Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2008 Oct 21.
Published in final edited form as: Curr Neurovasc Res. 2007 Nov;4(4):295–302. doi: 10.2174/156720207782446306

“SLY AS A FOXO”: New Paths with Forkhead Signaling in the Brain

Kenneth Maiese 1,2,*, Zhao Zhong Chong 1, Yan Chen Shang 1
PMCID: PMC2570708  NIHMSID: NIHMS73606  PMID: 18045156

Abstract

The Forkhead transcription factor FOXO3a has emerged as a versatile target for diseases that impact upon neuronal survival, vascular integrity, immune function, and cellular metabolism. Enthusiasm is high to fill a critical treatment void through FOXO3a signaling for several neurodegenerative disorders that include aging, neuromuscular disease, systemic lupus erythematosus, stroke, and diabetic complications. Here we discuss the influence of FOXO3a upon cell survival and longevity, the intricate signal transduction pathways of FOXO3a, insights into present disease models, and the potential clinical translation of FOXO3a signaling into novel therapeutic strategies.

Keywords: Amyotrophic lateral sclerosis, diabetes, apoptosis, FOXO3a, FKHRL1, Akt, erythropoietin, neuromuscular disease, oxidative stress, psychiatric, systemic lupus erythematosus, stroke, stem cells

INTRODUCTION

The mammalian forkhead transcription factor family regulates processes ranging from cell longevity to cell apoptosis and function as transcription factors by preferentially binding to the core consensus DNA sequence 5′-TTGTTTAG-3′, the Forkhead response element (Chong et al., 2005e, Chong et al., 2004c, Wijchers et al., 2006). The first member of this family was the Drosophila melanogaster gene Fork head. Since this time, greater than 100 forkhead genes and 19 human subgroups are known to exist that extend from FOXA to FOXS (Wijchers et al., 2006). The forkhead box (FOX) family of genes is characterized by a conserved forkhead domain commonly noted as a “forkhead box” or a “winged helix” as a result of the butterfly-like appearance on X-ray crystallography (Clark et al., 1993) and nuclear magnetic resonance (Jin et al., 1998). All Fox proteins contain the 100-amino acid winged helix domain, but it should be noted that not all winged helix domains are Fox proteins (Larson et al., 2007). The forkhead domain may have a number of variations from the usual three α-helices, three β-sheets, and two loops that are referred to as the wings. This domain defines the Fox class of transcription factors, but additional portions of the Fox proteins can be divergent. A unified nomenclature has been described that has now replaced prior terms such as forkhead in rhabdomyosarcoma (FKHR), the Drosophila gene fork head (fkh), and Forkhead RElated ACtivator (FREAC)-1 and -2 (Kaestner et al., 2000). Within the subclasses of the Fox proteins that are each designated by a letter, an Arabic number is provided such that the actual name of a Fox protein would follow the direction of “Fox, subclass or subgroup “Letter”, and member “Number”. In relation to human Fox proteins, all letters are capitalized, otherwise only the initial letter is listed as uppercase.

FOXO3a EXPRESSION

Of the Forkhead transcription factors, FOXO3a is one member that has emerged as a versatile target for a number of disorders. In tissues such as alveolar rhabdomyosarcoma tumor samples, the FOXO3a gene consists of at least three exons and leads to a protein containing 673 amino acids. The initial exon contains the N-terminal portion of the protein including the initiating ATG and 60 amino acids of the forkhead domain. This portion is GC rich and contains the methylation-sensitive restriction sites. The second exon contains the C-terminal 50 amino acids of the forkhead domain and the remaining translated region (Anderson et al., 1998).

FOXO3a, also known as FKHRL1 (forkhead in rhabdomyosarcoma like 1), is expressed in a wide variety of human cell types. For example, FOXO3a protein is present in human megakaryocytes (Tanaka et al., 2001) and in the in the lamina propria and muscular of the gastro intestine of the pig (Zhou et al., 2007). Furthermore, FOXO3a transcripts are expressed in multiple tissue that involve the ovary, prostate, skeletal muscle, brain, heart, lung, liver, pancreas, spleen, thymus, and testis. In murine models, mRNA expression is highest in muscle, adipose tissue, and embryonic liver (Furuyama et al., 2000). With in situ hybridization experiments on adult and embryonic mouse brains, Foxo3 is expressed throughout the brain including all hippocampal areas, cortex and cerebellum (Hoekman et al., 2006).

FOXO3a IN VIVO DELETION

Interestingly, Foxo1−/− mouse embryos usually die upon embryonic day 10.5 as a consequence of incomplete vascular development. In contrast, Foxo3a−/− as well as Foxo4−/− mice can be viable and similar in appearance to controls, suggesting that at least for vascular development Foxo3a and Foxo4 are not critical (Hosaka et al., 2004). Yet, the studies on Foxo3a −/− null mice suggest that in the absence of Foxo3a, development of specific organs can be affected that may later have consequences for the nervous system. Foxo3a −/− null mice become infertile and have been reported to experience follicular activation to the extent that ovarian follicles are depleted with the subsequent death of oocytes (Castrillon et al., 2003). In addition, absence of Foxo3a leads to lymphoproliferation, organ inflammation of the salivary glands, lung, and kidney, and increased activity of helper T cells, indicating an important role for Foxo3a in immune regulation and T cell tolerance (Lin et al., 2004). Foxo3a also regulates hematopoietic stem cell development (Miyamoto et al., 2007) and neutrophil activity. For example, hematopoietic stem cell numbers are significantly decreased in aged Foxo3−/− mice compared to the littermate controls, suggesting that Foxo3a plays a pivotal role in maintaining the these stem cell pool (Miyamoto et al., 2007). Foxo3a null mice also are resistant to two models of neutrophilic inflammation that involve immune complex-mediated inflammatory arthritis and thioglycollate-induced peritonitis (Jonsson et al., 2005). In addition, Foxo-deficient mice (Foxo1, Foxo3a, Foxo4) also exhibit an increase in reactive oxygen species (ROS) and changes in the expression of regulated-genes of ROS in hematopoietic stem cell populations. Thus, Foxo3a proteins may play an essential role in the response to physiologic oxidative stress (Tothova et al., 2007).

CLINICAL IMPLICATIONS FOR FOXO3a IN THE NERVOUS SYSTEM

Ischemia and Oxidative Stress

FOXO3a may play a significant role during injuries that involve cerebral ischemia and oxidative stress (Chong et al., 2005b, Chong et al., 2005e). Hypoxic-ischemic conditions as well as oxidative stress represent significant pathways for the apoptotic destruction of cells (Chong et al., 2005f, Li et al., 2006a). Apoptosis consists of membrane phosphatidylserine (PS) exposure and DNA fragmentation (Maiese et al., 2005a) and can contribute to a variety of disease states that especially involve the nervous system such as diabetes, stroke, vascular dementia, Alzheimer’s disease, and trauma (Chong et al., 2004b, Han and Suk, 2005, Li et al., 2004a). As an early event in the dynamics of cellular apoptosis, membrane PS externalization can become a signal for the phagocytosis of cells (Chong et al., 2005e, Hong et al., 2004). Cells expressing externalized PS may be removed by microglia (Li et al., 2004b, Lin and Maiese, 2001) as well as lead to the dysfunction of inflammatory cells (Chong et al., 2007b, Li et al., 2006b).

In clinical work, the effects of the genetic variance in FOXO3a on metabolic profile, age-related diseases, fertility, fecundity, and mortality were analyzed in 1245 aged persons. Increased risk of stroke was observed in two haplotypes of FOXO3a block-A suggesting an association with FOXO3a (Kuningas et al., 2007). In animal studies, Foxo3a expression has been associated with c-Jun N-terminal kinase 3 activity and the potential to promote stroke (Pirianov et al., 2007). Other studies of cerebral ischemia also have indicated that inhibitory phosphorylation of Foxo3a may be associated with the neuroprotective effects of estradiol during stroke (Won et al., 2006). In neuronal cell culture models, inhibition or gene knockdown of Foxo3a activity can mediate the ischemic protective effects of metabotropic glutamate receptors (Chong et al., 2006), enhance neuronal survival through NAD+ precursors (Chong et al., 2004c, Li et al., 2006a, Li et al., 2004b, Maiese and Chong, 2003, Maiese et al., 2007), and offer trophic factor protection in both neurons and vascular cells during oxidative stress such as with erythropoietin (Chong and Maiese, 2007a), insulin-like growth factor-1 (Zheng et al., 2000), neurotrophins (Zheng et al., 2002), and vascular endothelial growth factor (Abid et al., 2004). In fact, agents that rely upon the modulation of FOXO3a, such as erythropoietin, commonly employ a number of pathways tied to FOXO3a signaling such as protein kinase B (Akt) (Chong et al., 2003a, Chong et al., 2003b, Chong et al., 2002a, Chong et al., 2005d, Chong et al., 2003c, Chong and Maiese, 2007a).

Diabetes, Metabolism, and Cell Senescence

Diabetes mellitus is believed to affect more than 165 million individuals worldwide (Quinn, 2001) and by the year 2030 more than 360 million individuals will be afflicted with diabetes (Maiese et al., 2007) and its debilitating conditions (Wild et al., 2004). Type 2 diabetes represents at least 80 percent of all diabetics and is dramatically increasing in incidence as a result of changes in human behavior and increased body mass index (Laakso, 2001). Type 1 insulin-dependent diabetes mellitus accounts for only 5–10 percent of all diabetics. Yet, disease of the nervous system can become one of the most debilitating complications for diabetes and affect sensitive neuronal and vascular regions of the brain (Chong et al., 2007c) resulting in significant functional impairment and dementia (Awad et al., 2004, Gerozissis, 2003). Furthermore, both focal and generalized neuropathies, especially in conjunction with vascular disease, can result in severe disability (Perkins and Bril, 2002).

The forkhead transcription factor family may be involved in pathways responsible for cell metabolism, diabetes mellitus onset, and diabetic complications (Chong and Maiese, 2007b, Maiese et al., 2005a, Maiese et al., 2004, Maiese et al., 2005b). In a prospective population based study of 1245 participants aged 85 years or more, haplotype analyses of FOXO1a revealed that carriers of haplotype 3 ‘TCA’ had higher HbA1c levels to suggest evidence of at least disorders with glucose intolerance (Kuningas et al., 2007). In particular for FOXO1, this transcription factor may be responsible for the prevention of beta cell injury during against oxidative stress. This process requires the complex formation with the pro-myelocytic leukemia protein Pml and the NAD-dependent deacetylase SIRT1 to activate expression of the insulin2 gene transcription factors NeuroD and MafA (Kitamura et al., 2005). Yet, the role of Forkhead transcription factors can vary among different cells and tissues. For example, mice overexpressing FOXO1 in skeletal muscle suffered from reduced skeletal muscle mass and poor glycemic control (Kamei et al., 2004). In regards to FOXO3a, a clinical exam of 734 individuals that sequenced all exons of the FOXO genes FOXO1a, FOXO3a, and FOXO4 found one promoter single nucleotide polymorphism in the 5′ flanking region of FOXO3a that displayed a significant association with body mass index such that the highest body mass index was present in individuals homozygous for this allele (Kim et al., 2006). In experimental models, administration of a high-fat diet induced hyperinsulinemic insulin-resistant obesity, increased expression of Foxo3a, and impaired cardiomyocyte function (Relling et al., 2006). Other work has linked diabetic nephropathy to Foxo3a by demonstrating that phosphorylation of Foxo3a increased in rat and mouse renal cortical tissues two weeks after the induction of diabetes by streptozotocin (Kato et al., 2006). Furthermore, enteric neurons can be protected from hyperglycemia by glial cell line-derived neurotrophic factor that can affect protein kinase B signaling and prevent Foxo3a activation and nuclear translocation (Anitha et al., 2006). Yet, some studies in other experimental models suggest that Foxo3a may block oxidative stress during diabetes. For example, interferon –gamma driven expression of tryptophan catabolism by cytotoxic T lymphocyte antigen 4 may activate Foxo3a to protect dendritic cells from oxidative stress in nonobese diabetic mice (Fallarino et al., 2004).

Closely linked to the diabetic complications that can affect the nervous system is the regulation of cellular metabolism. One potential pathway to consider is the precursor for the coenzyme β-nicotinamide adenine dinucleotide (NAD+), namely nicotinamide (Li et al., 2006a, Li et al., 2004b, Maiese and Chong, 2003). Nicotinamide can be directly utilized by cells to synthesize NAD+ (Jackson et al., 1995) and participates in energy metabolism through the tricarboxylic acid cycle by utilizing NAD+ in the mitochondrial respiratory electron transport chain for the production of ATP, DNA synthesis, and DNA repair (Hageman and Stierum, 2001, Magni et al., 2004). Furthermore, nicotinamide can significantly increase NAD+ levels in vulnerable regions of the ischemic brain, suggesting that nicotinamide may offer cytoprotection of injured tissue through the maintenance of NAD+ levels (Sadanaga-Akiyoshi et al., 2003). Administration of nicotinamide can significantly improved glucose utilization, prevent excessive lactate production and improve electrophysiologic capacity in ischemic animal models (Li et al., 2006a).

Treatment with nicotinamide can approximate normal fasting blood glucose in animals with streptozotocin-induced diabetes (Hu et al., 1996, Reddy et al., 1995). Oral nicotinamide (1200mg/m2/day) protects β-cell function and prevents clinical disease in islet-cell antibody-positive first-degree relatives of type-1 diabetes (Olmos et al., 2005). In addition, treatment with nicotinamide (25mg/kg) in patients with recent onset type-1 diabetes mellitus combined with intensive insulin therapy for up to two years after diagnosis can reduce HbA1c levels (Crino et al., 2005). Potentially relevant to diabetic patients with renal failure, nicotinamide also has been shown to reduce intestinal absorption of phosphate and prevent the development of hyperphosphatemia and progressive renal dysfunction (Eto et al., 2005). Therefore, it should be of interest that nicotinamide may derive its protective capacity through two separate mechanisms of post-translational modification of Foxo3a. Nicotinamide can not only maintain inhibitory phosphorylation of Foxo3a, but also preserve the integrity of total Foxo3a and phosphorylated Foxo3a (Chong et al., 2004c) to block Foxo3a proteolysis that can yield potentially apoptotic amino-terminal (Nt) fragments (Charvet et al., 2003).

However, nicotinamide has many roles especially in relation to sirtuins. Nicotinamide offers protection in millimole concentrations against free radicals (Lin et al., 2000), anoxia (Lin et al., 2001), and oxygen glucose deprivation (Chong et al., 2004c). Yet in relation to cell longevity, nicotinamide is an inhibitor of sirtuins in concentrations that range from 50–100 μM (Porcu and Chiarugi, 2005). Increased longevity in yeast has been shown to be dependent upon silent information regulator 2 (Sir2) protein. The Sir2 gene belongs to a family of genes which is a highly conserved group in the genomes of organisms ranging from archaebacteria to eukaryotes (Frye, 2000, Vaziri et al., 2001). Physiological concentrations of nicotinamide noncompetitively inhibit Sir2, suggesting that nicotinamide is a physiologically relevant regulator of Sir2 enzymes (Bitterman et al., 2002). Other studies have shown that stimulation of the NAD+-dependent deacetylase SIRT1, a mammalian ortholog of Sir2, in mammalian cells during starvation is dependent upon Foxo3a expression (Nemoto et al., 2004). This process may require a close feedback loop since SIRT1 can deacetylate Foxo3a during oxidative stress (Wang et al., 2007) and inhibit its activity (Motta et al., 2004).

Irrespective of the direct relationship to sirtuins, FOXO3a also has been linked to cell aging and senescence (Kyoung Kim et al., 2005, Ramnanan et al., 2007), suggesting a more broader impact for FOXO3a in studies that involve the effects of neurodegeneration associated with cognitive decline (Chong et al., 2007a, Chong et al., 2005c, Chong et al., 2005f, Maiese et al., 2005a). It has been shown that protein kinase B (Akt) activity can increase with cellular senescence in primary cultured human endothelial cells and inhibition of FOXO3a by Akt is essential for this growth arrest to occur (Miyauchi et al., 2004). Interestingly, FOXO3a inactivation in older animals may limit tissue antioxidant properties through decreased manganese-superoxide dismutase and lead to enhanced cell susceptibility to injury with aging (Li et al., 2006c).

Inflammatory and Neuromuscular Disease

As previously noted, FOXO3a has a significant role in the immune system as evidenced by studies that illustrate that Foxo3a-deficient mice are resistant to neutrophilic inflammation and immune complex-mediated inflammatory arthritis (Jonsson et al., 2005). For the nervous system, control of systemic lupus erythematosus (SLE) may require Foxo3a activation. Amelioration of SLE in animal models is achieved with administration of peptides based upon the complementary determining region-1 of an anti-DNA antibody. In these studies, transforming growth factor-beta (TGF-β) is increased with subsequent upregulation of Foxo3a that is believed to block NF-kappaB activation and interferon-gamma secretion (Sela et al., 2006). In addition, models of multiple sclerosis with experimental autoimmune encephalomyelitis have shown that osteopontin, a protein expressed in multiple sclerosis lesions, was found to block Foxo3a activity that subsequently leads to activation of nuclear factor kappaB and promote the activity of myelin-reactive T cells resulting in disease progression (Hur et al., 2007).

A number of disease entities are associated with skeletal muscle atrophy, such as cancer, diabetes, cardiac and renal failure, and sepsis. In muscles that undergo atrophy, the ubiquitin ligase atrogin-1 is increased that leads to muscle loss. In studies using cultured myotubes, atrophy was linked to the activation of Foxo3a (Sandri et al., 2004). However, in clinical studies involving biopsies from twenty-two patients with amyotrophic lateral sclerosis, patients had a significant increase in atrogin-1 mRNA and protein content, but no detectable increase in FOXO3a mRNA and protein content (Leger et al., 2006). This work may suggest that other factors in addition to FOXO3a are necessary to initiate neuromuscular disease.

FOXO3a SIGNALING IN THE NERVOUS SYSTEM

A number of regulatory pathways are essential to control cell survival, such as the phosphatidylinositol 3-kinase (PI 3-K) pathway through protein kinase B (Akt) (Chong et al., 2005e, Chong and Maiese, 2007a). Regulation of Akt activity can affect a number of cellular functions that relate to growth, survival, injury, and metabolism. In particular, Akt can prevent apoptotic cell injury during late genomic DNA destruction and early apoptotic signaling with membrane PS exposure (Chong et al., 2005f, Maiese and Chong, 2004). Enhanced Akt activity can increase cell survival during several injuries such as free radical exposure (Chong et al., 2003b, Matsuzaki et al., 1999), matrix detachment (Rytomaa et al., 2000), neuronal axotomy (Namikawa et al., 2000), DNA damage (Chong et al., 2004a, Chong et al., 2002a, Henry et al., 2001, Kang et al., 2003a), oxidative stress (Chong et al., 2003b, Kang et al., 2003a, Kang et al., 2003b, Yamaguchi and Wang, 2001), hypoxic preconditioning (Wick et al., 2002), β-amyloid exposure (Martin et al., 2001), metabotropic receptor signaling (Chong et al., 2005a, Chong et al., 2006, Maiese et al., 2005a), and cell metabolic pathways (Chong et al., 2005g, Maiese and Chong, 2003). Activation of Akt also can prevent membrane PS exposure on injured cells (Chong et al., 2004a, Li et al., 2006b) and block the activation of microglia during oxidative stress (Chong et al., 2005a, Kang et al., 2003a, Kang et al., 2003b).

Intimately linked to these survival pathways with Akt is FOXO3a (Fig. 1). Akt can phosphorylate FOXO3a and inhibit its activity to sequester FOXO3a in the cytoplasm by association with 14-3-3 proteins (Brunet et al., 2002, Chong and Maiese, 2007a, Dong et al., 2007, Kino et al., 2005, Munoz-Fontela et al., 2007). In the absence of inhibitory Akt1 phosphorylation, FOXO3a is activate, can translocate to the nucleus, and controls a variety of functions that involve cell cycle progression, cell longevity, and apoptosis (Lehtinen et al., 2006, Li et al., 2006a, Maiese et al., 2007). Site-directed mutagenesis of the potential Akt phosphorylation has shown that FOXO3a is phosphorylated by Akt on Thr-32, Ser-253 and Ser-315. Each phosphorylation site for FOXO3a may function independently that determines subcellar localization. Akt may first phosphorylate FOXO3a at Ser-253, create a docking motif for 14-3-3 proteins, which then may participate in the nuclear export of FOXO3a (Brunet et al., 1999, Cahill et al., 2001). A 14-3-3 dimer requires stable binding to the phosphorylated N-terminal Akt site prior to the other half of the dimer binding to the phosphorylated Akt motif in the forkhead domain (Van Der Heide et al., 2004).

Fig. 1. Potential cellular signaling pathways for FOXO3a.

Fig. 1

A representative illustration of possible “upstream” and “downstream” signal transduction pathways for the FOXO3a system that may be set in motion, such as during the presence of the growth factor erythropoietin (EPO).

Additional mechanisms exist independent from Akt that can phosphorylate and retain Foxo3a in the cytoplasm. As a close homolog to Akt, serum- and glucocorticoid-inducible protein kinase (Sgk) can increase cell survival through inhibition of Foxo3a (Leong et al., 2003). However, Sgk and Akt have different efficacies with which they phosphorylate Foxo3a. Both kinases can phosphorylate Thr-32, but Sgk preferentially phosphorylates Ser-315 whereas Akt favors Ser-253. This work suggests that Sgk and Akt may control Foxo3a through different phopshorylation sites (Brunet et al., 2001). The ability of Sgk to oversee Foxo3a activity also may involve p53. In studies that examine DNA damage, p53 led to the induction of Sgk that was necessary to maintain phosphorylation of Foxo3a (You et al., 2004). Additional work also has demonstrated that IkappaB kinase (IKK) in the absence of activated Akt can phosphorylate Foxo3a and retain Foxo3a in the cytoplasm (Finnberg and El-Deiry, 2004, Hu et al., 2004).

Downstream from the regulation of Foxo3a, a number of target genes for Foxo3a have been elucidated (Fig. 1). For example, FOXO3a activation in colonic cancer cells resulted in the expression of several members of Myc target genes that involve the Mad/Mxd family of transcriptional repressors and may account for the growth inhibitory effects of FOXO3a (Delpuech et al., 2007). In relation to apoptotic death, Foxo3a has been shown to modulate a ligand activating a Fas-mediated death pathway (Barthelemy et al., 2004) and to induce tumor-necrosis-factor-related apoptosis-induced ligand (TRAIL) and the BH3-only proteins Noxa and Bim (Obexer et al., 2007). FOXO3a also can modulate growth-arrest and DNA-damage-response protein 45 (Gadd45) and influence cell cycle regulation (Furukawa-Hibi et al., 2002, Tran et al., 2002). In regards to aging in a model of vascular smooth muscle cells, FOXO3a inactivation in older rats was found to reduce the transcription of manganese-superoxide dismutase, a major cellular anti-oxidant that may work to reduce cell disability with advancing age (Li et al., 2006c). Furthermore, immune system regulation by FOXO3a is believed to be mediated, in part, through the regulation of nuclear factor-kappa B (NF-kappaB) (Lin et al., 2004).

FUTURE PERSPECTIVES

There are several levels to target the activity of FOXO3a. FOXO3a may be regulated at three distinct levels. First, FOXO3a-dependent transcription depends upon the translocation of FOXO3a to the nucleus. As a result, kinases such as Akt exert control over FOXO3a through phosphorylation to sequester FOXO3a in the cytoplasm through association with 14-3-3 proteins (Brunet et al., 2002, Chong and Maiese, 2007a, Dong et al., 2007, Kino et al., 2005, Munoz-Fontela et al., 2007). Without Akt phosphorylation, FOXO3a is activate to translocate to the cell nucleus (Lehtinen et al., 2006, Li et al., 2006a, Maiese et al., 2007). Second, the integrity of the FOXO3a protein represents another level of regulation. IKK can phosphorylate and lead to proteolysis of FOXO3a via the Ub-dependent proteasome pathway (Hu et al., 2004). In addition, modulation of caspase 3 activity also appears to be closely associated with a unique regulatory mechanism that blocks the proteolytic degradation of phosphorylated FOXO3a. FOXO3a has been shown to be a substrate for caspase 3-like proteases at the consensus sequence DELD304A (Charvet et al., 2003). As a result, blockade of caspase 3 - like activity can prevent the destruction of phosphorylated FOXO3a during conditions of oxidative stress (Chong et al., 2004c, Chong and Maiese, 2007a). Agents that prevent FOXO3a degradation through caspase 3 inhibition (Chong et al., 2002b, Chong et al., 2004c, Chong and Maiese, 2007a) can be cytoprotective since proteolysis of FOXO3a has been shown to yield amino-terminal (Nt) fragments that may further precipitate cellular injury (Charvet et al., 2003). As additional mechanisms, FOXO3a can become acetylated, such as during growth factor deprivation (Mahmud et al., 2002), or deacetylated by the sirtuin Sir2 homolog SIRT1 (Brunet et al., 2004, Motta et al., 2004, Wang et al., 2007), to either disrupt FOXO3a association with target DNA or to alter the ability of FOXO3a to lead to cell cycle arrest and cell death.

Future directives that focus upon FOXO3a as a therapeutic target should consider the multiple pathways that can govern FOXO3a activity. In relation to cell survival, trophic factors such as erythropoietin, insulin growth factor, and brain-derived growth factor can inhibit activity of FOXO3a and sequester it in the cytoplasm of cells (Chong et al., 2006, Chong et al., 2004c, Chong and Maiese, 2007a, Zheng et al., 2000, Zhu et al., 2004). Although several agents that include growth factors, metabotropic receptors, and NAD+ precursors (Chong et al., 2006, Chong et al., 2004c, Chong and Maiese, 2007a) can prevent cell injury through reductions in FOXO3a activity, future work that intends to prevent apoptotic injury should strive to modulate FOXO3a at multiple levels that involve inhibitory phosphorylation of FOXO3a, prevention of phosphorylated FOXO3a destruction, cytoplasmic retention of FOXO3a, and gene knockdown of FOXO3a. Yet, it is clear that such a recipe is overly simplistic and must proceed with caution, since FOXO3a activity is required as a significant regulator of cell growth, cell senescence, and in some cases for the prevention of invasive neoplasms (Belguise et al., 2007, Real et al., 2005, Rice et al., 2007).

Acknowledgments

This research was supported by the following grants (KM): American Diabetes Association, American Heart Association (National), Bugher Foundation Award, Janssen Neuroscience Award, LEARN Foundation Award, MI Life Sciences Challenge Award, Nelson Foundation Award, NIH NIEHS (P30 ES06639), and NIH NINDS/NIA.

References

  1. Abid MR, Guo S, Minami T, Spokes KC, Ueki K, Skurk C, Walsh K, Aird WC. Vascular endothelial growth factor activates PI3K/Akt/forkhead signaling in endothelial cells. Arterioscler Thromb Vasc Biol. 2004;24:294–300. doi: 10.1161/01.ATV.0000110502.10593.06. [DOI] [PubMed] [Google Scholar]
  2. Anderson MJ, Viars CS, Czekay S, Cavenee WK, Arden KC. Cloning and characterization of three human forkhead genes that comprise an FKHR-like gene subfamily. Genomics. 1998;47:187–99. doi: 10.1006/geno.1997.5122. [DOI] [PubMed] [Google Scholar]
  3. Anitha M, Gondha C, Sutliff R, Parsadanian A, Mwangi S, Sitaraman SV, Srinivasan S. GDNF rescues hyperglycemia-induced diabetic enteric neuropathy through activation of the PI3K/Akt pathway. J Clin Invest. 2006;116:344–56. doi: 10.1172/JCI26295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Awad N, Gagnon M, Messier C. The relationship between impaired glucose tolerance, type 2 diabetes, and cognitive function. J Clin Exp Neuropsychol. 2004;26:1044–80. doi: 10.1080/13803390490514875. [DOI] [PubMed] [Google Scholar]
  5. Barthelemy C, Henderson CE, Pettmann B. Foxo3a induces motoneuron death through the Fas pathway in cooperation with JNK. BMC Neurosci. 2004;5:48. doi: 10.1186/1471-2202-5-48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Belguise K, Guo S, Sonenshein GE. Activation of FOXO3a by the green tea polyphenol epigallocatechin-3-gallate induces estrogen receptor {alpha} expression reversing invasive phenotype of breast cancer cells. Cancer Res. 2007;67:5763–70. doi: 10.1158/0008-5472.CAN-06-4327. [DOI] [PubMed] [Google Scholar]
  7. Bitterman KJ, Anderson RM, Cohen HY, Latorre-Esteves M, Sinclair DA. Inhibition of silencing and accelerated aging by nicotinamide, a putative negative regulator of yeast sir2 and human SIRT1. J Biol Chem. 2002;277:45099–107. doi: 10.1074/jbc.M205670200. [DOI] [PubMed] [Google Scholar]
  8. Brunet A, Bonni A, Zigmond MJ, Lin MZ, Juo P, Hu LS, Anderson MJ, Arden KC, Blenis J, Greenberg ME. Akt promotes cell survival by phosphorylating and inhibiting a Forkhead transcription factor. Cell. 1999;96:857–68. doi: 10.1016/s0092-8674(00)80595-4. [DOI] [PubMed] [Google Scholar]
  9. Brunet A, Kanai F, Stehn J, Xu J, Sarbassova D, Frangioni JV, Dalal SN, DeCaprio JA, Greenberg ME, Yaffe MB. 14-3-3 transits to the nucleus and participates in dynamic nucleocytoplasmic transport. J Cell Biol. 2002;156:817–28. doi: 10.1083/jcb.200112059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Brunet A, Park J, Tran H, Hu LS, Hemmings BA, Greenberg ME. Protein kinase SGK mediates survival signals by phosphorylating the forkhead transcription factor FKHRL1 (FOXO3a) Mol Cell Biol. 2001;21:952–65. doi: 10.1128/MCB.21.3.952-965.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Brunet A, Sweeney LB, Sturgill JF, Chua KF, Greer PL, Lin Y, Tran H, Ross SE, Mostoslavsky R, Cohen HY, Hu LS, Cheng HL, Jedrychowski MP, Gygi SP, Sinclair DA, Alt FW, Greenberg ME. Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science. 2004;303:2011–5. doi: 10.1126/science.1094637. [DOI] [PubMed] [Google Scholar]
  12. Cahill CM, Tzivion G, Nasrin N, Ogg S, Dore J, Ruvkun G, Alexander-Bridges M. Phosphatidylinositol 3-kinase signaling inhibits DAF-16 DNA binding and function via 14-3-3-dependent and 14-3-3-independent pathways. J Biol Chem. 2001;276:13402–10. doi: 10.1074/jbc.M010042200. [DOI] [PubMed] [Google Scholar]
  13. Castrillon DH, Miao L, Kollipara R, Horner JW, DePinho RA. Suppression of ovarian follicle activation in mice by the transcription factor Foxo3a. Science. 2003;301:215–8. doi: 10.1126/science.1086336. [DOI] [PubMed] [Google Scholar]
  14. Charvet C, Alberti I, Luciano F, Jacquel A, Bernard A, Auberger P, Deckert M. Proteolytic regulation of Forkhead transcription factor FOXO3a by caspase-3-like proteases. Oncogene. 2003;22:4557–68. doi: 10.1038/sj.onc.1206778. [DOI] [PubMed] [Google Scholar]
  15. Chong ZZ, 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. 2005a;2:197–211. doi: 10.2174/1567202054368317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Chong ZZ, Kang JQ, Maiese K. Akt1 drives endothelial cell membrane asymmetry and microglial activation through Bcl-x(L) and caspase 1, 3, and 9. Exp Cell Res. 2004a;296:196–207. doi: 10.1016/j.yexcr.2004.01.021. [DOI] [PubMed] [Google Scholar]
  17. Chong ZZ, Kang JQ, 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. 2003a;23:320–30. doi: 10.1097/01.WCB.0000050061.57184.AE. [DOI] [PubMed] [Google Scholar]
  18. Chong ZZ, Kang JQ, Maiese K. Erythropoietin fosters both intrinsic and extrinsic neuronal protection through modulation of microglia, Akt1, Bad, and caspase-mediated pathways. Br J Pharmacol. 2003b;138:1107–1118. doi: 10.1038/sj.bjp.0705161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Chong ZZ, Kang JQ, Maiese K. Erythropoietin is a novel vascular protectant through activation of Akt1 and mitochondrial modulation of cysteine proteases. Circulation. 2002a;106:2973–9. doi: 10.1161/01.cir.0000039103.58920.1f. [DOI] [PubMed] [Google Scholar]
  20. Chong ZZ, Kang JQ, Maiese K. Essential cellular regulatory elements of oxidative stress in early and late phases of apoptosis in the central nervous system. Antioxid Redox Signal. 2004b;6:277–87. doi: 10.1089/152308604322899341. [DOI] [PubMed] [Google Scholar]
  21. Chong ZZ, Li F, Maiese K. Activating Akt and the brain’s resources to drive cellular survival and prevent inflammatory injury. Histol Histopathol. 2005b;20:299–315. doi: 10.14670/hh-20.299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Chong ZZ, Li F, Maiese K. Cellular demise and inflammatory microglial activation during beta-amyloid toxicity are governed by Wnt1 and canonical signaling pathways. Cell Signal. 2007a;19:1150–62. doi: 10.1016/j.cellsig.2006.12.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Chong ZZ, Li F, Maiese K. Employing new cellular therapeutic targets for Alzheimer’s disease: a change for the better? Curr Neurovasc Res. 2005c;2:55–72. doi: 10.2174/1567202052773508. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Chong ZZ, Li F, Maiese K. Erythropoietin requires NF-kappaB and its nuclear translocation to prevent early and late apoptotic neuronal injury during beta-amyloid toxicity. Curr Neurovasc Res. 2005d;2:387–99. doi: 10.2174/156720205774962683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Chong ZZ, Li F, Maiese K. Group I metabotropic receptor neuroprotection requires Akt and its substrates that govern FOXO3a, Bim, and beta-catenin during oxidative stress. Curr Neurovasc Res. 2006;3:107–17. doi: 10.2174/156720206776875830. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Chong ZZ, Li F, Maiese K. Oxidative stress in the brain: Novel cellular targets that govern survival during neurodegenerative disease. Prog Neurobiol. 2005e;75:207–46. doi: 10.1016/j.pneurobio.2005.02.004. [DOI] [PubMed] [Google Scholar]
  27. Chong ZZ, Li F, Maiese K. Stress in the brain: novel cellular mechanisms of injury linked to Alzheimer’s disease. Brain Res Brain Res Rev. 2005f;49:1–21. doi: 10.1016/j.brainresrev.2004.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Chong ZZ, Li F, Maiese K. 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. 2007b;19:263–72. [PMC free article] [PubMed] [Google Scholar]
  29. Chong ZZ, Lin SH, Kang JQ, Maiese K. Erythropoietin prevents early and late neuronal demise through modulation of Akt1 and induction of caspase 1, 3, and 8. J Neurosci Res. 2003c;71:659–69. doi: 10.1002/jnr.10528. [DOI] [PubMed] [Google Scholar]
  30. Chong ZZ, Lin SH, 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. 2005g;2:271–85. doi: 10.2174/156720205774322584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Chong ZZ, Lin SH, Maiese K. Nicotinamide Modulates Mitochondrial Membrane Potential and Cysteine Protease Activity during Cerebral Vascular Endothelial Cell Injury. J Vasc Res. 2002b;39:131–47. doi: 10.1159/000057762. [DOI] [PubMed] [Google Scholar]
  32. Chong ZZ, Lin SH, 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. 2004c;24:728–43. doi: 10.1097/01.WCB.0000122746.72175.0E. [DOI] [PubMed] [Google Scholar]
  33. Chong ZZ, 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. 2007a;150:839–50. doi: 10.1038/sj.bjp.0707161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Chong ZZ, Maiese K. The Src homology 2 domain tyrosine phosphatases SHP-1 and SHP-2: diversified control of cell growth, inflammation, and injury. Histol Histopathol. 2007b;22:1251–67. doi: 10.14670/hh-22.1251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Chong ZZ, Shang YC, Maiese K. Vascular injury during elevated glucose can be mitigated by erythropoietin and Wnt signaling. Curr Neurovasc Res. 2007c;4:194–204. doi: 10.2174/156720207781387150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Clark KL, Halay ED, Lai E, Burley SK. Co-crystal structure of the HNF-3/fork head DNA-recognition motif resembles histone H5. Nature. 1993;364:412–20. doi: 10.1038/364412a0. [DOI] [PubMed] [Google Scholar]
  37. Crino A, Schiaffini R, Ciampalini P, Suraci MC, Manfrini S, Visalli N, Matteoli MC, Patera P, Buzzetti R, Guglielmi C, Spera S, Costanza F, Fioriti E, Pitocco D, Pozzilli P. A two year observational study of nicotinamide and intensive insulin therapy in patients with recent onset type 1 diabetes mellitus. J Pediatr Endocrinol Metab. 2005;18:749–54. doi: 10.1515/jpem.2005.18.8.749. [DOI] [PubMed] [Google Scholar]
  38. Delpuech O, Griffiths B, East P, Essafi A, Lam EW, Burgering B, Downward J, Schulze A. Induction of Mxi1-SR{alpha} by FOXO3a contributes to repression of Myc-dependent gene expression. Mol Cell Biol. 2007;27:4917–30. doi: 10.1128/MCB.01789-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Dong S, Kang S, Gu TL, Kardar S, Fu H, Lonial S, Khoury HJ, Khuri F, Chen J. 14-3-3 integrates prosurvival signals mediated by the AKT and MAPK pathways in ZNF198-FGFR1-transformed hematopoietic cells. Blood. 2007;110:360–9. doi: 10.1182/blood-2006-12-065615. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Eto N, Miyata Y, Ohno H, Yamashita T. Nicotinamide prevents the development of hyperphosphataemia by suppressing intestinal sodium-dependent phosphate transporter in rats with adenine-induced renal failure. Nephrol Dial Transplant. 2005;20:1378–84. doi: 10.1093/ndt/gfh781. [DOI] [PubMed] [Google Scholar]
  41. Fallarino F, Bianchi R, Orabona C, Vacca C, Belladonna ML, Fioretti MC, Serreze DV, Grohmann U, Puccetti P. CTLA-4-Ig activates forkhead transcription factors and protects dendritic cells from oxidative stress in nonobese diabetic mice. J Exp Med. 2004;200:1051–62. doi: 10.1084/jem.20040942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Finnberg N, El-Deiry WS. Activating FOXO3a, NF-kappaB and p53 by targeting IKKs: an effective multi-faceted targeting of the tumor-cell phenotype? Cancer Biol Ther. 2004;3:614–6. doi: 10.4161/cbt.3.7.1057. [DOI] [PubMed] [Google Scholar]
  43. Frye RA. Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochem Biophys Res Commun. 2000;273:793–8. doi: 10.1006/bbrc.2000.3000. [DOI] [PubMed] [Google Scholar]
  44. Furukawa-Hibi Y, Yoshida-Araki K, Ohta T, Ikeda K, Motoyama N. FOXO forkhead transcription factors induce G(2)-M checkpoint in response to oxidative stress. J Biol Chem. 2002;277:26729–32. doi: 10.1074/jbc.C200256200. [DOI] [PubMed] [Google Scholar]
  45. Furuyama T, Nakazawa T, Nakano I, Mori N. Identification of the differential distribution patterns of mRNAs and consensus binding sequences for mouse DAF-16 homologues. Biochem J. 2000;349:629–34. doi: 10.1042/0264-6021:3490629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Gerozissis K. Brain insulin: regulation, mechanisms of action and functions. Cell Mol Neurobiol. 2003;23:1–25. doi: 10.1023/A:1022598900246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Hageman GJ, Stierum RH. Niacin, poly(ADP-ribose) polymerase-1 and genomic stability. Mutat Res. 2001;475:45–56. doi: 10.1016/s0027-5107(01)00078-1. [DOI] [PubMed] [Google Scholar]
  48. Han HS, Suk K. The function and integrity of the neurovascular unit rests upon the integration of the vascular and inflammatory cell systems. Curr Neurovasc Res. 2005;2:409–23. doi: 10.2174/156720205774962647. [DOI] [PubMed] [Google Scholar]
  49. Henry MK, Lynch JT, Eapen AK, Quelle FW. DNA damage-induced cell-cycle arrest of hematopoietic cells is overridden by activation of the PI-3 kinase/Akt signaling pathway. Blood. 2001;98:834–41. doi: 10.1182/blood.v98.3.834. [DOI] [PubMed] [Google Scholar]
  50. Hoekman MF, Jacobs FM, Smidt MP, Burbach JP. Spatial and temporal expression of FoxO transcription factors in the developing and adult murine brain. Gene Expr Patterns. 2006;6:134–40. doi: 10.1016/j.modgep.2005.07.003. [DOI] [PubMed] [Google Scholar]
  51. Hong JR, Lin GH, Lin CJ, Wang WP, Lee CC, Lin TL, Wu JL. Phosphatidylserine receptor is required for the engulfment of dead apoptotic cells and for normal embryonic development in zebrafish. Development. 2004;131:5417–27. doi: 10.1242/dev.01409. [DOI] [PubMed] [Google Scholar]
  52. Hosaka T, Biggs WH, 3rd, Tieu D, Boyer AD, Varki NM, Cavenee WK, Arden KC. Disruption of forkhead transcription factor (FOXO) family members in mice reveals their functional diversification. Proc Natl Acad Sci U S A. 2004;101:2975–80. doi: 10.1073/pnas.0400093101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Hu MC, Lee DF, Xia W, Golfman LS, Ou-Yang F, Yang JY, Zou Y, Bao S, Hanada N, Saso H, Kobayashi R, Hung MC. IkappaB kinase promotes tumorigenesis through inhibition of forkhead FOXO3a. Cell. 2004;117:225–37. doi: 10.1016/s0092-8674(04)00302-2. [DOI] [PubMed] [Google Scholar]
  54. Hu Y, Wang Y, Wang L, Zhang H, Zhao B, Zhang A, Li Y. Effects of nicotinamide on prevention and treatment of streptozotocin-induced diabetes mellitus in rats. Chin Med J (Engl) 1996;109:819–22. [PubMed] [Google Scholar]
  55. Hur EM, Youssef S, Haws ME, Zhang SY, Sobel RA, Steinman L. Osteopontin-induced relapse and progression of autoimmune brain disease through enhanced survival of activated T cells. Nat Immunol. 2007;8:74–83. doi: 10.1038/ni1415. [DOI] [PubMed] [Google Scholar]
  56. Jackson TM, Rawling JM, Roebuck BD, Kirkland JB. Large supplements of nicotinic acid and nicotinamide increase tissue NAD+ and poly(ADP-ribose) levels but do not affect diethylnitrosamine-induced altered hepatic foci in Fischer-344 rats. J Nutr. 1995;125:1455–61. doi: 10.1093/jn/125.6.1455. [DOI] [PubMed] [Google Scholar]
  57. Jin C, Marsden I, Chen X, Liao X. Sequence specific collective motions in a winged helix DNA binding domain detected by 15N relaxation NMR. Biochemistry. 1998;37:6179–87. doi: 10.1021/bi980031v. [DOI] [PubMed] [Google Scholar]
  58. Jonsson H, Allen P, Peng SL. Inflammatory arthritis requires Foxo3a to prevent Fas ligand-induced neutrophil apoptosis. Nat Med. 2005;11:666–71. doi: 10.1038/nm1248. [DOI] [PubMed] [Google Scholar]
  59. Kaestner KH, Knochel W, Martinez DE. Unified nomenclature for the winged helix/forkhead transcription factors. Genes Dev. 2000;14:142–6. [PubMed] [Google Scholar]
  60. Kamei Y, Miura S, Suzuki M, Kai Y, Mizukami J, Taniguchi T, Mochida K, Hata T, Matsuda J, Aburatani H, Nishino I, Ezaki O. Skeletal muscle FOXO1 (FKHR) transgenic mice have less skeletal muscle mass, down-regulated Type I (slow twitch/red muscle) fiber genes, and impaired glycemic control. J Biol Chem. 2004;279:41114–23. doi: 10.1074/jbc.M400674200. [DOI] [PubMed] [Google Scholar]
  61. Kang JQ, Chong ZZ, Maiese K. Akt1 protects against inflammatory microglial activation through maintenance of membrane asymmetry and modulation of cysteine protease activity. J Neurosci Res. 2003a;74:37–51. doi: 10.1002/jnr.10740. [DOI] [PubMed] [Google Scholar]
  62. Kang JQ, Chong ZZ, Maiese K. Critical role for Akt1 in the modulation of apoptotic phosphatidylserine exposure and microglial activation. Mol Pharmacol. 2003b;64:557–69. doi: 10.1124/mol.64.3.557. [DOI] [PubMed] [Google Scholar]
  63. Kato M, Yuan H, Xu ZG, Lanting L, Li SL, Wang M, Hu MC, Reddy MA, Natarajan R. Role of the Akt/FoxO3a pathway in TGF-beta1-mediated mesangial cell dysfunction: a novel mechanism related to diabetic kidney disease. J Am Soc Nephrol. 2006;17:3325–35. doi: 10.1681/ASN.2006070754. [DOI] [PubMed] [Google Scholar]
  64. Kim JR, Jung HS, Bae SW, Kim JH, Park BL, Choi YH, Cho HY, Cheong HS, Shin HD. Polymorphisms in FOXO gene family and association analysis with BMI. Obesity (Silver Spring, Md) 2006;14:188–93. doi: 10.1038/oby.2006.24. [DOI] [PubMed] [Google Scholar]
  65. Kino T, De Martino MU, Charmandari E, Ichijo T, Outas T, Chrousos GP. HIV-1 accessory protein Vpr inhibits the effect of insulin on the Foxo subfamily of forkhead transcription factors by interfering with their binding to 14-3-3 proteins: potential clinical implications regarding the insulin resistance of HIV-1-infected patients. Diabetes. 2005;54:23–31. doi: 10.2337/diabetes.54.1.23. [DOI] [PubMed] [Google Scholar]
  66. Kitamura YI, Kitamura T, Kruse JP, Raum JC, Stein R, Gu W, Accili D. FoxO1 protects against pancreatic beta cell failure through NeuroD and MafA induction. Cell Metab. 2005;2:153–63. doi: 10.1016/j.cmet.2005.08.004. [DOI] [PubMed] [Google Scholar]
  67. Kuningas M, Magi R, Westendorp RG, Slagboom PE, Remm M, van Heemst D. Haplotypes in the human Foxo1a and Foxo3a genes; impact on disease and mortality at old age. Eur J Hum Genet. 2007;15:294–301. doi: 10.1038/sj.ejhg.5201766. [DOI] [PubMed] [Google Scholar]
  68. Kyoung Kim H, Kyoung Kim Y, Song IH, Baek SH, Lee SR, Hye Kim J, Kim JR. Down-regulation of a forkhead transcription factor, FOXO3a, accelerates cellular senescence in human dermal fibroblasts. J Gerontol A Biol Sci Med Sci. 2005;60:4–9. doi: 10.1093/gerona/60.1.4. [DOI] [PubMed] [Google Scholar]
  69. Laakso M. Cardiovascular disease in type 2 diabetes: challenge for treatment and prevention. J Intern Med. 2001;249:225–35. doi: 10.1046/j.1365-2796.2001.00789.x. [DOI] [PubMed] [Google Scholar]
  70. Larson ET, Eilers B, Menon S, Reiter D, Ortmann A, Young MJ, Lawrence CM. A winged-helix protein from sulfolobus turreted icosahedral virus points toward stabilizing disulfide bonds in the intracellular proteins of a hyperthermophilic virus. Virology. 2007 doi: 10.1016/j.virol.2007.06.040. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  71. Leger B, Vergani L, Soraru G, Hespel P, Derave W, Gobelet C, D’Ascenzio C, Angelini C, Russell AP. Human skeletal muscle atrophy in amyotrophic lateral sclerosis reveals a reduction in Akt and an increase in atrogin-1. FASEB J. 2006;20:583–5. doi: 10.1096/fj.05-5249fje. [DOI] [PubMed] [Google Scholar]
  72. Lehtinen MK, Yuan Z, Boag PR, Yang Y, Villen J, Becker EB, Di-Bacco S, de la Iglesia N, Gygi S, Blackwell TK, Bonni A. A conserved MST-FOXO signaling pathway mediates oxidative-stress responses and extends life span. Cell. 2006;125:987–1001. doi: 10.1016/j.cell.2006.03.046. [DOI] [PubMed] [Google Scholar]
  73. Leong ML, Maiyar AC, Kim B, O’Keeffe BA, Firestone GL. Expression of the serum- and glucocorticoid-inducible protein kinase, Sgk, is a cell survival response to multiple types of environmental stress stimuli in mammary epithelial cells. J Biol Chem. 2003;278:5871–82. doi: 10.1074/jbc.M211649200. [DOI] [PubMed] [Google Scholar]
  74. Li F, Chong ZZ, Maiese K. Cell life versus cell longevity: the mysteries surrounding the NAD(+) precursor nicotinamide. Curr Med Chem. 2006a;13:883–95. doi: 10.2174/092986706776361058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Li F, Chong ZZ, Maiese K. Erythropoietin on a tightrope: balancing neuronal and vascular protection between intrinsic and extrinsic pathways. Neurosignals. 2004a;13:265–89. doi: 10.1159/000081963. [DOI] [PubMed] [Google Scholar]
  76. Li F, Chong ZZ, 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. 2006b;3:187–201. doi: 10.2174/156720206778018758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Li F, Chong ZZ, Maiese K. Navigating novel mechanisms of cellular plasticity with the NAD+ precursor and nutrient nicotinamide. Front Biosci. 2004b;9:2500–2520. doi: 10.2741/1412. [DOI] [PubMed] [Google Scholar]
  78. Li M, Chiu JF, Mossman BT, Fukagawa NK. Down-regulation of manganese-superoxide dismutase through phosphorylation of FOXO3a by Akt in explanted vascular smooth muscle cells from old rats. J Biol Chem. 2006c;281:40429–39. doi: 10.1074/jbc.M606596200. [DOI] [PubMed] [Google Scholar]
  79. Lin L, Hron JD, Peng SL. Regulation of NF-kappaB, Th activation, and autoinflammation by the forkhead transcription factor Foxo3a. Immunity. 2004;21:203–13. doi: 10.1016/j.immuni.2004.06.016. [DOI] [PubMed] [Google Scholar]
  80. Lin SH, Chong ZZ, Maiese K. Nicotinamide: A nutritional supplement that provides protection against neuronal and vascular injury. J Med Food. 2001;4:27–38. doi: 10.1089/10966200152053686. [DOI] [PubMed] [Google Scholar]
  81. Lin SH, Maiese K. The metabotropic glutamate receptor system protects against ischemic free radical programmed cell death in rat brain endothelial cells. J Cereb Blood Flow Metab. 2001;21:262–75. doi: 10.1097/00004647-200103000-00010. [DOI] [PubMed] [Google Scholar]
  82. Lin SH, Vincent A, Shaw T, Maynard KI, Maiese K. Prevention of nitric oxide-induced neuronal injury through the modulation of independent pathways of programmed cell death. J Cereb Blood Flow Metab. 2000;20:1380–91. doi: 10.1097/00004647-200009000-00013. [DOI] [PubMed] [Google Scholar]
  83. Magni G, Amici A, Emanuelli M, Orsomando G, Raffaelli N, Ruggieri S. Enzymology of NAD+ homeostasis in man. Cell Mol Life Sci. 2004;61:19–34. doi: 10.1007/s00018-003-3161-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Mahmud DL, G-Amlak M, Deb DK, Platanias LC, Uddin S, Wickrema A. Phosphorylation of forkhead transcription factors by erythropoietin and stem cell factor prevents acetylation and their interaction with coactivator p300 in erythroid progenitor cells. Oncogene. 2002;21:1556–62. doi: 10.1038/sj.onc.1205230. [DOI] [PubMed] [Google Scholar]
  85. Maiese K, Chong ZZ. Insights into oxidative stress and potential novel therapeutic targets for Alzheimer disease. Restor Neurol Neurosci. 2004;22:87–104. [PubMed] [Google Scholar]
  86. Maiese K, Chong ZZ. Nicotinamide: necessary nutrient emerges as a novel cytoprotectant for the brain. Trends Pharmacol Sci. 2003;24:228–32. doi: 10.1016/S0165-6147(03)00078-6. [DOI] [PubMed] [Google Scholar]
  87. Maiese K, Chong ZZ, Li F. Driving cellular plasticity and survival through the signal transduction pathways of metabotropic glutamate receptors. Curr Neurovasc Res. 2005a;2:425–46. doi: 10.2174/156720205774962692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Maiese K, Chong ZZ, Shang YC. Mechanisitic insights into diabetes mellitus and oxidative stress. Curr Med Chem. 2007;14:1689–1699. doi: 10.2174/092986707781058968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Maiese K, Li F, Chong ZZ. Erythropoietin in the brain: can the promise to protect be fulfilled? Trends Pharmacol Sci. 2004;25:577–583. doi: 10.1016/j.tips.2004.09.006. [DOI] [PubMed] [Google Scholar]
  90. Maiese K, Li F, Chong ZZ. New avenues of exploration for erythropoietin. JAMA. 2005b;293:90–5. doi: 10.1001/jama.293.1.90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Martin D, Salinas M, Lopez-Valdaliso R, Serrano E, Recuero M, Cuadrado A. Effect of the Alzheimer amyloid fragment Abeta(25–35) on Akt/PKB kinase and survival of PC12 cells. J Neurochem. 2001;78:1000–8. doi: 10.1046/j.1471-4159.2001.00472.x. [DOI] [PubMed] [Google Scholar]
  92. Matsuzaki H, Tamatani M, Mitsuda N, Namikawa K, Kiyama H, Miyake S, Tohyama M. Activation of Akt kinase inhibits apoptosis and changes in Bcl-2 and Bax expression induced by nitric oxide in primary hippocampal neurons. J Neurochem. 1999;73:2037–46. [PubMed] [Google Scholar]
  93. Miyamoto K, Araki KY, Naka K, Arai F, Takubo K, Yamazaki S, Matsuoka S, Miyamoto T, Ito K, Ohmura M, Chen C, Hosokawa K, Nakauchi H, Nakayama K, Nakayama KI, Harada M, Motoyama N, Suda T, Hirao A. Foxo3a is essential for maintenance of the hematopoietic stem cell pool. Cell Stem Cell. 2007;1:101–112. doi: 10.1016/j.stem.2007.02.001. [DOI] [PubMed] [Google Scholar]
  94. Miyauchi H, Minamino T, Tateno K, Kunieda T, Toko H, Komuro I. Akt negatively regulates the in vitro lifespan of human endothelial cells via a p53/p21-dependent pathway. EMBO J. 2004;23:212–20. doi: 10.1038/sj.emboj.7600045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Motta MC, Divecha N, Lemieux M, Kamel C, Chen D, Gu W, Bultsma Y, McBurney M, Guarente L. Mammalian SIRT1 represses forkhead transcription factors. Cell. 2004;116:551–63. doi: 10.1016/s0092-8674(04)00126-6. [DOI] [PubMed] [Google Scholar]
  96. Munoz-Fontela C, Marcos-Villar L, Gallego P, Arroyo J, Da Costa M, Pomeranz KM, Lam EW, Rivas C. Latent protein LANA2 from Kaposi’s sarcoma-associated herpesvirus interacts with 14-3-3 proteins and inhibits FOXO3a transcription factor. J Virol. 2007;81:1511–6. doi: 10.1128/JVI.01816-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Namikawa K, Honma M, Abe K, Takeda M, Mansur K, Obata T, Miwa A, Okado H, Kiyama H. Akt/protein kinase B prevents injury-induced motoneuron death and accelerates axonal regeneration. J Neurosci. 2000;20:2875–86. doi: 10.1523/JNEUROSCI.20-08-02875.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Nemoto S, Fergusson MM, Finkel T. Nutrient availability regulates SIRT1 through a forkhead-dependent pathway. Science. 2004;306:2105–8. doi: 10.1126/science.1101731. [DOI] [PubMed] [Google Scholar]
  99. Obexer P, Geiger K, Ambros PF, Meister B, Ausserlechner MJ. FKHRL1-mediated expression of Noxa and Bim induces apoptosis via the mitochondria in neuroblastoma cells. Cell Death Differ. 2007;14:534–47. doi: 10.1038/sj.cdd.4402017. [DOI] [PubMed] [Google Scholar]
  100. Olmos PR, Hodgson MI, Maiz A, Manrique M, De Valdes MD, Foncea R, Acosta AM, Emmerich MV, Velasco S, Muniz OP, Oyarzun CA, Claro JC, Bastias MJ, Tora LA. Nicotinamide protected first-phase insulin response (FPIR) and prevented clinical disease in first-degree relatives of type-1 diabetics. Diabetes Res Clin Pract. 2005;71:320–33. doi: 10.1016/j.diabres.2005.07.009. [DOI] [PubMed] [Google Scholar]
  101. Perkins BA, Bril V. Diagnosis and management of diabetic neuropathy. Curr Diab Rep. 2002;2:495–500. doi: 10.1007/s11892-002-0119-x. [DOI] [PubMed] [Google Scholar]
  102. Pirianov G, Brywe KG, Mallard C, Edwards AD, Flavell RA, Hagberg H, Mehmet H. Deletion of the c-Jun N-terminal kinase 3 gene protects neonatal mice against cerebral hypoxic-ischaemic injury. J Cereb Blood Flow Metab. 2007;27:1022–32. doi: 10.1038/sj.jcbfm.9600413. [DOI] [PubMed] [Google Scholar]
  103. Porcu M, Chiarugi A. The emerging therapeutic potential of sirtuin-interacting drugs: from cell death to lifespan extension. Trends Pharmacol Sci. 2005;26:94–103. doi: 10.1016/j.tips.2004.12.009. [DOI] [PubMed] [Google Scholar]
  104. Quinn L. Type 2 diabetes: epidemiology, pathophysiology, and diagnosis. Nurs Clin North Am. 2001;36:175–92. v. [PubMed] [Google Scholar]
  105. Ramnanan CJ, Groom AG, Storey KB. Akt and its downstream targets play key roles in mediating dormancy in land snails. Comp Biochem Physiol B Biochem Mol Biol. 2007 doi: 10.1016/j.cbpb.2007.06.002. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  106. Real PJ, Benito A, Cuevas J, Berciano MT, de Juan A, Coffer P, Gomez-Roman J, Lafarga M, Lopez-Vega JM, Fernandez-Luna JL. Blockade of epidermal growth factor receptors chemosensitizes breast cancer cells through up-regulation of Bnip3L. Cancer Res. 2005;65:8151–7. doi: 10.1158/0008-5472.CAN-05-1134. [DOI] [PubMed] [Google Scholar]
  107. Reddy S, Bibby NJ, Wu D, Swinney C, Barrow G, Elliott RB. A combined casein-free-nicotinamide diet prevents diabetes in the NOD mouse with minimum insulitis. Diabetes Res Clin Pract. 1995;29:83–92. doi: 10.1016/0168-8227(95)01109-9. [DOI] [PubMed] [Google Scholar]
  108. Relling DP, Esberg LB, Fang CX, Johnson WT, Murphy EJ, Carlson EC, Saari JT, Ren J. High-fat diet-induced juvenile obesity leads to cardiomyocyte dysfunction and upregulation of Foxo3a transcription factor independent of lipotoxicity and apoptosis. J Hypertens. 2006;24:549–61. doi: 10.1097/01.hjh.0000203846.34314.94. [DOI] [PubMed] [Google Scholar]
  109. Rice L, Handayani R, Cui Y, Medrano T, Samedi V, Baker H, Szabo NJ, Rosser CJ, Goodison S, Shiverick KT. Soy isoflavones exert differential effects on androgen responsive genes in LNCaP human prostate cancer cells. J Nutr. 2007;137:964–72. doi: 10.1093/jn/137.4.964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Rytomaa M, Lehmann K, Downward J. Matrix detachment induces caspase-dependent cytochrome c release from mitochondria: inhibition by PKB/Akt but not Raf signalling. Oncogene. 2000;19:4461–8. doi: 10.1038/sj.onc.1203805. [DOI] [PubMed] [Google Scholar]
  111. Sadanaga-Akiyoshi F, Yao H, Tanuma S, Nakahara T, Hong JS, Ibayashi S, Uchimura H, Fujishima M. Nicotinamide attenuates focal ischemic brain injury in rats: with special reference to changes in nicotinamide and NAD+ levels in ischemic core and penumbra. Neurochem Res. 2003;28:1227–34. doi: 10.1023/a:1024236614015. [DOI] [PubMed] [Google Scholar]
  112. Sandri M, Sandri C, Gilbert A, Skurk C, Calabria E, Picard A, Walsh K, Schiaffino S, Lecker SH, Goldberg AL. Foxo transcription factors induce the atrophy-related ubiquitin ligase atrogin-1 and cause skeletal muscle atrophy. Cell. 2004;117:399–412. doi: 10.1016/s0092-8674(04)00400-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Sela U, Dayan M, Hershkoviz R, Cahalon L, Lider O, Mozes E. The negative regulators Foxj1 and Foxo3a are up-regulated by a peptide that inhibits systemic lupus erythematosus-associated T cell responses. Eur J Immunol. 2006;36:2971–80. doi: 10.1002/eji.200636137. [DOI] [PubMed] [Google Scholar]
  114. Tanaka M, Kirito K, Kashii Y, Uchida M, Watanabe T, Endo H, Endoh T, Sawada K, Ozawa K, Komatsu N. Forkhead family transcription factor FKHRL1 is expressed in human megakaryocytes. Regulation of cell cycling as a downstream molecule of thrombopoietin signaling. J Biol Chem. 2001;276:15082–9. doi: 10.1074/jbc.M007958200. [DOI] [PubMed] [Google Scholar]
  115. Tothova Z, Kollipara R, Huntly BJ, Lee BH, Castrillon DH, Cullen DE, McDowell EP, Lazo-Kallanian S, Williams IR, Sears C, Armstrong SA, Passegue E, DePinho RA, Gilliland DG. FoxOs are critical mediators of hematopoietic stem cell resistance to physiologic oxidative stress. Cell. 2007;128:325–39. doi: 10.1016/j.cell.2007.01.003. [DOI] [PubMed] [Google Scholar]
  116. Tran H, Brunet A, Grenier JM, Datta SR, Fornace AJ, Jr, DiStefano PS, Chiang LW, Greenberg ME. DNA repair pathway stimulated by the forkhead transcription factor FOXO3a through the Gadd45 protein. Science. 2002;296:530–4. doi: 10.1126/science.1068712. [DOI] [PubMed] [Google Scholar]
  117. Van Der Heide LP, Hoekman MF, Smidt MP. The ins and outs of FoxO shuttling: mechanisms of FoxO translocation and transcriptional regulation. Biochem J. 2004;380:297–309. doi: 10.1042/BJ20040167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Vaziri H, Dessain SK, Ng Eaton E, Imai SI, Frye RA, Pandita TK, Guarente L, Weinberg RA. hSIR2(SIRT1) functions as an NAD-dependent p53 deacetylase. Cell. 2001;107:149–59. doi: 10.1016/s0092-8674(01)00527-x. [DOI] [PubMed] [Google Scholar]
  119. Wang F, Nguyen M, Qin FX, Tong Q. SIRT2 deacetylates FOXO3a in response to oxidative stress and caloric restriction. Aging Cell. 2007;6:505–14. doi: 10.1111/j.1474-9726.2007.00304.x. [DOI] [PubMed] [Google Scholar]
  120. Wick A, Wick W, Waltenberger J, Weller M, Dichgans J, Schulz JB. Neuroprotection by hypoxic preconditioning requires sequential activation of vascular endothelial growth factor receptor and Akt. J Neurosci. 2002;22:6401–7. doi: 10.1523/JNEUROSCI.22-15-06401.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Wijchers PJ, Burbach JP, Smidt MP. In control of biology: of mice, men and Foxes. Biochem J. 2006;397:233–46. doi: 10.1042/BJ20060387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Wild S, Roglic G, Green A, Sicree R, King H. Global prevalence of diabetes: estimates for the year 2000 and projections for 2030. Diabetes Care. 2004;27:1047–53. doi: 10.2337/diacare.27.5.1047. [DOI] [PubMed] [Google Scholar]
  123. Won CK, Ji HH, Koh PO. Estradiol prevents the focal cerebral ischemic injury-induced decrease of forkhead transcription factors phosphorylation. Neurosci Lett. 2006;398:39–43. doi: 10.1016/j.neulet.2005.12.060. [DOI] [PubMed] [Google Scholar]
  124. Yamaguchi H, Wang HG. The protein kinase PKB/Akt regulates cell survival and apoptosis by inhibiting Bax conformational change. Oncogene. 2001;20:7779–86. doi: 10.1038/sj.onc.1204984. [DOI] [PubMed] [Google Scholar]
  125. You H, Jang Y, You-Ten AI, Okada H, Liepa J, Wakeham A, Zaugg K, Mak TW. p53-dependent inhibition of FKHRL1 in response to DNA damage through protein kinase SGK1. Proc Natl Acad Sci U S A. 2004;101:14057–62. doi: 10.1073/pnas.0406286101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Zheng WH, Kar S, Quirion R. FKHRL1 and its homologs are new targets of nerve growth factor Trk receptor signaling. J Neurochem. 2002;80:1049–61. doi: 10.1046/j.0022-3042.2002.00783.x. [DOI] [PubMed] [Google Scholar]
  127. Zheng WH, Kar S, Quirion R. Insulin-like growth factor-1-induced phosphorylation of the forkhead family transcription factor FKHRL1 is mediated by Akt kinase in PC12 cells. J Biol Chem. 2000;275:39152–8. doi: 10.1074/jbc.M002417200. [DOI] [PubMed] [Google Scholar]
  128. Zhou ZQ, Wang T, Pan LM, Huang RH, Shi FX. FoxO4 is the main forkhead transcriptional factor localized in the gastrointestinal tracts of pigs. J Zhejiang Univ Sci. 2007;8:39–44. doi: 10.1631/jzus.2007.B0039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Zhu W, Bijur GN, Styles NA, Li X. Regulation of FOXO3a by brain-derived neurotrophic factor in differentiated human SH-SY5Y neuroblastoma cells. Brain Res Mol Brain Res. 2004;126:45–56. doi: 10.1016/j.molbrainres.2004.03.019. [DOI] [PubMed] [Google Scholar]

RESOURCES