321–1332. [Google Scholar]
  227. Treins, C.; Alliouachene, S.; Hassouna, R.; Xie, Y.; Birnbaum, M.J.; Pende, M. The combined deletion of S6K1 and Akt2 deteriorates glycaemic control in high fat diet. Mol. Cell. Biol 2012, 32, 4004–4011. [Google Scholar]
  228. Fraenkel, M.; Ketzinel-Gilad, M.; Ariav, Y.; Pappo, O.; Karaca, M.; Castel, J.; Berthault, M.F.; Magnan, C.; Cerasi, E.; Kaiser, N.; et al. mTOR inhibition by rapamycin prevents beta-cell adaptation to hyperglycemia and exacerbates the metabolic state in type 2 diabetes. Diabetes 2008, 57, 945–957. [Google Scholar]
  229. Cho, H.J.; Park, J.; Lee, H.W.; Lee, Y.S.; Kim, J.B. Regulation of adipocyte differentiation and insulin action with rapamycin. Biochem. Biophys. Res. Commun 2004, 321, 942–948. [Google Scholar]
  230. Yeh, W.C.; Bierer, B.E.; McKnight, S.L. Rapamycin inhibits clonal expansion and adipogenic differentiation of 3T3-L1 cells. Proc. Natl. Acad. Sci. USA 1995, 92, 11086–11090. [Google Scholar]
  231. Chakrabarti, P.; English, T.; Shi, J.; Smas, C.M.; Kandror, K.V. Mammalian target of rapamycin complex 1 suppresses lipolysis, stimulates lipogenesis, and promotes fat storage. Diabetes 2010, 59, 775–781. [Google Scholar]
  232. Rovira, J.; Marcelo Arellano, E.; Burke, J.T.; Brault, Y.; Moya-Rull, D.; Banon-Maneus, E.; Ramirez-Bajo, M.J.; Gutierrez-Dalmau, A.; Revuelta, I.; Quintana, L.F.; et al. Effect of mTOR inhibitor on body weight: From an experimental rat model to human transplant patients. Transpl. Int 2008, 21, 992–998. [Google Scholar]
  233. Phan, J.; Peterfy, M.; Reue, K. Lipin expression preceding peroxisome proliferator-activated receptor-γ is critical for adipogenesis in vivo and in vitro. J. Biol. Chem 2004, 279, 29558–29564. [Google Scholar]
  234. Peterfy, M.; Phan, J.; Xu, P.; Reue, K. Lipodystrophy in the fld mouse results from mutation of a new gene encoding a nuclear protein, lipin. Nat. Genet 2001, 27, 121–124. [Google Scholar]
  235. Finck, B.N.; Gropler, M.C.; Chen, Z.; Leone, T.C.; Croce, M.A.; Harris, T.E.; Lawrence, J.C., Jr; Kelly, D.P. Lipin 1 is an inducible amplifier of the hepatic PGC-1alpha/PPARalpha regulatory pathway. Cell Metab 2006, 4, 199–210. [Google Scholar]
  236. Huffman, T.A.; Mothe-Satney, I.; Lawrence, J.C., Jr. Insulin-stimulated phosphorylation of lipin mediated by the mammalian target of rapamycin. Proc. Natl. Acad. Sci. USA 2002, 99, 1047–1052. [Google Scholar]
  237. Cota, D.; Proulx, K.; Smith, K.A.; Kozma, S.C.; Thomas, G.; Woods, S.C.; Seeley, R.J. Hypothalamic mTOR signaling regulates food intake. Science 2006, 312, 927–930. [Google Scholar]
  238. Cota, D.; Matter, E.K.; Woods, S.C.; Seeley, R.J. The role of hypothalamic mammalian target of rapamycin complex 1 signaling in diet-induced obesity. J. Neurosci 2008, 28, 7202–7208. [Google Scholar]
  239. Chakrabarti, P.; Anno, T.; Manning, B.D.; Luo, Z.; Kandror, K.V. The mammalian target of rapamycin complex 1 regulates leptin biosynthesis in adipocytes at the level of translation: The role of the 5′-untranslated region in the expression of leptin messenger ribonucleic acid. Mol. Endocrinol 2008, 22, 2260–2267. [Google Scholar]
  240. Inoki, K.; Zhu, T.; Guan, K.L. TSC2 mediates cellular energy response to control cell growth and survival. Cell 2003, 115, 577–590. [Google Scholar]
  241. Chong, Z.Z.; Kang, J.Q.; Maiese, K. Apaf-1, Bcl-xL, Cytochrome c, and Caspase-9 form the critical elements for cerebral vascular protection by erythropoietin. J. Cereb. Blood Flow Metab 2003, 23, 320–330. [Google Scholar]
  242. Toba, H.; Kojima, Y.; Wang, J.; Noda, K.; Tian, W.; Kobara, M.; Nakata, T. Erythropoietin attenuated vascular dysfunction and inflammation by inhibiting NADPH oxidase-derived superoxide production in nitric oxide synthase-inhibited hypertensive rat aorta. Eur. J. Pharmacol 2012, 691, 190–197. [Google Scholar]
  243. Sheng, B.; Liu, J.; Li, G.H. Metformin preconditioning protects Daphnia pulex from lethal hypoxic insult involving AMPK, HIF and mTOR signaling. Comp. Biochem. Physiol. B Biochem. Mol. Biol 2012, 163, 51–58. [Google Scholar]
  244. Koh, P.O. Melatonin prevents ischemic brain injury through activation of the mTOR/p70S6 kinase signaling pathway. Neurosci. Lett 2008, 444, 74–78. [Google Scholar]
  245. Liu, G.; Detloff, M.R.; Miller, K.N.; Santi, L.; Houle, J.D. Exercise modulates microRNAs that affect the PTEN/mTOR pathway in rats after spinal cord injury. Exp. Neurol 2012, 233, 447–456. [Google Scholar]
  246. Walker, C.L.; Walker, M.J.; Liu, N.K.; Risberg, E.C.; Gao, X.; Chen, J.; Xu, X.M. Systemic bisperoxovanadium activates Akt/mTOR, reduces autophagy, and enhances recovery following cervical spinal cord injury. PLoS One 2012, 7, e30012. [Google Scholar]
  247. Hu, L.Y.; Sun, Z.G.; Wen, Y.M.; Cheng, G.Z.; Wang, S.L.; Zhao, H.B.; Zhang, X.R. ATP-mediated protein kinase B Akt/mammalian target of rapamycin mTOR/p70 ribosomal S6 protein p70S6 kinase signaling pathway activation promotes improvement of locomotor function after spinal cord injury in rats. Neuroscience 2010, 169, 1046–1062. [Google Scholar]
  248. Sun, F.; Park, K.K.; Belin, S.; Wang, D.; Lu, T.; Chen, G.; Zhang, K.; Yeung, C.; Feng, G.; Yankner, B.A.; et al. Sustained axon regeneration induced by co-deletion of PTEN and SOCS3. Nature 2011, 480, 372–375. [Google Scholar] [Green Version]
  249. Liu, K.; Lu, Y.; Lee, J.K.; Samara, R.; Willenberg, R.; Sears-Kraxberger, I.; Tedeschi, A.; Park, K.K.; Jin, D.; Cai, B.; et al. PTEN deletion enhances the regenerative ability of adult corticospinal neurons. Nat. Neurosci 2010, 13, 1075–1081. [Google Scholar]
  250. Park, K.K.; Liu, K.; Hu, Y.; Smith, P.D.; Wang, C.; Cai, B.; Xu, B.; Connolly, L.; Kramvis, I.; Sahin, M.; et al. Promoting axon regeneration in the adult CNS by modulation of the PTEN/mTOR pathway. Science 2008, 322, 963–966. [Google Scholar]
  251. Erlich, S.; Alexandrovich, A.; Shohami, E.; Pinkas-Kramarski, R. Rapamycin is a neuroprotective treatment for traumatic brain injury. Neurobiol. Dis 2007, 26, 86–93. [Google Scholar]
  252. Sekiguchi, A.; Kanno, H.; Ozawa, H.; Yamaya, S.; Itoi, E. Rapamycin promotes autophagy and reduces neural tissue damage and locomotor impairment after spinal cord injury in mice. J. Neurotrauma 2012, 29, 946–956. [Google Scholar]
  253. Shi, G.D.; OuYang, Y.P.; Shi, J.G.; Liu, Y.; Yuan, W.; Jia, L.S. PTEN deletion prevents ischemic brain injury by activating the mTOR signaling pathway. Biochem. Biophys. Res. Commun 2011, 404, 941–945. [Google Scholar]
  254. Zhang, W.; Khatibi, N.H.; Yamaguchi-Okada, M.; Yan, J.; Chen, C.; Hu, Q.; Meng, H.; Han, H.; Liu, S.; Zhou, C. Mammalian target of rapamycin (mTOR) inhibition reduces cerebral vasospasm following a subarachnoid hemorrhage injury in canines. Exp. Neurol 2012, 233, 799–806. [Google Scholar]
  255. Zeng, L.H.; Rensing, N.R.; Wong, M. The mammalian target of rapamycin signaling pathway mediates epileptogenesis in a model of temporal lobe epilepsy. J. Neurosci 2009, 29, 6964–6972. [Google Scholar]
  256. Buckmaster, P.S.; Ingram, E.A.; Wen, X. Inhibition of the mammalian target of rapamycin signaling pathway suppresses dentate granule cell axon sprouting in a rodent model of temporal lobe epilepsy. J. Neurosci 2009, 29, 8259–8269. [Google Scholar]
  257. Holmes, G.L.; Stafstrom, C.E. Tuberous sclerosis complex and epilepsy: Recent developments and future challenges. Epilepsia 2007, 48, 617–630. [Google Scholar]
  258. Waltereit, R.; Welzl, H.; Dichgans, J.; Lipp, H.P.; Schmidt, W.J.; Weller, M. Enhanced episodic-like memory and kindling epilepsy in a rat model of tuberous sclerosis. J. Neurochem 2006, 96, 407–413. [Google Scholar]
  259. Zeng, L.H.; Xu, L.; Gutmann, D.H.; Wong, M. Rapamycin prevents epilepsy in a mouse model of tuberous sclerosis complex. Ann. Neurol 2008, 63, 444–453. [Google Scholar]
  260. Lee, S.T.; Chu, K.; Park, J.E.; Jung, K.H.; Jeon, D.; Lim, J.Y.; Lee, S.K.; Kim, M.; Roh, J.K. Erythropoietin improves memory function with reducing endothelial dysfunction and amyloid-beta burden in Alzheimer’s disease models. J. Neurochem 2012, 120, 115–124. [Google Scholar]
  261. Burgos-Ramos, E.; Martos-Moreno, G.A.; Lopez, M.G.; Herranz, R.; Aguado-Llera, D.; Egea, J.; Frechilla, D.; Cenarruzabeitia, E.; Leon, R.; Arilla-Ferreiro, E.; et al. The N-terminal tripeptide of insulin-like growth factor-I protects against beta-amyloid-induced somatostatin depletion by calcium and glycogen synthase kinase 3 beta modulation. J. Neurochem 2009, 109, 360–370. [Google Scholar]
  262. Echeverria, V.; Zeitlin, R.; Burgess, S.; Patel, S.; Barman, A.; Thakur, G.; Mamcarz, M.; Wang, L.; Sattelle, D.B.; Kirschner, D.A.; et al. Cotinine reduces amyloid-beta aggregation and improves memory in Alzheimer’s disease mice. J. Alzheimers Dis 2011, 24, 817–835. [Google Scholar]
  263. Li, L.; Xu, B.; Zhu, Y.; Chen, L.; Sokabe, M. DHEA prevents Abeta25-35-impaired survival of newborn neurons in the dentate gyrus through a modulation of PI3K-Akt-mTOR signaling. Neuropharmacology 2010, 59, 323–333. [Google Scholar]
  264. Chano, T.; Okabe, H.; Hulette, C.M. RB1CC1 insufficiency causes neuronal atrophy through mTOR signaling alteration and involved in the pathology of Alzheimer’s diseases. Brain Res 2007, 1168, 97–105. [Google Scholar]
  265. Paccalin, M.; Pain-Barc, S.; Pluchon, C.; Paul, C.; Besson, M.N.; Carret-Rebillat, A.S.; Rioux-Bilan, A.; Gil, R.; Hugon, J. Activated mTOR and PKR kinases in lymphocytes correlate with memory and cognitive decline in Alzheimer’s disease. Dementia Geriatr. Cognit. Disord 2006, 22, 320–326. [Google Scholar]
  266. Slipczuk, L.; Bekinschtein, P.; Katche, C.; Cammarota, M.; Izquierdo, I.; Medina, J.H. BDNF activates mTOR to regulate GluR1 expression required for memory formation. PLoS One 2009, 4, e6007. [Google Scholar]
  267. Ma, T.; Hoeffer, C.A.; Capetillo-Zarate, E.; Yu, F.; Wong, H.; Lin, M.T.; Tampellini, D.; Klann, E.; Blitzer, R.D.; Gouras, G.K. Dysregulation of the mTOR pathway mediates impairment of synaptic plasticity in a mouse model of Alzheimer’s disease. PLoS One 2010, 5, e12845. [Google Scholar]
  268. Lafay-Chebassier, C.; Paccalin, M.; Page, G.; Barc-Pain, S.; Perault-Pochat, M.C.; Gil, R.; Pradier, L.; Hugon, J. mTOR/p70S6k signalling alteration by Aβ exposure as well as in APP-PS1 transgenic models and in patients with Alzheimer’s disease. J. Neurochem 2005, 94, 215–225. [Google Scholar]
  269. Griffin, R.J.; Moloney, A.; Kelliher, M.; Johnston, J.A.; Ravid, R.; Dockery, P.; O’Connor, R.; O’Neill, C. Activation of Akt/PKB, increased phosphorylation of Akt substrates and loss and altered distribution of Akt and PTEN are features of Alzheimer’s disease pathology. J. Neurochem 2005, 93, 105–117. [Google Scholar]
  270. Esteras, N.; Munoz, U.; Alquezar, C.; Bartolome, F.; Bermejo-Pareja, F.; Martin-Requero, A. Altered calmodulin degradation and signaling in non-neuronal cells from Alzheimer’s disease patients. Curr. Alzheimer Res 2012, 9, 267–277. [Google Scholar]
  271. An, W.L.; Cowburn, R.F.; Li, L.; Braak, H.; Alafuzoff, I.; Iqbal, K.; Iqbal, I.G.; Winblad, B.; Pei, J.J. Up-regulation of phosphorylated/activated p70 S6 kinase and its relationship to neurofibrillary pathology in Alzheimer’s disease. Am. J. Pathol 2003, 163, 591–607. [Google Scholar]
  272. Sarkar, S.; Ravikumar, B.; Rubinsztein, D.C. Autophagic clearance of aggregate-prone proteins associated with neurodegeneration. Methods Enzymol 2009, 453, 83–110. [Google Scholar]
  273. Berger, Z.; Ravikumar, B.; Menzies, F.M.; Oroz, L.G.; Underwood, B.R.; Pangalos, M.N.; Schmitt, I.; Wullner, U.; Evert, B.O.; O’Kane, C.J.; et al. Rapamycin alleviates toxicity of different aggregate-prone proteins. Hum. Mol. Genet 2006, 15, 433–442. [Google Scholar]
  274. Ravikumar, B.; Vacher, C.; Berger, Z.; Davies, J.E.; Luo, S.; Oroz, L.G.; Scaravilli, F.; Easton, D.F.; Duden, R.; O’Kane, C.J.; et al. Inhibition of mTOR induces autophagy and reduces toxicity of polyglutamine expansions in fly and mouse models of Huntington disease. Nat. Genet 2004, 36, 585–595. [Google Scholar]
  275. Floto, R.A.; Sarkar, S.; Perlstein, E.O.; Kampmann, B.; Schreiber, S.L.; Rubinsztein, D.C. Small molecule enhancers of rapamycin-induced TOR inhibition promote autophagy, reduce toxicity in Huntington’s disease models and enhance killing of mycobacteria by macrophages. Autophagy 2007, 3, 620–622. [Google Scholar]
  276. Roscic, A.; Baldo, B.; Crochemore, C.; Marcellin, D.; Paganetti, P. Induction of autophagy with catalytic mTOR inhibitors reduces huntingtin aggregates in a neuronal cell model. J. Neurochem 2011, 119, 398–407. [Google Scholar]
  277. Fox, J.H.; Connor, T.; Chopra, V.; Dorsey, K.; Kama, J.A.; Bleckmann, D.; Betschart, C.; Hoyer, D.; Frentzel, S.; Difiglia, M.; et al. The mTOR kinase inhibitor Everolimus decreases S6 kinase phosphorylation but fails to reduce mutant huntingtin levels in brain and is not neuroprotective in the R6/2 mouse model of Huntington’s disease. Mol. Neurodegener 2010, 5, 26. [Google Scholar]
  278. Crews, L.; Spencer, B.; Desplats, P.; Patrick, C.; Paulino, A.; Rockenstein, E.; Hansen, L.; Adame, A.; Galasko, D.; Masliah, E. Selective molecular alterations in the autophagy pathway in patients with Lewy body disease and in models of alpha-synucleinopathy. PLoS One 2010, 5, e9313. [Google Scholar]
  279. Tain, L.S.; Mortiboys, H.; Tao, R.N.; Ziviani, E.; Bandmann, O.; Whitworth, A.J. Rapamycin activation of 4E-BP prevents parkinsonian dopaminergic neuron loss. Nat. Neurosci 2009, 12, 1129–1135. [Google Scholar]
  280. DeYoung, M.P.; Horak, P.; Sofer, A.; Sgroi, D.; Ellisen, L.W. Hypoxia regulates TSC1/2-mTOR signaling and tumor suppression through REDD1–mediated 14-3-3 shuttling. Genes Dev 2008, 22, 239–251. [Google Scholar]
  281. Malagelada, C.; Ryu, E.J.; Biswas, S.C.; Jackson-Lewis, V.; Greene, L.A. RTP801 is elevated in Parkinson brain substantia nigral neurons and mediates death in cellular models of Parkinson’s disease by a mechanism involving mammalian target of rapamycin inactivation. J. Neurosci 2006, 26, 9996–10005. [Google Scholar]
  282. Imai, Y.; Gehrke, S.; Wang, H.Q.; Takahashi, R.; Hasegawa, K.; Oota, E.; Lu, B. Phosphorylation of 4E-BP by LRRK2 affects the maintenance of dopaminergic neurons in Drosophila. EMBO J 2008, 27, 2432–2443. [Google Scholar]
  283. Santini, E.; Heiman, M.; Greengard, P.; Valjent, E.; Fisone, G. Inhibition of mTOR signaling in Parkinson’s disease prevents L-DOPA-induced dyskinesia. Sci. Signal. 2009, 2. [Google Scholar] [CrossRef]
  284. Fokas, E.; Yoshimura, M.; Prevo, R.; Higgins, G.; Hackl, W.; Maira, S.M.; Bernhard, E.J.; McKenna, W.G.; Muschel, R.J. NVP-BEZ235 and NVP-BGT226, dual phosphatidylinositol 3-kinase/Mammalian target of rapamycin inhibitors, enhance tumor and endothelial cell radiosensitivity. Radiat. Oncol 2012, 7, 48. [Google Scholar]
  285. Maiese, K.; Hou, J.; Chong, Z.Z.; Shang, Y.C. A fork in the path: Developing therapeutic inroads with FoxO proteins. Oxid. Med. Cell. Longev 2009, 2, 119–129. [Google Scholar]
  286. Musteanu, M.; Blaas, L.; Zenz, R.; Svinka, J.; Hoffmann, T.; Grabner, B.; Schramek, D.; Kantner, H.P.; Muller, M.; Kolbe, T.; et al. A mouse model to identify cooperating signaling pathways in cancer. Nat. Methods 2012, 9, 897–900. [Google Scholar]
  287. Pavel, M.E.; Hainsworth, J.D.; Baudin, E.; Peeters, M.; Horsch, D.; Winkler, R.E.; Klimovsky, J.; Lebwohl, D.; Jehl, V.; Wolin, E.M.; et al. Everolimus plus octreotide long-acting repeatable for the treatment of advanced neuroendocrine tumours associated with carcinoid syndrome (RADIANT-2): A randomised, placebo-controlled, phase 3 study. Lancet 2011, 378, 2005–2012. [Google Scholar]
  288. Baryawno, N.; Sveinbjornsson, B.; Eksborg, S.; Chen, C.S.; Kogner, P.; Johnsen, J.I. Small-molecule inhibitors of phosphatidylinositol 3-kinase/Akt signaling inhibit Wnt/beta-catenin pathway cross-talk and suppress medulloblastoma growth. Cancer Res 2010, 70, 266–276. [Google Scholar]
  289. Majumder, S.; Richardson, A.; Strong, R.; Oddo, S. Inducing autophagy by rapamycin before, but not after, the formation of plaques and tangles ameliorates cognitive deficits. PLoS One 2011, 6, e25416. [Google Scholar]
  290. Pattingre, S.; Tassa, A.; Qu, X.; Garuti, R.; Liang, X.H.; Mizushima, N.; Packer, M.; Schneider, M.D.; Levine, B. Bcl-2 antiapoptotic proteins inhibit Beclin 1-dependent autophagy. Cell 2005, 122, 927–939. [Google Scholar]
  291. Carayol, N.; Vakana, E.; Sassano, A.; Kaur, S.; Goussetis, D.J.; Glaser, H.; Druker, B.J.; Donato, N.J.; Altman, J.K.; Barr, S.; et al. Critical roles for mTORC2- and rapamycin-insensitive mTORC1-complexes in growth and survival of BCR-ABL-expressing leukemic cells. Proc. Natl. Acad. Sci. USA 2010, 107, 12469–12474. [Google Scholar]
Figure 1. Signal transduction pathways of the PI 3-K, Akt, and mTOR cascade. During oxidative stress, multiple pathways are affected that involve PI 3-K, Akt, and mTOR that ultimately interface with programmed cell death pathways of apoptosis and autophagy. Activation of phosphoinositide 3 kinase (PI 3-K), such as by tropic factors that include erythropoietin can promote the production of phosphatidylinositide (3,4)-biphosphate (PI-3,4-P2) and phosphatidylinositide (3,4,5)-triphosphate (PI-3,4,5-P3) that recruits Akt to the plasma membrane. This recruitment activates phosphoinositide dependent kinase 1 (PDK1) and PDK2, leading to Akt phosphorylation. Akt activity can be blocked by the phosphatase and tensin homolog deleted from chromosome 10 (PTEN), SH2 domain-containing inositol phosphatase (SHIP), and carboxyl-terminal modulator protein (CTMP). Akt activity can be enhanced by the T cell leukemia/lymphoma 1 (TCL1) and 90 kDa heat shock protein (Hsp90) that can inhibit protein phosphatase 2A (PP2A). Akt can activate mTORC1 through phosphorylating TSC2 and disrupting the interaction between TSC2 and TSC1. Akt may also activate mTORC1 through I-kappaB kinase (IKK). IKKα associates with Raptor and IKKβ that can phosphorylate TSC1 and suppress TSC1 and its interaction with TSC2. In addition, Akt can directly phosphorylate proline rich Akt substrate 40 kDa (PRAS40) to reduce PRAS40 binding to regulatory associated protein of mTOR (Raptor) and thereby activate mTORC1. Upon activation, mTORC1 phosphorylates its downstream targets p70 ribosome S6 kinase (p70S6K) to phosphorylate pro-apoptotic protein BAD and increase the expression of Bcl-2/Bcl-xL which functions as an anti-apoptotic protein. mTORC1 activation also inhibits autophagic proteins autophagy related gene 13 (Atg13) and UNC-51 like kinase 1/2(ULK1/2) through phosphorylation to prevent autophagy. Rapamycin, an inhibitor of mTOR, can prevent this process and foster autophagy. mTOR signaling inhibits apoptosis though activation of Akt that inhibits “pro-apoptotic” proteins FoxO3a, glycogen synthase-3β (GSK-3β), BAD, and PRAS40.
Figure 1. Signal transduction pathways of the PI 3-K, Akt, and mTOR cascade. During oxidative stress, multiple pathways are affected that involve PI 3-K, Akt, and mTOR that ultimately interface with programmed cell death pathways of apoptosis and autophagy. Activation of phosphoinositide 3 kinase (PI 3-K), such as by tropic factors that include erythropoietin can promote the production of phosphatidylinositide (3,4)-biphosphate (PI-3,4-P2) and phosphatidylinositide (3,4,5)-triphosphate (PI-3,4,5-P3) that recruits Akt to the plasma membrane. This recruitment activates phosphoinositide dependent kinase 1 (PDK1) and PDK2, leading to Akt phosphorylation. Akt activity can be blocked by the phosphatase and tensin homolog deleted from chromosome 10 (PTEN), SH2 domain-containing inositol phosphatase (SHIP), and carboxyl-terminal modulator protein (CTMP). Akt activity can be enhanced by the T cell leukemia/lymphoma 1 (TCL1) and 90 kDa heat shock protein (Hsp90) that can inhibit protein phosphatase 2A (PP2A). Akt can activate mTORC1 through phosphorylating TSC2 and disrupting the interaction between TSC2 and TSC1. Akt may also activate mTORC1 through I-kappaB kinase (IKK). IKKα associates with Raptor and IKKβ that can phosphorylate TSC1 and suppress TSC1 and its interaction with TSC2. In addition, Akt can directly phosphorylate proline rich Akt substrate 40 kDa (PRAS40) to reduce PRAS40 binding to regulatory associated protein of mTOR (Raptor) and thereby activate mTORC1. Upon activation, mTORC1 phosphorylates its downstream targets p70 ribosome S6 kinase (p70S6K) to phosphorylate pro-apoptotic protein BAD and increase the expression of Bcl-2/Bcl-xL which functions as an anti-apoptotic protein. mTORC1 activation also inhibits autophagic proteins autophagy related gene 13 (Atg13) and UNC-51 like kinase 1/2(ULK1/2) through phosphorylation to prevent autophagy. Rapamycin, an inhibitor of mTOR, can prevent this process and foster autophagy. mTOR signaling inhibits apoptosis though activation of Akt that inhibits “pro-apoptotic” proteins FoxO3a, glycogen synthase-3β (GSK-3β), BAD, and PRAS40.
Ijms 13 13830f1
Figure 2. Modulation of neurodegenerative disorders through PI 3-K, Akt, mTOR, and associated pathways of SIRT1. Oxidative stress leads to cell injury in multiple neurodegenerative disorders. In Parkinson’s disease (PD), oxidative stress can lead to the induction of the stress response protein REDD1 that can inhibit the activation of mTOR. The accumulation of amyloid (Aβ) during Alzheimer’s disease (AD) also can block the activation of mTOR. In AD, retinoblastoma tumor suppressor (RB1) inducible Coiled-Coil 1 (RB1CC1), which functions to activate mTOR, is reduced, contributing to neuronal atrophy in AD. The activation of the downstream target of mTOR, p70 ribosome S6 kinase (p70S6K), by phosphorylation (p) prevents acute neuronal injury during stroke. However, inhibition of mTOR and p70S6K is required to promote autophagy and the clearance of aggregate prone proteins, such as α-synuclein, Aβ, and Huntingtin to prevent neuronal loss. A fine balance of mTOR activation is necessary in these disorders since mTOR can lead to dyskinesia in PD and activation of p70S6K has been associated with the promotion of the phosphorylation of tau protein contributing to formation of neurofibrillary tangles. During diabetes mellitus (DM), increasing oxidative stress results in insulin resistance, which can be ameliorated by SIRT1. Activation of SIRT1 can increase the secretion of insulin by repressing the mitochondrial uncoupling protein 2 (UCP2), promoting lipolysis, and increasing gluconeogenesis. SIRT1 also can increase insulin sensitivity by inhibiting tyrosine phosphatase 1B (PTP1B). Elevated levels of oxidative stress can reduce insulin sensitivity and enhance the activity of FoxO3a. FoxO3a has multiple roles that can interact with SIRT1 and influence β cell function as well as modulate inflammation. SIRT1 can increase FoxO3a activity through deacetylation (dAc). SIRT1 also can activate Akt that decreases the activity of FoxO3a through phosphorylation (p). Enhanced activity of mTOR increases insulin secretion, induces adipogenesis, and inhibits lipolysis that can influence nervous system complications of DM.
Figure 2. Modulation of neurodegenerative disorders through PI 3-K, Akt, mTOR, and associated pathways of SIRT1. Oxidative stress leads to cell injury in multiple neurodegenerative disorders. In Parkinson’s disease (PD), oxidative stress can lead to the induction of the stress response protein REDD1 that can inhibit the activation of mTOR. The accumulation of amyloid (Aβ) during Alzheimer’s disease (AD) also can block the activation of mTOR. In AD, retinoblastoma tumor suppressor (RB1) inducible Coiled-Coil 1 (RB1CC1), which functions to activate mTOR, is reduced, contributing to neuronal atrophy in AD. The activation of the downstream target of mTOR, p70 ribosome S6 kinase (p70S6K), by phosphorylation (p) prevents acute neuronal injury during stroke. However, inhibition of mTOR and p70S6K is required to promote autophagy and the clearance of aggregate prone proteins, such as α-synuclein, Aβ, and Huntingtin to prevent neuronal loss. A fine balance of mTOR activation is necessary in these disorders since mTOR can lead to dyskinesia in PD and activation of p70S6K has been associated with the promotion of the phosphorylation of tau protein contributing to formation of neurofibrillary tangles. During diabetes mellitus (DM), increasing oxidative stress results in insulin resistance, which can be ameliorated by SIRT1. Activation of SIRT1 can increase the secretion of insulin by repressing the mitochondrial uncoupling protein 2 (UCP2), promoting lipolysis, and increasing gluconeogenesis. SIRT1 also can increase insulin sensitivity by inhibiting tyrosine phosphatase 1B (PTP1B). Elevated levels of oxidative stress can reduce insulin sensitivity and enhance the activity of FoxO3a. FoxO3a has multiple roles that can interact with SIRT1 and influence β cell function as well as modulate inflammation. SIRT1 can increase FoxO3a activity through deacetylation (dAc). SIRT1 also can activate Akt that decreases the activity of FoxO3a through phosphorylation (p). Enhanced activity of mTOR increases insulin secretion, induces adipogenesis, and inhibits lipolysis that can influence nervous system complications of DM.
Ijms 13 13830f2

Share and Cite

MDPI and ACS Style

Maiese, K.; Chong, Z.Z.; Wang, S.; Shang, Y.C. Oxidant Stress and Signal Transduction in the Nervous System with the PI 3-K, Akt, and mTOR Cascade. Int. J. Mol. Sci. 2012, 13, 13830-13866. https://doi.org/10.3390/ijms131113830

AMA Style

Maiese K, Chong ZZ, Wang S, Shang YC. Oxidant Stress and Signal Transduction in the Nervous System with the PI 3-K, Akt, and mTOR Cascade. International Journal of Molecular Sciences. 2012; 13(11):13830-13866. https://doi.org/10.3390/ijms131113830

Chicago/Turabian Style

Maiese, Kenneth, Zhao Zhong Chong, Shaohui Wang, and Yan Chen Shang. 2012. "Oxidant Stress and Signal Transduction in the Nervous System with the PI 3-K, Akt, and mTOR Cascade" International Journal of Molecular Sciences 13, no. 11: 13830-13866. https://doi.org/10.3390/ijms131113830

Article Metrics

Back to TopTop