Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

PRAS40 Is an Integral Regulatory Component of Erythropoietin mTOR Signaling and Cytoprotection

  • Zhao Zhong Chong,

    Affiliations Laboratory of Cellular and Molecular Signaling, Newark, New Jersey, United States of America, New Jersey Health Sciences University, Newark, New Jersey, United States of America

  • Yan Chen Shang,

    Affiliations Laboratory of Cellular and Molecular Signaling, Newark, New Jersey, United States of America, New Jersey Health Sciences University, Newark, New Jersey, United States of America

  • Shaohui Wang,

    Affiliations Laboratory of Cellular and Molecular Signaling, Newark, New Jersey, United States of America, New Jersey Health Sciences University, Newark, New Jersey, United States of America

  • Kenneth Maiese

    wntin75@yahoo.com

    Affiliations Laboratory of Cellular and Molecular Signaling, Newark, New Jersey, United States of America, Cancer Institute of New Jersey, New Brunswick, New Jersey, United States of America, New Jersey Health Sciences University, Newark, New Jersey, United States of America

Abstract

Emerging strategies that center upon the mammalian target of rapamycin (mTOR) signaling for neurodegenerative disorders may bring effective treatment for a number of difficult disease entities. Here we show that erythropoietin (EPO), a novel agent for nervous system disorders, prevents apoptotic SH-SY5Y cell injury in an oxidative stress model of oxygen-glucose deprivation through phosphatidylinositol-3-kinase (PI 3-K)/protein kinase B (Akt) dependent activation of mTOR signaling and phosphorylation of the downstream pathways of p70 ribosomal S6 kinase (p70S6K), eukaryotic initiation factor 4E-binding protein 1 (4EBP1), and proline rich Akt substrate 40 kDa (PRAS40). PRAS40 is an important regulatory component either alone or in conjunction with EPO signal transduction that can determine cell survival through apoptotic caspase 3 activation. EPO and the PI 3-K/Akt pathways control cell survival and mTOR activity through the inhibitory post-translational phosphorylation of PRAS40 that leads to subcellular binding of PRAS40 to the cytoplasmic docking protein 14-3-3. However, modulation and phosphorylation of PRAS40 is independent of other protective pathways of EPO that involve extracellular signal related kinase (ERK 1/2) and signal transducer and activator of transcription (STAT5). Our studies highlight EPO and PRAS40 signaling in the mTOR pathway as potential therapeutic strategies for development against degenerative disorders that lead to cell demise.

Introduction

Neurodegenerative disease leads to either severe disability or death for a significant proportion of the world's population. For example, in regards to cognitive disease, it is estimated that greater than twenty-four million people are afflicted with Alzheimer's disease, pre-senile dementia, and associated disease that involve memory loss [1], [2], [3]. Although multiple factors may contribute to the onset and progression of neurodegenerative disease, oxidative stress is considered to be an important component in neurodegenerative disorders. Oxidative stress can lead to cognitive disorders [4], [5], [6], movement disorders [5], [7], [8], and neurovascular complications associated with metabolic disease [9], [10], [11].

Given that effective treatments for the majority of neurodegenerative disorders do not exist, new strategies that can offer protection in the nervous system during oxidative stress are of great interest [12], [13]. In particular, erythropoietin (EPO) represents a novel therapy that may provide robust protection for both neuronal and non-neuronal cells in the nervous system. EPO prevents neuronal cell injury [14], [15], [16], [17], [18], [19], [20], maintains vascular integrity [21], [22], [23], [24], and modulates inflammatory cell activation [25], [26], [27], [28]. EPO promotes cellular survival through the phosphatidylinositol-3-kinase (PI 3-K) and protein kinase B (Akt) pathways [29], [30], [31], [32]. More recent studies have demonstrated that EPO also relies upon mammalian target of rapamycin (mTOR) signaling to modulate inflammatory cell survival [27], [33], osteoblastogenesis, and osteoclastogenesis [34].

In a number of scenarios, mTOR activation may be necessary to prevent apoptotic neuronal cell death during oxidative stress. Cell death following exposure to oxidative stress in dopaminergic neurons can be prevented during application of agents that increase mTOR activity [35]. In contrast, loss of mTOR activity during oxidative stress leads to apoptotic neuronal death [36] and injury in non-neuronal inflammatory cells [33], [37]. One of the central pathways that can control mTOR signaling is the proline rich Akt substrate 40 kDa (PRAS40). Through mTOR Complex 1 (mTORC1), PRAS40 prevents mTOR activity and inhibits the binding of the downstream mTOR proteins p70 ribosomal S6 kinase (p70S6K) and the eukaryotic initiation factor 4E-binding protein 1 (4EBP1) to Raptor [38], [39], [40]. PRAS40 activity is inhibited during post-translational phosphorylation [41] and this has been associated with increased cell survival [42], [43], [44]. We therefore examined if PRAS40 was a critical regulatory pathway for EPO to foster neuroprotection during oxidative stress. We show that in a model of oxygen-glucose deprivation (OGD) that can lead to oxidative stress [45], [46], EPO activates mTOR signaling through PI 3-K/Akt pathways to phosphorylate p70S6K and 4EBP1 that is necessary for protection in differentiated SH-SY5Y cells. EPO controls cell survival and mTOR activity through the post-translational phosphorylation of PRAS40 and the binding of PRAS40 to 14-3-3 protein. Furthermore, inhibition of PRAS40 is an integral cytoprotective component of EPO that can increase cell survival and limit apoptotic caspase 3 activity independent of other protective pathways of EPO that involve extracellular signal related kinase (ERK 1/2) and signal transducer and activator of transcription (STAT5). Our work highlights PRAS40 in the cytoprotective pathways of EPO as a potential target for novel therapeutic strategies directed against degenerative disorders.

Materials and Methods

Human neuroblastoma SH-SY5Y cell culture and differentiation

Per our prior protocols [47], [48], human adrenergic neuroblastoma SH-SY5Y cells were purchased from ATCC (American Type Culture Collection) and maintained in regular Dulbecco's modified Eagle medium (DMEM) (Life Technologies Corp, Carlsbad, CA), supplemented with 10% heat-inactivated fetal bovine serum, 1 mM pyruvate, 1.5 g/L sodium bicarbonate, 100 IU/ml penicillin, 100 µg/ml streptomycin at 37°C in 95%/5% (v/v) mixture of humidified atmospheric air and CO2. Cell suspension was prepared at a density of 3–4×104 (24 well plate) or 1–1.5×105 (35 mm2 Petri dish). When confluent at 50–60%, cells were differentiated by MEM growth medium containing 10 μM all-trans retinoic acid (RA) (Sigma, St. Louis, MO) for 48 h. Experiments were initiated until cells grew to 60% – 70% confluence between passages 4–10 after differentiation.

Experimental treatments

Per our prior experimental protocols, oxygen-glucose deprivation (OGD) in SH-SY5Y cells was performed by replacing the media of the cultures in 35 mm2 dishes with cells of 60–70% confluence with glucose-free Hank's balanced salt solution (HBSS) containing 116 mmol/l NaCl, 5.4 mmol/l KCl, 0.8 mmol/l MgSO4, 1 mmol/l NaH2PO4, 0.9 mmol/l CaCl2, and 10 mg/l phenol red (pH 7.4). SH-SY5Y cultures were then placed into a Bactron II anaerobic glove box (Sheldon manufacturing, Inc, Cornelius, OR) and were maintained in an anoxic environment (95% N2 and 5% CO2) at 37°C for 6 hours. Following this period, the cultures were removed from the anoxic chamber and the glucose-free HBSS was replaced with media containing Dulbecco's modified Eagle medium (DMEM) (Life Technologies Corp, Carlsbad, CA), supplemented with 10% heat-inactivated fetal bovine serum, 1 mM pyruvate, 1.5 g/L sodium bicarbonate, 100 IU/ml penicillin, and 100 µg/ml streptomycin and maintained at 37°C in 95%/5% (v/v) mixture of humidified atmospheric air and CO2. For treatments applied prior to OGD, erythropoietin (Sigma, St. Louis, MO), the phosphatidylinositol-3-kinase (PI 3-K) inhibitors wortmannin (0.5 μM, EMD Biochemicals Inc, Gibbstown, NJ) and LY294002 (10 μM, Sigma, St Louis, MO), extracellular signal-regulated kinase (ERK) inhibitor 3-(2-aminoethyl)-5-((4-ethoxyphenyl) methylene)-2,4-thiazolidinedione (ERK-1) (100 μM) (EMD Biochemicals Inc, Gibbstown, NJ), HCl (ERK-I) (EMD Biochemicals Inc, Gibbstown, NJ), and signal transducer and activator of transcription- 5 (STAT5) inhibitor N′-((4-Oxo-4H-chromen-3-yl)methylene) nicotinohydrazide (STAT5-I) (100 μM) (EMD Biochemicals Inc, Gibbstown, NJ) were each administered directly to the cultures 1 hour prior to OGD and treatments were continuous.

Assessment of cell survival

SH-SY5Y cell injury was determined by bright field microscopy using a 0.4% trypan blue dye exclusion method 24 hours following treatment with OGD per our previous protocols [22], [49]. For each experimental condition, 8×35 mm2 dishes were used, and for each dish, the mean survival was determined by counting eight randomly selected non-overlapping fields with each containing approximately 20 cells (viable + non-viable). Each experiment was replicated 6 times with different cultures.

Assessment of DNA Fragmentation

Genomic DNA fragmentation was determined by the terminal deoxynucleotidyl transferase nick end labeling (TUNEL) assay [47], [50], [51]. Briefly, SH-SY5Y cells were fixed in 4% paraformaldehyde/0.2% picric acid/0.05% glutaraldehyde and the 3′-hydroxy ends of cut DNA were labeled with biotinylated dUTP using the enzyme terminal deoxytransferase (Promega, Madison, WI) followed by streptavidin-peroxidase and visualized with 3,3′-diaminobenzidine (Vector Laboratories, Burlingame, CA).

Expression of mTOR, p70S6K, Akt1, STAT5, 4EBP1, PRAS40, ERK1/2, and caspase 3 with relevant phosphorylated moieties

Cells were homogenized and following protein determination, each sample (25–50 μg/lane) was then subjected to 7.5% (p-mTOR, mTOR, p-p70S6K, p70S6K, p-Akt1, Akt1, and p-STAT5, STAT5) or 12.5% (p-4EBP1, 4EBP1, PRAS40, p-PRAS40, ERK 1/2, p-ERK 1/2, and caspase 3) SDS-polyacrylamide gel electrophoresis separation. After blocking for 1 hour at room temperature with 5% skim milk, the membranes were incubated overnight at 4°C with a rabbit antibody against (p-  =  phosphorylated) p-mTOR (Ser 2448, 1∶1000), mTOR (1∶1000), p-p70S6K (Thr389, 1∶1000), p70S6K (1∶1000), p-Akt1 (Ser473, 1∶1000), Akt1 (1∶1000), p-STAT5 (Tyr694, 1∶1000), STAT5 (1∶1000), p-4EBP1 (Ser 65/Thr70, 1∶1000), 4EBP1 (1∶1000), PRAS40 (1∶1000), p-PRAS40 (Thr246, 1∶1000), p-ERK 1/2 (Thr202/Tyr204), and cleaved caspase 3 (1∶1000). All antibodies were obtained from Cell Signaling, Beverly, MA. Following incubation, the membranes were incubated with a horseradish peroxidase (HRP) conjugated secondary antibody goat anti-rabbit IgG (goat anti-rabbit IgG, 1∶5000, Thermo Scientific, Rockford, IL). The antibody-reactive bands were revealed by chemiluminescence (Amersham Pharmacia Biotech, Piscataway, NJ) and band density was performed using the public domain NIH Image program (developed at the U.S. National Institutes of Health and available at http://rsb.info.nih.gov/nih-image/).

In vitro assay of phosphorylation of mTOR and PRAS40

Recombinant human mTOR (EMD Biochemicals Inc, Gibbstown, NJ) or PRAS40 protein (Enzo Life Sciences, Plymouth Meeting, PA) 1 μg was incubated with 10 ng/ml EPO for 30 min at 30°C under continuous agitation in 30 μl kinase buffer containing 200 μM ATP (Cell Signaling Technology, Beverly, MA). Samples were analyzed by Western blot analysis using SDS-polyacrylamide gel and rabbit antibody against p-mTOR (Ser 2448) or p-PRAS40 (Thr246) (Cell Signaling Technology, Beverly, MA).

Gene silencing of Akt1 and PRAS40 with small interfering RNA (siRNA)

Cells were plated into 35 mm dishes or 24-well plates. To silence Akt1 and PRAS40 gene expression, commercial reagents using the SMARTpool Akt1 siRNA kit (Millipore, Billerica, MA) and siRNA pool for PRAS40 (Santa Cruz, Santa Cruz, CA) were used respectively. Transfection of siRNA duplexes was performed with Lipofectamine 2000 reagent according to manufacturer guidelines (Life Technologies, Carlsbad, CA). Experimental assays were performed 72 hours post-transfection. For each siRNA assay, scrambled siRNA was used as control.

Immunoprecipitation of 14-3-3, mTOR or PRAS40

Cell lysates of total protein (200 μg) were incubated with primary antibody against protein 14-3-3 (1∶100, Santa Cruz Biotech, Santa Cruz, CA) or mTOR (1∶100, Cell Signaling Technology, Beverly, MA) overnight at 4°C. The complexes were collected with protein A/G-agarose beads, centrifuged, and then prepared for 14-3-3, mTOR, PRAS40, and p-PRAS40 western analysis.

Statistical analysis

For each experiment, the mean and standard deviation (SD) was determined. Statistical differences among groups were assessed by means of analysis of variance (ANOVA) with the post-hoc Dunnett's test. Statistical significance was considered at P<0.05.

Results

EPO prevents cellular injury and apoptotic genomic DNA degradation during oxygen glucose deprivation (OGD)

EPO (10 ng/ml) was administered to cell cultures 1 hour prior to a 6 hour period of OGD and cell injury was determined 24 hours later through trypan blue dye exclusion method and TUNEL assay. This concentration of EPO was chosen since it previously was shown to provide significant cytoprotection in neuronal cells and vascular cells [21], [22], [50], [52]. In Figures 1A and 1B, untreated SH-SY5Y cells were not significantly stained with trypan blue and TUNEL, but exposure to OGD resulted in significant trypan blue staining and nuclear DNA damage 24 hours following OGD exposure in neurons. In contrast, EPO (10 ng/ml) significantly reduced trypan blue staining and nuclear DNA degradation.

EPO provides cellular protection through mTOR and its signaling pathways

Application of rapamycin (50 nM) or transfection with mTOR siRNA during EPO (10 ng/ml) exposure increased cell injury and DNA fragmentation following OGD when compared to OGD exposure alone (Figures 1A and 1B). Protection by EPO was significantly reduced during blockade of mTOR and its signaling pathways, suggesting that EPO relies upon mTOR to offer cellular protection during oxidant stress. As a control, non-specific scrambled siRNA during EPO treatment did not alter survival or DNA fragmentation when compared to EPO treatment and OGD exposure alone.

thumbnail
Figure 1. EPO promotes cellular protection through mTOR activation and phosphorylation of p70S6K and 4EBP1.

(A) EPO (10 ng/ml) was applied to SH-SY5Y cultures 1 hour prior to a 6 hour-period of OGD and cell survival and DNA fragmentation was determined 24 hours later through trypan blue dye exclusion (TB) and TUNEL assay. Representative images show that EPO significantly reduces trypan blue and TUNEL staining following OGD. However, application of the mTOR inhibitor rapamycin (50 nM) or transfection of mTOR siRNA prior to EPO administration resulted in the loss of EPO cytoprotection during OGD. (B) Quantitative results illustrate that EPO (10 ng/ml) significantly reduces percent cell staining of trypan blue (TB) and TUNEL in SH-SY5Y cells 24 hours following OGD. Rapamycin (50 nM) or mTOR siRNA transfection prior to OGD attenuated the ability of EPO to offer cytoprotection during OGD leading to an increase in the percent staining of trypan blue and TUNEL. Transfection of scrambled siRNA did not alter EPO cytoprotection or cell injury when compared to treatment with EPO or OGD alone (*P<0.01 vs. OGD; P<0.01 vs. EPO/OGD). For B, each data point represents the mean and SD from 6 experiments. (C) Western blot was performed for phosphorylated (p-) mTOR (p-mTOR, Ser2448), phosphorylated (p-)-p70S6K (p-p70S6K, Thr389), phosphorylated (p-)-4EBP1 (p-4EBP1, Ser65/Thr70) in SH-SY5Y cells at 1, 3, or 24 hours (Hrs) following a 6 hour period of OGD exposure. OGD resulted in a transient increase in the expression of p-mTOR, p-p70S6K, and p-4EBP1 at 1 and 3 hours (*P<0.01 vs. Control). (D) EPO (10 ng/ml) administration to SH-SY5Y cells significantly increased the expression of p-mTOR, p-p70S6K, and p-4EBP1 3 hours later. EPO (10 ng/ml) applied to SH-SY5Y cells 1 hour prior to a 6 hour period of OGD significantly increased the expression of p-mTOR, p-p70S6K, and p-4EBP1 3 hours following OGD when compared to OGD treated alone (*P<0.01 vs. untreated control; P<0.01 vs. OGD treated alone). For C and D, Cont =  Control and each data point represents the mean and SD from 3 experiments.

https://doi.org/10.1371/journal.pone.0045456.g001

Treatment with EPO leads to mTOR activation and phosphorylation of p70S6K and 4EBP1

Given that inhibition of mTOR can block cytoprotection by EPO during OGD, we investigated whether EPO can activate mTOR and control the activity of its downstream targets p70S6K and 4EBP1. The carboxy-terminal (C-terminal) kinase domain of mTOR consists of a conserved sequence with homology to the catalytic domain of phosphoinositide 3 –kinase (PI 3-K) with phosphorylation sites of mTOR for its activation [41], [53] that include serine2448 [54]. Serine2448 is an important target for both Akt and p70S6K [54], [55]. mTOR phosphorylates and activates p70S6K at threonine389 that serves as a marker of mTOR activity [56]. 4EBP1 is phosphorylated by mTOR at serine65 and threonine70 [57]. Phosphorylation of 4EBP1 leads to the dissociation of 4EBP1 from eukaryotic translation initiation factor 4 epsilon (eIF4E) to allow the eukaryotic translation initiation factor 4 gamma (eIF4G) to begin mRNA translation [58], [59].

We assessed the expression of phosphorylated mTOR (p-mTOR, Ser2448, active form) and phosphorylated forms of its downstream targets p-p70S6K and p-4EBP1 (p-p70S6K, Thr389; p-4EBP1, Ser65/Thr70) during OGD exposure. In Figure 1C, the expression of p-mTOR, p-p70S6K, and p-4EBP1 was significantly increased at 1 and 3 hours following OGD exposure, but returned to the level of untreated controls within 24 hours following OGD exposure. In the next series of studies, treatment with EPO (10 ng/ml) alone significantly increased the expression of p-mTOR, p-p70S6K, and p-4EBP1 within 3 hours after EPO exposure (Figure 1D).

EPO employs mTOR to phosphorylate p70S6K and 4EBP1

EPO (10 ng/ml) given 1 hour prior to OGD also significantly increased the expression of phosphorylated p-mTOR, p-p70S6K, and p-4EBP1 3 hours following OGD exposure, illustrating that EPO activates mTOR and p70S6K through phosphorylation but inhibits 4EBP1 activation through phosphorylation. The phosphorylation of mTOR, p70S6K, and 4EBP1 during either exposure to OGD alone or exposure to EPO and OGD were prevented during rapamycin administration or the transfection with mTOR siRNA (Figure 2A). During loss of mTOR activity or the gene silencing of mTOR, phosphorylation of mTOR, p70S6K, and 4EBP1 is minimal or absent during administration of EPO with OGD, demonstrating that EPO relies upon mTOR to phosphorylate p70S6K and 4EBP1 (Figure 2A). Furthermore, transfection with mTOR siRNA is confirmed by western analysis to prevent the expression of phosphorylated p-mTOR and total mTOR (Figure 2A).

thumbnail
Figure 2. EPO phosphorylation of mTOR, p70S6K, and 4EBP1 is dependent upon PI 3-K and Akt.

(A) Rapamycin (50 nM) administration or mTOR siRNA transfection during OGD exposure alone or treatment with EPO (10 ng/ml) applied 1 hour prior to OGD prevented the phosphorylation (p-) of p-mTOR, p-p70S6K, and p-4EBP1 3 hours following OGD exposure. Transfection with mTOR siRNA also significantly limited the expression of total mTOR (*P<0.01 vs. OGD treated alone; P<0.01 vs. EPO/OGD). (B) EPO (10 ng/ml) was incubated with recombinant mTOR protein (1 μg) for 30, 60, and 180 min. No significant expression of phosphorylated (p-) mTOR was detected. (C) Akt1 siRNA transfection in SH-SY5Y cells prior to OGD or prior to treatment with EPO (10 ng/ml) applied 1 hour prior to OGD significantly reduced the expression of total Akt1, phosphorylated (p-) p-mTOR, p-p70S6K, and p-4EBP1 3 hours following OGD (*P<0.01 vs. OGD treated alone; P<0.01 vs. EPO/OGD). (D) EPO (10 ng/ml) or EPO combined with wortmannin (500 nM) or LY294002 (20 μM) was applied to SH-SY5Y cells. Western blot analysis was performed to detect the expression of phosphorylated (p-) p-p70S6K and p-4EBP1 3 hours later. EPO phosphorylation (p-) of p-p70S6K and p-4EBP1 was blocked by the PI 3-K inhibitors wortmannin or LY294002 (*P<0.01 vs. EPO treated alone). (E) EPO (10 ng/ml) or EPO combined with wortmannin (500 nM) or LY294002 (20 μM) was applied to SH-SY5Y cells 1 hour prior to OGD and western blot was performed to detect the expression phosphorylated (p-) p-p70S6K and p-4EBP1 3 hours following OGD. EPO phosphorylation (p-) of p-p70S6K and p-4EBP1 during OGD was blocked by the PI 3-K inhibitors wortmannin or LY294002 (*P<0.01 vs. OGD treated alone; P<0.01 vs. EPO/OGD). In all cases above, each data point represents the mean and SD from 3 experiments.

https://doi.org/10.1371/journal.pone.0045456.g002

Phosphorylation of mTOR, p70S6K, and 4EBP1 by EPO are dependent upon PI 3-K and Akt pathways

Under some conditions, pathways for cellular proliferation and survival require the involvement of mTOR and phosphoinositide 3 –kinase (PI 3-K)/Akt pathways [60]. Since EPO employs mTOR to modulate the phosphorylation of p-p70S6K and p-4EBP1 (Figure 2A), we assessed whether EPO also relies upon the PI 3-K/Akt pathway to phosphorylate p-p70S6K and p-4EBP1. We initially examined whether EPO could directly phosphorylate mTOR. In Figure 2B, recombinant mTOR protein was incubated with EPO (10 ng/ml) for 30, 60, and 180 minutes in kinase assay buffer with western analysis subsequently performed. No significant expression of p-mTOR was detected over a 3 hour period of incubation, illustrating that EPO does not directly phosphorylate and activate mTOR. In Figure 2C, phosphorylation of p-mTOR by EPO (10 ng/ml) during OGD exposure was prevented with application of the PI 3-K inhibitors wortmannin (500 nM) or LY294002 (20 μM), illustrating that EPO phosphorylation of mTOR was dependent upon activation of the PI 3-K pathway. The inhibitor wortmannin forms a covalent link with the lysine residue of PI 3-K [61] and the inhibitor LY294002 (20 μM) reversibly competes for ATP binding to inhibit the PI 3-K pathway [62]. In addition, transfection of Akt1 siRNA also prevents EPO (10 ng/ml) phosphorylation of p-mTOR, p-p70S6K, and p-4EBP1 during OGD exposure (Figure 2C), demonstrating that Akt1 also is required for EPO to phosphorylate and activate mTOR as well as phosphorylate p-p70S6K and p-4EBP1. In Figures 2D and 2E, EPO (10 ng/ml) significantly increased the expression of phosphorylated p-p70S6K and p-4EBP1 alone and during exposure to OGD (Figures 2D and 2E). However, phosphorylation of p-p70S6K and p-4EBP1 by EPO alone or during OGD exposure was blocked during application of the PI 3-K inhibitors wortmannin (500 nM) and or LY294002 (20 μM), suggesting that EPO also employs the PI 3-K pathway to phosphorylate p70S6K and 4EBP1.

Loss of PRAS40 prevents cell injury, improves EPO cytoprotection, and increases phosphorylation of mTOR, p70S6K, and 4EBP1

PRAS40 can inhibit mTOR signaling by associating with Raptor [39], [40] and prevent p70S6K and 4EBP1 binding to Raptor [38], [40]. We therefore examined whether loss of PRAS40 could alter cell survival and apoptotic DNA degradation following OGD exposure and EPO administration. Cell survival was assessed with trypan dye blue exclusion and DNA degradation with TUNEL 24 hours after OGD exposure. In Figures 3A and 3B, transfection with PRAS40 siRNA during OGD exposure significantly decreased trypan blue staining and TUNEL staining, illustrating that loss of PRAS40 is protective during OGD exposure. In addition, gene silencing of PRAS40 during EPO (10 ng/ml) administration and OGD exposure improved cell survival and limited genomic DNA degradation to a greater extent than EPO alone (Figures 3A and 3B), suggesting that EPO employs inhibition of PRAS40 to some degree to offer cytoprotection during OGD exposure. As a control, non-specific scrambled siRNA during EPO treatment did not alter survival or DNA fragmentation when compared to EPO treatment and OGD exposure alone.

thumbnail
Figure 3. Loss of PRAS40 prevents cell demise, improves EPO cytoprotection, and increases phosphorylation of mTOR, p70S6K, and 4EBP1.

(A) Representative images of trypan blue (TB) dye staining and TUNEL assay performed 24 hours following OGD exposure in SH-SY5Y cells show that PRAS40 siRNA transfection prior to OGD significantly reduced trypan blue and TUNEL staining and mildly improved cytoprotection by EPO (10 ng/ml) applied 1 hour prior to OGD. (B) Quantitative results of cell survival assessed by trypan blue dye exclusion (TB) and DNA fragmentation by TUNEL assay demonstrate that PRAS40 siRNA transfection during OGD alone or during EPO (10 ng/ml) with OGD significantly decreased percent staining of trypan blue and TUNEL 24 hours following OGD (*P<0.01 vs. OGD treated alone). Scrambled siRNA transfection did not alter cell injury during OGD exposure or EPO treatment with OGD. (C and D) PRAS40 siRNA was transfected into SH-SY5Y cells prior to OGD and western blot analysis for phosphorylated (p-) p-mTOR, p-p70S6K, p-4EBP1, and total PRAS40 was performed in cell extracts at 3 hour following OGD. EPO (10 ng/ml) applied 1 hour prior to OGD significantly increased the expression of p-mTOR, p-p70S6K and p-4EBP1. Transfection with PRAS40 siRNA increased, although the increase is not statistically significant, the expression of p-mTOR, p-p70S6K and p-4EBP1 during OGD alone and during EPO (10 ng/ml) with OGD exposure. Transfection with PRAS40 siRNA significantly limited the expression of total PRAS40. Scrambled siRNA did not alter expression of p-mTOR, p-p70S6K. p-4EBP1, total PRAS40 during OGD alone and during EPO (10 ng/ml) with OGD exposure (*P<0.01 vs. OGD; P<0.01 vs. EPO/OGD). In all cases above, each data point represents the mean and SD from 3 experiments.

https://doi.org/10.1371/journal.pone.0045456.g003

In Figure 3C, we examined the effects of PRAS40 gene silencing upon the expression of phosphorylated p-mTOR, p-p70S6K, and p-4EBP1 during OGD exposure and exposure to EPO and OGD. Transfection with PRAS40 siRNA significantly limited the expression of PRAS40 protein and increased the phosphorylation of p-mTOR and its activity as suggested by increased phosphorylation of p-p70S6K and p-4EBP1 (Figure 3C). Loss of PRAS40 with siRNA transfection also increased the phosphorylation of p-mTOR, p-p70S6K, and p-4EBP1 during EPO (10 ng/ml) administration with OGD, suggesting that EPO phosphorylation of mTOR, p70S6K, and 4EBP1 can be fostered by the inhibition or loss of PRAS40 (Figure 3C). As a control, non-specific scrambled siRNA did not alter mTOR, p70S6K, and 4EBP1 phosphorylation.

EPO maintains PRAS40 phosphorylation through PI 3-K and Akt during OGD exposure

Western blot analysis for phosphorylated p-PRAS40 (Thr246) was performed at 1, 3, and 24 hours following OGD exposure. In Figure 4A, phosphorylated p-PRAS40 expression was initially increased within 1 hour following OGD exposure, but over a 24 hour course returned to approximately untreated control levels. In contrast, application of EPO (10 ng/ml) during OGD exposure, phosphorylation of PRAS40 was significantly increased and maintained over a 24 hour course when compared to exposure to OGD alone (Figure 4A). In addition, EPO (10 ng/ml) in normoxic cells not exposed to OGD also significantly increased phosphorylation of PRAS40 within 3 hours of treatment (Figure 4B). To determine whether EPO could directly phosphorylate PRAS40, we incubated recombinant PRAS40 protein with EPO (10 ng/ml) for 30, 60, and 180 minutes in kinase assay buffer and western analysis was subsequently performed. EPO did not directly phosphorylate PRAS40 over a 3 hour time period (Figure 4C).

thumbnail
Figure 4. EPO promotes PRAS40 phosphorylation through PI 3-K and Akt during OGD exposure.

(A) Western blot analysis was performed for phosphorylated (p-)-PRAS40 (p-PRAS40, Thr246) in SH-SY5Y cells at 1, 3, or 24 hours (hrs) following OGD exposure. EPO (10 ng/ml) applied to cell cultures 1 hour prior to OGD maintained p-PRAS40 expression at significant levels over 24 hours following OGD (*P<0.01 vs. Control; P<0.01 vs. OGD at corresponding time points). (B) EPO (10 ng/ml) applied to SH-SY5Y cells significantly increased the expression of phosphorylated (p-) p-PRAS40 3 hours later. EPO (10 ng/ml) applied to SH-SY5Y cultures 1 hour prior OGD significantly increased the expression of p-PRAS40 3 hours following OGD (*P<0.01 vs. untreated control; P<0.01 vs. OGD treated alone). (C) EPO (10 ng/ml) was incubated with recombinant PRAS40 protein for 30, 60, and 180 min. No significant expression of phosphorylated (p-) PRAS40 was detected. (D) EPO (10 ng/ml) or EPO combined with the P I3-K inhibitors wortmannin (500 nM) or LY294002 (20 μM) were applied to SH-SY5Y cells and western blot analysis for phosphorylated (p-) p-PRAS40 and p-Akt1 (p-Akt1, Ser473) was performed 3 hours later. Wortmannin or LY294002 prevented the expression of p-PRAS40 and p-Akt1 during EPO (10 ng/ml) administration (*P<0.01 vs. untreated control; P<0.01 vs. EPO treated alone). (E) EPO (10 ng/ml) was applied to SH-SY5Y cells 1 hour prior to OGD and western blot analysis for phosphorylated (p) p-PRAS40 and p-Akt1 was performed 3 hours following OGD. EPO significantly increased the expression of p-PRAS40 and p-Akt1 during OGD exposure. Wortmannin or LY294002 prevented the phosphorylation of PRAS40 and Akt1 during EPO administration following OGD (*P<0.01 vs. OGD treated alone; P<0.01 vs. EPO/OGD). (F) Transfection of Akt1 siRNA prior to the application of EPO (10 ng/ml) in SH-SY5Y cells significantly limited the expression of Akt1 and significantly reduced the expression of phosphorylated (p-) p-PRAS40 3 hours after administration of EPO. Scrambled siRNA transfection did not alter the expression of Akt1 and p-PRAS40 during EPO application (*P<0.01 vs. untreated control; P<0.01 vs. EPO treated alone). (G) Akt1 siRNA was transfected into SH-SY5Y cells prior to EPO (10 ng/ml) application and OGD exposure. Western analysis expression of phosphorylated (p-) p-PRAS40 and Akt1 was determined 3 hour following OGD. EPO (10 ng/ml) increased p-PRAS40 expression following OGD. Transfection of Akt1 siRNA significantly limited p-PRAS40 expression during OGD alone and during EPO treatment with OGD (*P<0.01 vs. OGD; P<0.01 vs. EPO/OGD). In all cases above, each data point represents the mean and SD from 3 experiments.

https://doi.org/10.1371/journal.pone.0045456.g004

We next examined the role of the PI 3-K pathway for EPO to phosphorylate PRAS40. EPO (10 ng/ml) was administered without (Figure 4D) and with OGD exposure (Figure 4E) and the expression of phosphorylated p-PRAS40 and phosphorylated Akt1 (p-Akt1) was determined 3 hours later. EPO significantly increased the expression of both p-PRAS40 and p-Akt1 either alone or in the presence of OGD exposure (Figures 4D and 4E). Yet, co-administration of the PI 3-K inhibitors wortmannin (500 nM) or LY294002 (20 μM) prevented EPO from significantly phosphorylating PRAS40 or Akt1 either alone or during OGD exposure (Figures 4D and 4E), illustrating that EPO was dependent upon the PI 3-K pathway to phosphorylate PRAS40.

To assess whether Akt1 played a role during EPO phosphorylation of PRAS40, we examined the effects of gene silencing of Akt1 on PRAS40 phosphorylation during EPO administration alone and during EPO application with OGD exposure. In Figures 4F and 4G, transfection with Akt1 siRNA significantly prevented the expression of Akt1 during EPO administration alone and during OGD exposure. In addition, gene silencing of Akt1 eliminated the ability of EPO to phosphorylate PRAS40 without OGD exposure and during OGD exposure, illustrating that EPO also relies upon Akt1 to phosphorylate PRAS40 (Figures 4F and 4G). Non-specific scrambled siRNA did not alter PRAS40 phosphorylation illustrating the specificity for Akt1 in relation to PRAS40 phosphorylation.

EPO promotes PRAS40 binding to 14-3-3 protein that is PI 3-K dependent

The PI 3-K/Akt pathway can phosphorylate threonine246 on PRAS40 and result in the dissociation of PRAS40 from the mTOR complex mTORC1 [38]. This ultimately leads to the binding of phosphorylated PRAS40 to the docking protein 14-3-3 to inhibit PRAS40 and activate mTOR signaling [63], [64]. We therefore examined whether EPO alone or during OGD exposure altered the binding of PRAS40 to protein 14-3-3 by immunoprecipitation. EPO (10 ng/ml) significantly increased the expression of p-PRAS40 in the lysate that was immunoprecipitated by antibody against 14-3-3 protein, but decreased the expression of PRAS40 in the lysate that was immunoprecipitated by anti-mTOR the antibody (Figure 5A), suggesting that EPO dissociates PRAS40 from mTOR and increases the binding of p-PRAS40 to protein 14-3-3. Application of EPO with the PI 3-K inhibitors wortmannin (500 nM) or LY294002 (20 μM) reduced the expression of p-PRAS40 in the lysate that was immunoprecipitated by antibody against 14-3-3 protein and increased the expression of PRAS40 in the lysate that was immunoprecipitated by anti-mTOR antibody (Figure 5A), illustrating that EPO relies upon the PI 3-K pathway to foster PRAS40 binding to 14-3-3. Inhibition of the mTOR pathway with application of rapamycin (50 nM) did not alter the ability of EPO to promote PRAS40 binding to protein 14-3-3, further supporting the role of the PI 3-K pathway that is responsible for phosphorylation of PRAS40 (Figures 4D and 4E). In Figure 5B, EPO (10 ng/ml) during OGD exposure also significantly increased the binding of p-PRAS40 to 14-3-3 protein. Furthermore, EPO treatment during OGD with the PI 3-K inhibitors wortmannin (500 nM) or LY294002 (20 μM) limited PRAS40 binding to 14-3-3 protein (Figure 5B), demonstrating that EPO during OGD exposure also relies upon the PI 3-K pathway to promote PRAS40 binding to 14-3-3.

thumbnail
Figure 5. EPO promotes the binding of PRAS40 to protein 14-3-3 through PI 3-K.

(A) EPO (10 ng/ml) or EPO combined with wortmannin (500 nM), LY294002 (20 μM) or rapamycin (50 nM) was applied to SH-SY5Y cells and cell extracts 3 hours later were immunoprecipitated by antibodies against protein 14-3-3 or mTOR. Western blot analysis was performed to detect the expression phosphorylated (p-) p-PRAS40 and total PRAS40 in the precipitates (*P<0.01 vs. untreated control; P<0.01 vs. EPO treated alone). Application of EPO increased the expression of p-PRAS40 in the precipitate. In contrast, wortmannin or LY294002 significantly reduced expression of p-PRAS40 in the precipitate. (B) EPO (10 ng/ml) or EPO combined with wortmannin (500 nM) or LY294002 (20 μM) was applied to SH-SY5Y cells 1 hour prior to OGD and cell extracts 3 hours following OGD were immunoprecipitated by antibodies against protein 14-3-3 or mTOR. Western blot analysis was performed to detect the expression phosphorylated (p-) p-PRAS40 and total PRAS40 in the precipitates (*P<0.01 vs. OGD; P<0.01 vs. EPO/OGD). Application of EPO increased the expression of p-PRAS40 in the precipitate during OGD. In contrast, wortmannin or LY294002 significantly reduced expression of p-PRAS40 in the precipitate during OGD. In all cases above, each data point represents the mean and SD from 3 experiments.

https://doi.org/10.1371/journal.pone.0045456.g005

EPO and gene suppression of PRAS40 limit caspase 3 activation during OGD

Since EPO can modulate apoptotic cell injury and DNA degradation by limiting caspase activation [17], [21], [30], [52], [65], [66], we assessed whether PRAS40 played a role in this mechanism of cytoprotection by EPO. The expression of cleaved (active) caspase 3 on western analysis was assessed at 6 hours following OGD exposure and demonstrates significant caspase 3 activity (Figures 6A and 6B). Treatment with EPO (10 ng/ml) prevented caspase 3 activation following OGD exposure. In addition, transfection with PRAS40 siRNA significantly decreased caspase 3 activation after OGD exposure and further prevented caspase 3 activation during EPO treatment following OGD exposure (Figures 6A and 6B), suggesting that EPO relies in part on the inhibition of PRAS40 to prevent caspase 3 activation during OGD exposure. Non-specific scrambled siRNA did not alter caspase 3 activity during OGD exposure.

thumbnail
Figure 6. EPO and gene silencing of PRAS40 reduce caspase-3 activation during OGD.

(A) EPO (10 ng/ml) was applied 1 hour prior to OGD in SH-SY5Y cells. Western blot analysis was performed for the cleaved fragments of caspase 3 with a antibody that identifies both the 19 kDa and 17 kDa fragments of caspase 3 hours following OGD exposure. PRAS40 siRNA transfection significantly reduced the expression of caspase 3 cleaved fragments during OGD exposure and during EPO administration with OGD exposure. (B) Quantitative results of the band density of the western blot analysis for caspase 3 fragments show that PRAS40 siRNA transfection significantly reduced the expression of caspase 3 cleaved fragments during OGD exposure and during EPO administration with OGD exposure (*P<0.01 vs. untreated control; P<0.01 vs. OGD treated alone). Scrambled siRNA transfection did not alter the expression of the caspase 3 fragments during OGD alone or during EPO with OGD. PRAS40 siRNA transfection did not significantly alter caspase 3 fragment expression in untreated cells when compared to cells receiving no treatments. Each data point represents the mean and SD from 3 experiments.

https://doi.org/10.1371/journal.pone.0045456.g006

EPO linked pathways of ERK 1/2 and STAT5 do not alter phosphorylation of PRAS40

Prior studies have demonstrated that EPO can activate extracellular signal related kinase (ERK 1/2) [29], [67], [68] and signal transducer and activator of transcription (STAT5) [29], [69] that may contribute to increased cellular survival during oxidative stress. We therefore investigated whether ERK 1/2 and STAT5 could modulate phosphorylation of PRAS40. Western blot for phosphorylated p-ERK 1/2 (Thr 202/204) and phosphorylated p-STAT5 (Tyr694) (activate forms of ERK ½ and STAT5) were performed at 1, 3, and 24 hors following OGD exposure. As shown in Figures 7A and 7C, the expression of p-ERK 1/2 and p-STAT5 was initially increased at 1 and 3 hours following OGD exposure. Treatment with EPO (10 ng/ml) further significantly increased the expression of p-ERK 1/2 and p-STAT5 during these time periods following OGD exposure, illustrating that EPO increases the activation of these pathways. In addition, the ability of EPO to increase phosphorylation of ERK 1/2 and STAT5 during OGD exposure was eliminated during application of the ERK inhibitor ERK-I (100 μM) [70] and the STAT5 inhibitor STAT5-I (100 μM) [71] (Figures 7B and 7D). However, application of the ERK inhibitor ERK-I (100 μM) and the STAT5 inhibitor STAT5-I (100 μM) during EPO administration and OGD exposure did not alter the ability of EPO to phosphorylate PRAS40, illustrating that EPO does not rely upon ERK 1/2 or STAT5 to modulate the phosphorylation and activity of PRAS40 (Figures 7B and 7D).

thumbnail
Figure 7. EPO pathways of ERK 1/2 and STAT5 do not alter phosphorylation of PRAS40.

(A) Western blot analysis for phosphorylated (p-) ERK 1/2 (p-ERK 1/2, Thr202/Tyr204) in SH-SY5Y cells was performed at 1, 3, or 24 hours (hr) following OGD. EPO (10 ng/ml) that was applied to cell cultures 1 hour prior to OGD significantly increased p-ERK 1/2 expression at 1 and 3 hours following OGD (*P<0.01 vs. Control; P<0.01 vs. OGD at corresponding time points). (B) EPO (10 ng/ml) that was applied to SH-SY5Y cells significantly increased the expression of phosphorylated (p-) p-ERK 1/2 either alone or during OGD exposure 3 hours later. Expression of p-ERK 1/2 was significantly limited during the application of the ERK inhibitor (ERK-I, 100 μM) applied to cultures 30 min prior to EPO administration. Inhibition of ERK 1/2 did not alter the ability of EPO to significantly phosphorylate (p-) p-PRAS40 with or without OGD exposure (*P<0.01 vs. EPO; P<0.01 vs. EPO/OGD). (C) Western blot analysis for phosphorylated (p-) STAT5 (p-STAT5, Tyr694) in SH-SY5Y cells was performed at 1, 3, or 24 hours (hr) following OGD. EPO (10 ng/ml) that was applied to cell cultures 1 hour prior to OGD significantly increased p-STAT5 expression at 1 and 3 hours following OGD (*P<0.01 vs. Control; P<0.01 vs. OGD at corresponding time points). (D) EPO (10 ng/ml) that was applied to SH-SY5Y cells significantly increased the expression of phosphorylated (p-) p-STAT5 either alone or during OGD exposure 3 hours later. Expression of p-STAT5 was significantly limited during the application of the STAT5 inhibitor (STAT5-I, 100 μM) applied to cultures 30 min prior to EPO administration. Inhibition of STAT5-I, 100 μM did not alter the ability of EPO to significantly phosphorylate (p-) p-PRAS40 with or without OGD exposure (*P<0.01 vs. EPO; P<0.01 vs. EPO/OGD). In all cases above, each data point represents the mean and SD from 3 experiments.

https://doi.org/10.1371/journal.pone.0045456.g007

Discussion

Neurodegenerative disease can lead to disability in multiple systems of the body [72], [73], [74]. In addition, the release of reactive oxygen species during oxidative stress can significantly impact the onset and course of neurodegenerative disorders to influence the outcome of cerebral ischemia [49], [75], [76], neurodevelopment [77], [78], inflammation [79], [80], [81], [82], [83], and cognitive disorders [33], [84], [85]. Novel therapeutic strategies such as EPO may offer great promise to develop new treatments for disease of the nervous system [14], [15], [26], [27], [52], [86], [87]. However, EPO is not without detrimental effects such as during hypertension [88], [89], vascular disease [90], [91], [92], and cancer progression [93], [94], [95], [96]. Therefore it is crucial to identify and target the cellular mechanisms of EPO that can provide robust cytoprotection without detrimental consequences.

We show that treatment with EPO in SH-SY5Y cells prevents cellular injury and apoptotic DNA degradation during exposure to OGD. This cellular protection by EPO is dependent upon the activity of mTOR and its signaling pathways. Prior studies have shown that oxidative stress can block the activity of mTOR signaling pathways to alter cell metabolism and longevity [10], [41], [97] as well as lead to cell death [27], [36], [98]. In contrast, activation of mTOR during oxidative stress can result in cytoprotection [27], [37], [99]. We show that administration of the mTOR inhibitor rapamycin or gene silencing of mTOR during EPO application significantly prevented cellular protection by EPO, illustrating that EPO relies upon the activation of mTOR to protect neurons against oxidative stress.

Consistent with prior studies that have demonstrated that EPO increases mTOR signaling in inflammatory cells [27], [33] and in osteoblastic phenotypes in human bone marrow stromal cells [34], we illustrate that EPO in neuronal cells requires mTOR to phosphorylate p70S6K and 4EBP1. Phosphorylation of p70S6K activates this protein and can result in mRNA biogenesis, translation of ribosomal proteins, and cellular proliferation [53], [100]. In regards to 4EBP1, hypophosphorylation of 4EBP1 can block protein translation by allowing 4EBP1 to bind to the eukaryotic translation initiation factor 4 epsilon (eIF4E) through the eukaryotic translation initiation factor 4 gamma (eIF4G), a protein that transfers mRNA to ribosomes. Phosphorylation of 4EBP1 leads to the dissociation of 4EBP1 from eIF4E to allow eIF4G to begin mRNA translation [41], [59]. Prior studies have shown that mTOR depends upon the modulation of p70S6K and 4EBP1 to prevent cell death during apoptosis. Loss of mTOR signaling prevents phosphorylation of both p70S6K and 4EBP1 to lead to apoptosis [101]. In non-neuronal cells of the nervous system, activation of p70S6K by mTOR in astrocytes is cytoprotective through expression of Bcl-2/Bcl-xL expression to block BAD activity that can result in apoptosis [102]. Without significant mTOR activity, 4EBP1 binds to eIF4E that leads to the translation of apoptotic promoting proteins [103].

We also show that phosphorylation of mTOR, p70S6K, and 4EBP1 by EPO are dependent upon the PI 3-K/Akt pathways. Our studies demonstrate that EPO cannot directly phosphorylate mTOR, but requires the PI 3-K/Akt pathways to activate mTOR and phosphorylate p70S6K, and 4EBP1. Prior work has shown that cellular growth and protection can require the involvement of mTOR and the PI 3-K/Akt pathways [60]. The PI 3-K/Akt pathways are principal mediators for cell survival [104], [105], [106], [107], cellular metabolism [108], [109], [110], [111], and tumor progression [112], [113], [114]. EPO also utilizes the PI 3-K/Akt pathway to advance cellular survival in multiple systems of the body [15], [21], [22], [32], [68], [115], [116]. In addition, prior studies suggest that mTOR requires activation of the PI 3-K/Akt pathway to block apoptotic cell death [27], [33], [117], [118]. mTOR also in conjunction with the PI 3-K/Akt pathways can inactivate “pro-apoptotic” forkhead transcription factors to block cellular injury [97], [119], [120].

Given that mTOR signaling through p70S6K and 4EBP1 forms an important component for EPO neuronal protection during oxidative stress, we examined the role of PRAS40 that blocks mTOR activity and prevents the binding of p70S6K and 4EBP1 to Raptor [38], [39], [40]. We show that gene silencing of PRAS40 prevents both cellular injury and neuronal apoptosis during oxidative stress. Furthermore, gene silencing of PRAS40 during EPO administration and OGD exposure enhanced cell survival and further reduced genomic DNA degradation to a greater extent than EPO alone, suggesting that EPO relies upon PRAS40 inhibition for cellular protection. This increased degree of protection by EPO during the loss of PRAS40 appears to be tied to the promotion of mTOR signaling for EPO, since gene silencing of PRAS40 increased the phosphorylation of p-mTOR, p-p70S6K, and p-4EBP1 during EPO treatment. EPO also maintained the phosphorylation and inhibition of PRAS40 over a 24 hour course following OGD exposure to a significantly greater extent that during exposure to OGD alone. In other cell systems, phosphorylation of PRAS40 has resulted in decreased apoptotic cell death [42], [43], [44]. Loss of PRAS40 through gene silencing in HeLa cells also has been shown to prevent apoptosis against tumor necrosis factor and cyclohexamide [121].

EPO controls PRAS40 activity through the post-translational phosphorylation of PRAS40 and the subcellular binding of PRAS40 to protein 14-3-3. However, EPO does not appear to directly control the post-translational phosphorylation of PRAS40, but requires activation of the PI 3-K/Akt pathway similar to the regulation of mTOR by EPO. We show that in vitro incubation of EPO with PRAS40 does not lead to phosphorylation of PRAS40. In contrast, co-administration of the PI 3-K inhibitors wortmannin or LY294002 or gene silencing of Akt1 eliminated the ability of EPO to phosphorylate PRAS40, suggesting that EPO was dependent upon the PI 3-K/Akt pathways to phosphorylate PRAS40. Akt can phosphorylate threonine246 on PRAS40. This results in the dissociation of PRAS40 from mTORC1 [38] and the binding of PRAS40 to protein 14-3-3 to allow activation of mTOR signaling [63], [64]. Consistent with these studies for PRAS40, we demonstrate that EPO also fosters the binding of phosphorylated PRAS40 to protein 14-3-3. Application of EPO with the PI 3-K inhibitors wortmannin or LY294002 significantly prevented phosphorylated PRAS40 binding to protein 14-3-3, illustrating that EPO also employed the PI 3-K pathway to compartmentalize PRAS40 in the cell with protein 14-3-3. Yet, treatment with rapamycin did not affect phosphorylated PRAS40 binding to protein 14-3-3 during EPO administration, further demonstrating that control of phosphorylated PRAS40 binding to protein 14-3-3 by was modulated at the level of the PI 3-K pathway.

Control of neuronal apoptosis by EPO involves suppression of caspase 3 activation through PRAS40 but not through ERK 1/2 or STAT5 pathways. Both the early and late phases of apoptotic cell injury can be the result of caspase 3 activation [4], [113], [122], [123], [124], [125] and EPO has been shown to effectively control caspase activity [15], [21], [30], [50], [66], [126], [127]. We show that loss of PRAS40 during transfection with PRAS40 siRNA significantly limits caspase 3 activation after OGD exposure. In addition, caspase 3 activation following OGD exposure is decreased to a greater degree by EPO during gene silencing of PRAS40 than with EPO alone, suggesting that PRAS40 controls caspase activation and mediates the ability of EPO to block caspase 3 activity. EPO also uses the pathways of ERK 1/2 and STAT5 to prevent apoptotic cell injury [29], [67], [68]. Yet, we found that phosphorylation of PRAS40 by EPO was not associated with ERK 1/2 and STAT5, suggesting that ERK 1/2, STAT5, and PRAS40 represent independent cytoprotective pathways for EPO.

Our studies offer new insight into the neuroprotective pathways for EPO during OGD that can lead to oxidative stress. EPO prevents cell injury through activation of mTOR signaling and is dependent upon the inhibition of PRAS40 as vital regulatory pathway. EPO controls the mTOR signaling pathways of p70S6K, 4EBP1, and PRAS40 through PI 3-K and Akt. PRAS40 either alone or as a component of EPO signal transduction governs cell survival through caspase 3 activation that is independent of other cytoprotective pathways of EPO that involve ERK 1/2 and STAT5. Future studies that can further elucidate the role of PRAS40 and the ability of EPO to regulate this pathway may offer novel approaches for the treatment of a variety of multi-system disorders, such as those that involve neurodegeneration.

Author Contributions

Conceived and designed the experiments: KM ZC. Performed the experiments: KM ZC YS SW. Analyzed the data: KM ZC. Contributed reagents/materials/analysis tools: KM. Wrote the paper: KM ZC.

References

  1. 1. Bajda M, Guzior N, Ignasik M, Malawska B (2011) Multi-target-directed ligands in Alzheimer's disease treatment. Curr Med Chem 18: 4949–4975.
  2. 2. Enz R (2012) Metabotropic glutamate receptors and interacting proteins: evolving drug targets. Curr Drug Targets 13: 145–156.
  3. 3. Maiese K, Chong ZZ, Hou J, Shang YC (2009) New strategies for Alzheimer's disease and cognitive impairment. Oxid Med Cell Longev 2: 279–289.
  4. 4. Chong ZZ, Li F, Maiese K (2005) Oxidative stress in the brain: Novel cellular targets that govern survival during neurodegenerative disease. Prog Neurobiol 75: 207–246.
  5. 5. Srivastava S, Haigis MC (2011) Role of sirtuins and calorie restriction in neuroprotection: implications in Alzheimer's and Parkinson's diseases. Curr Pharm Des 17: 3418–3433.
  6. 6. Zhang G, Zhao Z, Gao L, Deng J, Wang B, et al. (2011) Gypenoside attenuates white matter lesions induced by chronic cerebral hypoperfusion in rats. Pharmacol Biochem Behav 99: 42–51.
  7. 7. Chong ZZ, Shang YC, Wang S, Maiese K (2012) SIRT1: New avenues of discovery for disorders of oxidative stress. Expert Opin Ther Targets 16: 167–178.
  8. 8. Shacka JJ, Roth KA, Zhang J (2008) The autophagy-lysosomal degradation pathway: role in neurodegenerative disease and therapy. Front Biosci 13: 718–736.
  9. 9. Kashihara N, Haruna Y, Kondeti VK, Kanwar YS (2010) Oxidative stress in diabetic nephropathy. Curr Med Chem 17: 4256–4269.
  10. 10. Maiese K, Chong ZZ, Shang YC, Hou J (2011) Novel Avenues of Drug Discovery and Biomarkers for Diabetes Mellitus. J Clin Pharmacol 51: 128–152.
  11. 11. Yang H, Jin X, Kei Lam CW, Yan SK (2011) Oxidative stress and diabetes mellitus. Clin Chem Lab Med 49: 1773–1782.
  12. 12. Kumral A, Tuzun F, Oner MG, Genc S, Duman N, et al. (2011) Erythropoietin in neonatal brain protection: the past, the present and the future. Brain Dev 33: 632–643.
  13. 13. Maiese K, Chong ZZ, Li F, Shang YC (2008) Erythropoietin: elucidating new cellular targets that broaden therapeutic strategies. Prog Neurobiol 85: 194–213.
  14. 14. Chong ZZ, Li F, Maiese K (2005) Erythropoietin requires NF-kappaB and its nuclear translocation to prevent early and late apoptotic neuronal injury during beta-amyloid toxicity. Curr Neurovasc Res 2: 387–399.
  15. 15. Chong ZZ, Lin SH, Kang JQ, Maiese K (2003) Erythropoietin prevents early and late neuronal demise through modulation of Akt1 and induction of caspase 1, 3, and 8. J Neurosci Res 71: 659–669.
  16. 16. Digicaylioglu M, Lipton SA (2001) Erythropoietin-mediated neuroprotection involves cross-talk between Jak2 and NF-kappaB signalling cascades. Nature 412: 641–647.
  17. 17. Park KH, Choi NY, Koh SH, Park HH, Kim YS, et al. (2011) L-DOPA neurotoxicity is prevented by neuroprotective effects of erythropoietin. Neurotoxicology 32: 879–887.
  18. 18. Ruscher K, Freyer D, Karsch M, Isaev N, Megow D, et al. (2002) Erythropoietin is a paracrine mediator of ischemic tolerance in the brain: evidence from an in vitro model. J Neurosci 22: 10291–10301.
  19. 19. Simon F, Scheuerle A, Groger M, Vcelar B, McCook O, et al. (2011) Comparison of carbamylated erythropoietin-FC fusion protein and recombinant human erythropoietin during porcine aortic balloon occlusion-induced spinal cord ischemia/reperfusion injury. Intensive Care Med 37: 1525–1533.
  20. 20. Velly L, Pellegrini L, Guillet B, Bruder N, Pisano P (2010) Erythropoietin 2nd cerebral protection after acute injuries: a double-edged sword? Pharmacol Ther 128: 445–459.
  21. 21. Chong ZZ, Hou J, Shang YC, Wang S, Maiese K (2011) 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 8: 103–120.
  22. 22. Chong ZZ, Kang JQ, Maiese K (2002) Erythropoietin is a novel vascular protectant through activation of Akt1 and mitochondrial modulation of cysteine proteases. Circulation 106: 2973–2979.
  23. 23. Xu Y, Tian Y, Wei HJ, Chen J, Dong JF, et al. (2011) Erythropoietin increases circulating endothelial progenitor cells and reduces the formation and progression of cerebral aneurysm in rats. Neuroscience 181: 292–299.
  24. 24. Yamada M, Burke C, Colditz P, Johnson DW, Gobe GC (2011) Erythropoietin protects against apoptosis and increases expression of non-neuronal cell markers in the hypoxia-injured developing brain. J Pathol 224: 101–109.
  25. 25. Hamed S, Egozi D, Kruchevsky D, Teot L, Gilhar A, et al. (2010) Erythropoietin improves the survival of fat tissue after its transplantation in nude mice. PLoS ONE 5: e13986.
  26. 26. Loeliger MM, Mackintosh A, De Matteo R, Harding R, Rees SM (2011) Erythropoietin protects the developing retina in an ovine model of endotoxin-induced retinal injury. Invest Ophthalmol Vis Sci 52: 2656–2661.
  27. 27. Shang YC, Chong ZZ, Wang S, Maiese K (2011) Erythropoietin and Wnt1 Govern Pathways of mTOR, Apaf-1, and XIAP in Inflammatory Microglia. Curr Neurovasc Res 8: 270–285.
  28. 28. Solling C (2011) Organ-Protective and Immunomodulatory Effects of Erythropoietin – An Update on Recent Clinical Trials. Basic Clin Pharmacol Toxicol.
  29. 29. Chong ZZ, Maiese K (2007) Erythropoietin involves the phosphatidylinositol 3-kinase pathway, 14-3-3 protein and FOXO3a nuclear trafficking to preserve endothelial cell integrity. Br J Pharmacol 150: 839–850.
  30. 30. Hou J, Wang S, Shang YC, Chong ZZ, Maiese K (2011) Erythropoietin Employs Cell Longevity Pathways of SIRT1 to Foster Endothelial Vascular Integrity During Oxidant Stress. Curr Neurovasc Res 8: 220–235.
  31. 31. Su KH, Shyue SK, Kou YR, Ching LC, Chiang AN, et al. (2011) beta Common receptor integrates the erythropoietin signaling in activation of endothelial nitric oxide synthase. J Cell Physiol 226: 3330–3339.
  32. 32. Takeyama T, Takemura G, Kanamori H, Kawaguchi T, Ogino A, et al. (2012) Asialoerythropoietin, a nonerythropoietic derivative of erythropoietin, displays broad anti-heart failure activity. Circ Heart Fail 5: 274–285.
  33. 33. Shang YC, Chong ZZ, Wang S, Maiese K (2012) 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) 4: 187–201.
  34. 34. Kim J, Jung Y, Sun H, Joseph J, Mishra A, et al. (2012) Erythropoietin mediated bone formation is regulated by mTOR signaling. J Cell Biochem 113: 220–228.
  35. 35. Choi KC, Kim SH, Ha JY, Kim ST, Son JH (2010) A novel mTOR activating protein protects dopamine neurons against oxidative stress by repressing autophagy related cell death. J Neurochem 112: 366–376.
  36. 36. Chen L, Xu B, Liu L, Luo Y, Yin J, et al. (2010) Hydrogen peroxide inhibits mTOR signaling by activation of AMPKalpha leading to apoptosis of neuronal cells. Lab Invest 90: 762–773.
  37. 37. Chong ZZ, Li F, Maiese K (2007) The pro-survival pathways of mTOR and protein kinase B target glycogen synthase kinase-3beta and nuclear factor-kappaB to foster endogenous microglial cell protection. Int J Mol Med 19: 263–272.
  38. 38. Sancak Y, Thoreen CC, Peterson TR, Lindquist RA, Kang SA, et al. (2007) PRAS40 is an insulin-regulated inhibitor of the mTORC1 protein kinase. Mol Cell 25: 903–915.
  39. 39. Wang H, Zhang Q, Wen Q, Zheng Y, Philip L, et al. (2012) Proline-rich Akt substrate of 40kDa (PRAS40): a novel downstream target of PI3k/Akt signaling pathway. Cell Signal 24: 17–24.
  40. 40. Wang L, Harris TE, Roth RA, Lawrence JC (2007) PRAS40 regulates mTORC1 kinase activity by functioning as a direct inhibitor of substrate binding. J Biol Chem 282: 20036–20044.
  41. 41. Chong ZZ, Maiese K (2012) Mammalian Target of Rapamycin Signaling in Diabetic Cardiovascular Disease. Cardiovasc Diabetol 11: 45.
  42. 42. Das F, Dey N, Venkatesan B, Kasinath BS, Ghosh-Choudhury N, et al. (2011) High glucose upregulation of early-onset Parkinson's disease protein DJ-1 integrates the PRAS40/TORC1 axis to mesangial cell hypertrophy. Cell Signal 23: 1311–1319.
  43. 43. Kapoor V, Zaharieva MM, Das SN, Berger MR (2012) Erufosine simultaneously induces apoptosis and autophagy by modulating the Akt-mTOR signaling pathway in oral squamous cell carcinoma. Cancer Lett 319: 39–48.
  44. 44. Wang J, Xiao X, Zhang Y, Shi D, Chen W, et al. (2012) Simultaneous modulation of COX-2, p300, Akt, and Apaf-1 signaling by melatonin to inhibit proliferation and induce apoptosis in breast cancer cells. J Pineal Res.
  45. 45. Furuichi T, Liu W, Shi H, Miyake M, Liu KJ (2005) Generation of hydrogen peroxide during brief oxygen-glucose deprivation induces preconditioning neuronal protection in primary cultured neurons. J Neurosci Res 79: 816–824.
  46. 46. Weih M, Schmitt M, Gieche J, Harms C, Ruscher K, et al. (2001) Proteolysis of oxidized proteins after oxygen-glucose deprivation in rat cortical neurons is mediated by the proteasome. J Cereb Blood Flow Metab 21: 1090–1096.
  47. 47. Balan V, Miller GS, Kaplun L, Balan K, Chong ZZ, et al. (2008) Life span extension and neuronal cell protection by Drosophila nicotinamidase. J Biol Chem 283: 27810–27819.
  48. 48. Chong ZZ, Li F, Maiese K (2007) Cellular demise and inflammatory microglial activation during beta-amyloid toxicity are governed by Wnt1 and canonical signaling pathways. Cell Signal 19: 1150–1162.
  49. 49. Chong ZZ, Shang YC, Hou J, Maiese K (2010) Wnt1 neuroprotection translates into improved neurological function during oxidant stress and cerebral ischemia through AKT1 and mitochondrial apoptotic pathways. Oxid Med Cell Longev 3: 153–165.
  50. 50. Chong ZZ, Kang JQ, Maiese K (2003) Apaf-1, Bcl-xL, Cytochrome c, and Caspase-9 Form the Critical Elements for Cerebral Vascular Protection by Erythropoietin. J Cereb Blood Flow Metab 23: 320–330.
  51. 51. Shang YC, Chong ZZ, Hou J, Maiese K (2009) The forkhead transcription factor FoxO3a controls microglial inflammatory activation and eventual apoptotic injury through caspase 3. Curr Neurovasc Res 6: 20–31.
  52. 52. Chong ZZ, Kang JQ, Maiese K (2003) Erythropoietin fosters both intrinsic and extrinsic neuronal protection through modulation of microglia, Akt1, Bad, and caspase-mediated pathways. Br J Pharmacol 138: 1107–1118.
  53. 53. Chong ZZ, Shang YC, Zhang L, Wang S, Maiese K (2010) Mammalian target of rapamycin: hitting the bull's-eye for neurological disorders. Oxid Med Cell Longev 3: 374–391.
  54. 54. Reynolds THt, Bodine SC, Lawrence JC Jr (2002) Control of Ser2448 phosphorylation in the mammalian target of rapamycin by insulin and skeletal muscle load. J Biol Chem 277: 17657–17662.
  55. 55. Chiang GG, Abraham RT (2005) Phosphorylation of mammalian target of rapamycin (mTOR) at Ser-2448 is mediated by p70S6 kinase. J Biol Chem 280: 25485–25490.
  56. 56. Pearson RB, Dennis PB, Han JW, Williamson NA, Kozma SC, et al. (1995) The principal target of rapamycin-induced p70s6k inactivation is a novel phosphorylation site within a conserved hydrophobic domain. Embo J 14: 5279–5287.
  57. 57. Magagnin MG, van den Beucken T, Sergeant K, Lambin P, Koritzinsky M, et al. (2008) The mTOR target 4E-BP1 contributes to differential protein expression during normoxia and hypoxia through changes in mRNA translation efficiency. Proteomics 8: 1019–1028.
  58. 58. Gingras AC, Kennedy SG, O'Leary MA, Sonenberg N, Hay N (1998) 4E-BP1, a repressor of mRNA translation, is phosphorylated and inactivated by the Akt(PKB) signaling pathway. Genes Dev 12: 502–513.
  59. 59. Bhandari BK, Feliers D, Duraisamy S, Stewart JL, Gingras AC, et al. (2001) Insulin regulation of protein translation repressor 4E-BP1, an eIF4E-binding protein, in renal epithelial cells. Kidney Int 59: 866–875.
  60. 60. Sato A, Sunayama J, Matsuda K, Tachibana K, Sakurada K, et al. (2010) Regulation of neural stem/progenitor cell maintenance by PI3K and mTOR. Neurosci Lett 470: 115–120.
  61. 61. Wymann MP, Bulgarelli-Leva G, Zvelebil MJ, Pirola L, Vanhaesebroeck B, et al. (1996) Wortmannin inactivates phosphoinositide 3-kinase by covalent modification of Lys-802, a residue involved in the phosphate transfer reaction. Mol Cell Biol 16: 1722–1733.
  62. 62. Vlahos CJ, Matter WF, Hui KY, Brown RF (1994) A specific inhibitor of phosphatidylinositol 3-kinase, 2-(4-morpholinyl)-8-phenyl-4H-1-benzopyran-4-one (LY294002). J Biol Chem 269: 5241–5248.
  63. 63. Kovacina KS, Park GY, Bae SS, Guzzetta AW, Schaefer E, et al. (2003) Identification of a proline-rich Akt substrate as a 14-3-3 binding partner. J Biol Chem 278: 10189–10194.
  64. 64. Vander Haar E, Lee SI, Bandhakavi S, Griffin TJ, Kim DH (2007) Insulin signalling to mTOR mediated by the Akt/PKB substrate PRAS40. Nat Cell Biol 9: 316–323.
  65. 65. Ammar HI, Saba S, Ammar RI, Elsayed LA, Ghaly WB, et al. (2011) Erythropoietin protects against doxorubicin-induced heart failure. Am J Physiol Heart Circ Physiol 301: H2413–2421.
  66. 66. Canadillas S, Ortega R, Estepa JC, Egea J, Gonzalez-Menchen A, et al. (2010) Darbepoetin-alpha treatment enhances glomerular regenerative process in the Thy-1 glomerulonephritis model. Am J Physiol Renal Physiol 299: F1278–1287.
  67. 67. Wang Y, Hu X, Xie X, He A, Liu X, et al. (2011) Effects of mesenchymal stem cells on matrix metalloproteinase synthesis in cardiac fibroblasts. Exp Biol Med (Maywood) 236: 1197–1204.
  68. 68. Xie Z, Chen F, Wu X, Zhuang C, Zhu J, et al. (2012) Effects of supplemental erythropoietin on its receptor expression and signal transduction pathways in rat model of retinal detachment. Curr Eye Res 37: 138–144.
  69. 69. Ji YQ, Zhang YQ, Li MQ, Du MR, Wei WW, et al. (2011) EPO improves the proliferation and inhibits apoptosis of trophoblast and decidual stromal cells through activating STAT-5 and inactivating p38 signal in human early pregnancy. Int J Clin Exp Pathol 4: 765–774.
  70. 70. Muller J, Sperl B, Reindl W, Kiessling A, Berg T (2008) Discovery of chromone-based inhibitors of the transcription factor STAT5. Chembiochem 9: 723–727.
  71. 71. Hancock CN, Macias A, Lee EK, Yu SY, Mackerell AD, et al. (2005) Identification of novel extracellular signal-regulated kinase docking domain inhibitors. J Med Chem 48: 4586–4595.
  72. 72. Caprara C, Grimm C (2012) From oxygen to erythropoietin: relevance of hypoxia for retinal development, health and disease. Prog Retin Eye Res 31: 89–119.
  73. 73. Holopainen IE, Lauren HB (2012) Glutamate signaling in the pathophysiology and therapy of prenatal insults. Pharmacol Biochem Behav 100: 825–834.
  74. 74. Maiese K, Chong ZZ, Shang YC, Wang S (2011) Translating cell survival and cell longevity into treatment strategies with SIRT1. Rom J Morphol Embryol 52: 1173–1185.
  75. 75. Singh N, Sharma G, Mishra V, Raghubir R (2012) Hypoxia Inducible Factor-1: Its Potential Role In Cerebral Ischemia. Cell Mol Neurobiol.
  76. 76. Uzbay TI (2012) The pharmacological importance of agmatine in the brain. Neurosci Biobehav Rev 36: 502–519.
  77. 77. Ullah N, Lee HY, Naseer MI, Ullah I, Suh JW, et al. (2011) Nicotinamide inhibits alkylating agent-induced apoptotic neurodegeneration in the developing rat brain. PLoS ONE 6: e27093.
  78. 78. Ullah N, Ullah I, Lee HY, Naseer MI, Seok PM, et al. (2012) Protective Function of Nicotinamide Against Ketamine-induced Apoptotic Neurodegeneration in the Infant Rat Brain. J Mol Neurosci 47: 67–75.
  79. 79. Kato S, Aoyama M, Kakita H, Hida H, Kato I, et al. (2011) Endogenous erythropoietin from astrocyte protects the oligodendrocyte precursor cell against hypoxic and reoxygenation injury. J Neurosci Res 89: 1566–1574.
  80. 80. Kigerl KA, Ankeny DP, Garg SK, Wei P, Guan Z, et al. (2012) System x(c)(-) regulates microglia and macrophage glutamate excitotoxicity in vivo. Exp Neurol 233: 333–341.
  81. 81. L'Episcopo F, Tirolo C, Testa N, Caniglia S, Morale MC, et al. (2012) 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 32: 2062–2085.
  82. 82. Shang YC, Chong ZZ, Hou J, Maiese K (2009) FoxO3a governs early microglial proliferation and employs mitochondrial depolarization with caspase 3, 8, and 9 cleavage during oxidant induced apoptosis. Curr Neurovasc Res 6: 223–238.
  83. 83. Shang YC, Chong ZZ, Hou J, Maiese K (2010) Wnt1, FoxO3a, and NF-kappaB oversee microglial integrity and activation during oxidant stress. Cell Signal 22: 1317–1329.
  84. 84. Hong YK, Lee S, Park SH, Lee JH, Han SY, et al. (2012) Inhibition of JNK/dFOXO pathway and caspases rescues neurological impairments in Drosophila Alzheimer's disease model. Biochem Biophys Res Commun 419: 49–53.
  85. 85. Liu T, Jin H, Sun QR, Xu JH, Hu HT (2010) The neuroprotective effects of tanshinone IIA on beta-amyloid-induced toxicity in rat cortical neurons. Neuropharmacology 59: 595–604.
  86. 86. Kollensperger M, Krismer F, Pallua A, Stefanova N, Poewe W, et al. (2011) Erythropoietin is neuroprotective in a transgenic mouse model of multiple system atrophy. Mov Disord 26: 507–515.
  87. 87. Kook YH, Ka M, Um M (2011) Neuroprotective cytokines repress PUMA induction in the 1-methyl-4-phenylpyridinium (MPP(+)) model of Parkinson's disease. Biochem Biophys Res Commun 411: 370–374.
  88. 88. Ioka T, Tsuruoka S, Ito C, Iwaguro H, Asahara T, et al. (2009) Hypertension induced by erythropoietin has a correlation with truncated erythropoietin receptor mRNA in endothelial progenitor cells of hemodialysis patients. Clin Pharmacol Ther 86: 154–159.
  89. 89. Maiese K, Li F, Chong ZZ (2005) New avenues of exploration for erythropoietin. Jama 293: 90–95.
  90. 90. Lanfranconi S, Locatelli F, Corti S, Candelise L, Comi GP, et al. (2011) Growth factors in ischemic stroke. J Cell Mol Med 15: 1645–1687.
  91. 91. Maiese K, Chong ZZ, Shang YC (2008) Raves and risks for erythropoietin. Cytokine Growth Factor Rev 19: 145–155.
  92. 92. Mastromarino V, Volpe M, Musumeci MB, Autore C, Conti E (2011) Erythropoietin and the heart: facts and perspectives. Clin Sci (Lond) 120: 51–63.
  93. 93. Hardee ME, Cao Y, Fu P, Jiang X, Zhao Y, et al. (2007) Erythropoietin blockade inhibits the induction of tumor angiogenesis and progression. PLoS ONE 2: e549.
  94. 94. Hedley BD, Allan AL, Xenocostas A (2011) The role of erythropoietin and erythropoiesis-stimulating agents in tumor progression. Clin Cancer Res 17: 6373–6380.
  95. 95. Lombardero M, Kovacs K, Scheithauer BW (2011) Erythropoietin: a hormone with multiple functions. Pathobiology 78: 41–53.
  96. 96. Maiese K, Li F, Chong ZZ (2005) Erythropoietin and cancer. JAMA 293: 1858–1859.
  97. 97. Wang RH, Kim HS, Xiao C, Xu X, Gavrilova O, et al. (2011) Hepatic Sirt1 deficiency in mice impairs mTorc2/Akt signaling and results in hyperglycemia, oxidative damage, and insulin resistance. J Clin Invest 121: 4477–4490.
  98. 98. Andreucci M, Fuiano G, Presta P, Lucisano G, Leone F, et al. (2009) Downregulation of cell survival signalling pathways and increased cell damage in hydrogen peroxide-treated human renal proximal tubular cells by alpha-erythropoietin. Cell Prolif 42: 554–561.
  99. 99. Di Nardo A, Kramvis I, Cho N, Sadowski A, Meikle L, et al. (2009) Tuberous sclerosis complex activity is required to control neuronal stress responses in an mTOR-dependent manner. J Neurosci 29: 5926–5937.
  100. 100. Jastrzebski K, Hannan KM, Tchoubrieva EB, Hannan RD, Pearson RB (2007) Coordinate regulation of ribosome biogenesis and function by the ribosomal protein S6 kinase, a key mediator of mTOR function. Growth Factors 25: 209–226.
  101. 101. Hou G, Xue L, Lu Z, Fan T, Tian F, et al. (2007) 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 253: 236–248.
  102. 102. Pastor MD, Garcia-Yebenes I, Fradejas N, Perez-Ortiz JM, Mora-Lee S, et al. (2009) mTOR/S6 kinase pathway contributes to astrocyte survival during ischemia. J Biol Chem 284: 22067–22078.
  103. 103. Zhang D, Contu R, Latronico MV, Zhang J, Rizzi R, et al. (2010) MTORC1 regulates cardiac function and myocyte survival through 4E-BP1 inhibition in mice. J Clin Invest 120: 2805–2816.
  104. 104. Chong ZZ, Maiese K (2007) The Src homology 2 domain tyrosine phosphatases SHP-1 and SHP-2: diversified control of cell growth, inflammation, and injury. Histol Histopathol 22: 1251–1267.
  105. 105. Fokas E, Yoshimura M, Prevo R, Higgins G, Hackl W, et al. (2012) NVP-BEZ235 and NVP-BGT226, dual phosphatidylinositol 3-kinase/Mammalian target of rapamycin inhibitors, enhance tumor and endothelial cell radiosensitivity. Radiat Oncol 7: 48.
  106. 106. Hou J, Chong ZZ, Shang YC, Maiese K (2010) Early apoptotic vascular signaling is determined by Sirt1 through nuclear shuttling, forkhead trafficking, bad, and mitochondrial caspase activation. Curr Neurovasc Res 7: 95–112.
  107. 107. Kang JQ, Chong ZZ, Maiese K (2003) Critical role for Akt1 in the modulation of apoptotic phosphatidylserine exposure and microglial activation. Mol Pharmacol 64: 557–569.
  108. 108. Chen JX, Tuo Q, Liao DF, Zeng H (2012) Inhibition of Protein Tyrosine Phosphatase Improves Angiogenesis via Enhancing Ang-1/Tie-2 Signaling in Diabetes. Exp Diabetes Res 2012: 836759.
  109. 109. Deblon N, Bourgoin L, Veyrat-Durebex C, Peyrou M, Vinciguerra M, et al. (2012) Chronic mTOR inhibition by rapamycin induces muscle insulin resistance despite weight loss in rats. Br J Pharmacol 165: 2325–2340.
  110. 110. Hou J, Chong ZZ, Shang YC, Maiese K (2010) FoxO3a governs early and late apoptotic endothelial programs during elevated glucose through mitochondrial and caspase signaling. Mol Cell Endocrinol 321: 194–206.
  111. 111. Maiese K, Chong ZZ, Hou J, Shang YC (2009) The vitamin nicotinamide: translating nutrition into clinical care. Molecules 14: 3446–3485.
  112. 112. Chong ZZ, Li F, Maiese K (2005) Activating Akt and the brain's resources to drive cellular survival and prevent inflammatory injury. Histol Histopathol 20: 299–315.
  113. 113. Chung CY, Park YL, Song YA, Myung E, Kim KY, et al. (2012) Knockdown of RON Inhibits AP-1 Activity and Induces Apoptosis and Cell Cycle Arrest Through the Modulation of Akt/FoxO Signaling in Human Colorectal Cancer Cells. Dig Dis Sci 57: 371–380.
  114. 114. Janku F, Wheler JJ, Westin SN, Moulder SL, Naing A, et al. (2012) PI3K/AKT/mTOR inhibitors in patients with breast and gynecologic malignancies harboring PIK3CA mutations. J Clin Oncol 30: 777–782.
  115. 115. Eliopoulos N, Zhao J, Forner K, Birman E, Young YK, et al. (2011) Erythropoietin Gene-enhanced Marrow Mesenchymal Stromal Cells Decrease Cisplatin-induced Kidney Injury and Improve Survival of Allogeneic Mice. Mol Ther 19: 2072–2083.
  116. 116. Shen J, Wu Y, Xu JY, Zhang J, Sinclair SH, et al. (2010) ERK- and Akt-dependent neuroprotection by erythropoietin (EPO) against glyoxal-AGEs via modulation of Bcl-xL, Bax, and BAD. Invest Ophthalmol Vis Sci 51: 35–46.
  117. 117. Hernandez G, Lal H, Fidalgo M, Guerrero A, Zalvide J, et al. (2011) A novel cardioprotective p38-MAPK/mTOR pathway. Exp Cell Res 317: 2938–2949.
  118. 118. Magri L, Cambiaghi M, Cominelli M, Alfaro-Cervello C, Cursi M, et al. (2011) Sustained activation of mTOR pathway in embryonic neural stem cells leads to development of tuberous sclerosis complex-associated lesions. Cell Stem Cell 9: 447–462.
  119. 119. Dormond O, Madsen JC, Briscoe DM (2007) The effects of mTOR-Akt interactions on anti-apoptotic signaling in vascular endothelial cells. J Biol Chem 282: 23679–23686.
  120. 120. Maiese K, Chong ZZ, Shang YC (2008) OutFOXOing disease and disability: the therapeutic potential of targeting FoxO proteins. Trends Mol Med 14: 219–227.
  121. 121. Thedieck K, Polak P, Kim ML, Molle KD, Cohen A, et al. (2007) PRAS40 and PRR5-like protein are new mTOR interactors that regulate apoptosis. PLoS One 2: e1217.
  122. 122. Maiese K, Chong ZZ, Hou J, Shang YC (2010) Oxidative stress: Biomarkers and novel therapeutic pathways. Exp Gerontol 45: 217–234.
  123. 123. Sung JH, Kim MO, Koh PO (2012) Nicotinamide Prevents the Down-Regulation of MEK/ERK/p90RSK Signaling Cascade in Brain Ischemic Injury. J Vet Med Sci 74: 35–41.
  124. 124. Troy CM, Akpan N, Jean YY (2011) Regulation of caspases in the nervous system implications for functions in health and disease. Prog Mol Biol Transl Sci 99: 265–305.
  125. 125. Wang J, Sun P, Bao Y, Dou B, Song D, et al. (2012) Vitamin E renders protection to PC12 cells against oxidative damage and apoptosis induced by single-walled carbon nanotubes. Toxicol In Vitro 26: 32–41.
  126. 126. Dang J, Jia R, Tu Y, Xiao S, Ding G (2010) Erythropoietin prevents reactive oxygen species generation and renal tubular cell apoptosis at high glucose level. Biomed Pharmacother 64: 681–685.
  127. 127. Wu Y, Shang Y, Sun S, Liang H, Liu R (2007) Erythropoietin prevents PC12 cells from 1-methyl-4-phenylpyridinium ion-induced apoptosis via the Akt/GSK-3beta/caspase-3 mediated signaling pathway. Apoptosis 12: 1365–1375.