Synopsis
The successful treatment for multiple disease entities can rest heavily upon the ability to elucidate the intricate relationships that govern cellular proliferation, metabolism, survival, and inflammation. Here we discuss the therapeutic potential of the mammalian forkhead transcription factors predominantly in the O class, FoxO1, FoxO3, FoxO4, and FoxO6, which play a significant role during normal cellular function as well as during progressive disease. These transcription factors are integrated with several signal transduction pathways, such as Wnt proteins, that can regulate a broad array of cellular process that include stem cell proliferation, aging, and malignancy. FoxO transcription factors are attractive considerations for strategies directed against human cancer in light of their pro-apoptotic effects and ability to lead to cell cycle arrest. Yet, FoxO proteins can be associated with infertility, cellular degeneration, and unchecked cellular proliferation. As our knowledge continues to develop for this novel family of proteins, potential clinical applications for the FoxO family should heighten our ability to limit disease progression without clinical compromise.
Keywords: cancer, diabetes, immune system, oxidative stress, stem cells
1. Origin and structure of FoxO transcription factors
More than 100 forkhead genes and 19 human subgroups that extend from FOXA to FOXS are now known to exist since the initial discovery of the fly Drosophila melanogaster gene fork head 1,2. A current nomenclature has replaced prior terms, such as forkhead in rhabdomyosarcoma (FKHR), the Drosophila gene fork head (fkh), and Forkhead RElated ACtivator (FREAC)-1 and -2. 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 designation of “Fox”, then a subclass or subgroup “Letter” is provided, and finally the member “Number” is listed. In relation to the nomenclature for human Fox proteins, all letters are capitalized, otherwise only the initial letter is listed as uppercase for the mouse, and for all other chordates the initial and subclass letters are in uppercase.
Of the mammalian forkhead transcription factors in the O class, FoxO1, FoxO3, FoxO4, and FoxO6 proteins can play a significant role during normal cellular function as well as during progressive disease. The most recently cloned member is FoxO6, but progressive interest in FoxO1, FoxO3, and FoxO4 has shown that these transcription factors can promote cell proliferation as well as cell death 3. For example, FoxO proteins are homologous to the transcription factor DAuer Formation-16 (DAF-16) in the worm Caenorhabditis elegans that can determine metabolic insulin signaling and lead to lifespan extension 2,4. It is believed that FoxO proteins can influence cellular function in multiple species, since metabolic signaling with FoxO proteins is conserved among Caenorhabditis elegans, Drosophila melanogaster, and mammals.
The forkhead box (FOX) family of genes have a conserved forkhead domain (the “forkhead box”) described as a “winged helix” as a result of the butterfly-like appearance on X-ray crystallography 5 and nuclear magnetic resonance 6. The forkhead domain in FoxO proteins consists of three α-helices, three β-sheets, and two loops that are referred to as the wings 2,3, but it should be noted that not all winged helix domains are considered to be Fox proteins 7. High sequence homology is present in the α-helices and β-sheets with variations described in either absent β-sheets and loops or additional α-helices. FoxO proteins bind DNA through the FoxO-recognized element with the consensus sequence T/C-G/A-A-A-A-C-A-A in the C-terminal basic region of the forkhead DNA binding domain8,9. The forkhead proteins either activate or repress target gene expression through fourteen protein-DNA contacts with the primary recognition site located at α-helix H3 5. Although both the first and second loops make contact with DNA, it is the second loop that can enhance the specificity and stability of the binding. It is believed that post-translational modification of FoxO proteins, such as phosphorylation or acetylation that block FoxO activity, alter the binding of the C-terminal basic region to DNA to prevent transcriptional activity 10. Yet, the mechanisms that lead to DNA binding with FoxO proteins are not completely defined and may depend upon several factors, such as variations in the N-terminal region of the recognition helix, changes in electrostatic distribution, and the ability of FoxO proteins to be shuttled to the cell nucleus that can be controlled by the C-terminal region of the forkhead domain 11,12.
2. Tissue expression of FoxO proteins
FoxO proteins are expressed throughout the body and are found in the ovary, prostate, skeletal muscle, brain, heart, lung, liver, pancreas, spleen, thymus, and testis 3,13-17. Initially, FOXO1, termed forkhead in rhabdomyosarcoma (FKHR), and FOXO3a, also known as FKHRL1 (forkhead in rhabdomyosarcoma like protein 1), and their genes were identified through chromosomal translocations in alveolar rhabdomyosarcoma tumors 2. The acute leukemia fusion gene located in chromosome X (AFX), also known as the FOXO4 gene, was described as a gene that fused to MLL transcription factor as a result of the t(X; 11) chromosomal translocation in acute lymphoblastic leukemia 18. A fusion between FOXO2 and MLL also occurs in some cases of acute myeloid leukemia that also is believed to be identical to FOXO3a 19.
Interestingly, FoxO proteins are not equally expressed in all tissues, suggesting that individual FoxO proteins may have specificity in regards to cellular function. For example, Foxo6 expression is found throughout several regions of the brain that play a significant role in cognitive function and emotion, such as the hippocampus, the amygdala, and the nucleus accumbens 16. However, Foxo1 may have a greater role in motor pathways with some memory formation, since its expression is primarily in the striatum and sub-regions of the hippocampus 16. On the other hand, Foxo3 is more diffusely represented in the hippocampus, cortex, and cerebellum, suggesting a complementary role for this FoxO protein to control cognitive and motor function. Furthermore, in mouse embryos and adults, mRNA expression of Foxo1, Foxo3a, and Foxo4 have a significant presence in muscle, adipose tissue, and liver with Foxo3a displaying a greater distribution in the heart, brain, and kidney 14.
3. FoxO proteins and cellular signaling
3.2 Post-translational control of FoxO proteins
Post-translational modification of FoxO proteins is critical for the regulation of these transcription factors and employs the biochemical pathways associated with phosphorylation, acetylation, and ubiquitylation 2,4,20,21. In regards to the inhibition of FoxO protein activity, the serine-threonine kinase protein kinase B (Akt) is a primary mediator of phosphorylation of FoxO1, FoxO3a, and FoxO4 2,22. Activation of Akt is usually cytoprotective, such as during free radical exposure 23,24, hyperglycemia 25, hypoxia 26,27, β-amyloid toxicity 28-30, and oxidative stress 31-33. Akt can prevent cellular apoptosis through the phosphorylation of FoxO proteins 34. Post-translational phosphorylation of FoxO proteins will maintain FoxO transcription factors in the cytoplasm by association with 14-3-3 proteins and prevent the transcription of pro-apoptotic target genes 35,36. An exception in regards to the subcellular trafficking of FoxO proteins involves FoxO6. This FoxO protein usually resides in the nucleus of cells and is phosphorylated by Akt in the nucleus. FoxO6 does not contain a conserved C-terminal Akt motif which limits nuclear shuttling of this protein. Yet, FoxO6 transcriptional activity can be blocked by growth factors independent of shuttling to the cytosol through a FoxO6 N-terminal Akt site 37.
Modulation of Akt activity during oxidative stress can control the apoptotic pathways of the caspase family that may offer an alternative mechanism to regulate FoxO proteins. Caspases are a family of cysteine proteases that are synthesized as inactive zymogens which are proteolytically cleaved into subunits at the onset of apoptosis 38,39. The caspases 1 and 3 have each been linked to the apoptotic pathways of genomic DNA cleavage and cellular membrane PS exposure 23,40-42. These caspases, in addition to caspase 8 and 9, are also tied to the direct activation and proliferation of microglia 23,32,33. Furthermore, caspase 9 is activated through a process that involves the cytochrome c -apoptotic protease-activating factor-1 (Apaf-1) complex 43,44. Caspase pathways may be tied to the forkhead transcription factor FoxO3a since increased activity of FoxO3a can result in cytochrome c release and caspase-induced apoptotic death 35,45-47. Pathways that can inhibit caspase 3 activity appear to offer a unique regulatory mechanism for FoxO3a that blocks the proteolytic degradation of inactive phosphorylated FoxO3a to prevent apoptotic cell injury during oxidative stress 35,45,46.
In addition to phosphorylation of forkhead transcription factors, post-translational modification of FoxO proteins also relies upon biochemical pathways associated with ubiquitylation and acetylation 48,49. Akt phosphorylation of FoxO proteins not only retains these transcription factors in the cytoplasm, but also leads to ubiquitination and degradation through the 26S proteasome 4,49. In the absence of Akt, IκB kinase (IKK) also can directly phosphorylate and block the activity of FoxO proteins, such as FoxO3a. This leads to the proteolysis of FoxO3a via the Ub-dependent proteasome pathway 50. The serum- and glucocorticoid-inducible protein kinase (Sgk), a member of a family of kinases termed AGC (protein kinase A/protein kinase G/protein kinase C) kinases which includes Akt, also can phosphorylate and retain FoxO3a in the cytoplasm 51. Knowledge that Sgk and Akt can phosphorylate FoxO3a at different sites may offer new opportunities to more effectively prevent apoptotic cell injury that may be mediated by FoxO3a activity. Yet, phosphorylation of FoxO proteins does not always lead to negative regulation. Interestingly, c-Jun N-terminal kinase (JNK) phosphorylates 14-3-3 protein leading to the nuclear localization of FoxO proteins, such as FoxO3a 52, suggesting that JNK promotes apoptosis through increased FoxO protein activity. The protein kinase mammalian sterile 20-like kinase-1 also can phosphorylate FoxO proteins directly and lead to their activation 53. The ability of sterile 20-like kinase-1 to activate FoxO proteins may be linked to JNK, since sterile 20-like kinase-1 can increase JNK activation 54. FoxO proteins also are acetylated by histone acetyltransferases that include p300, the CREB-binding protein (CBP), and the CBP-associated factor and are deacetylated by histone deacetylases, such as SIRT1 2,4,20,21. Acetylation of FoxO proteins provides another avenue for the control of these proteins. Once acetylated such as by CBP, FoxO proteins may translocate to the cell nucleus but have diminished activity since acetylation of lysine residues on FoxO proteins has been shown to limit the ability of FoxO proteins to bind to DNA 55. In addition, acetylation can increase phosphorylation of FoxO proteins by Akt 55.
3.2 FoxO proteins, oxidative stress, and apoptosis
In many diseases, cellular survival and cellular longevity are intimately dependent upon exposure to oxidative stress and the induction of apoptotic pathways. Oxidative stress is a result of the release of reactive oxygen species (ROS) that include superoxide free radicals, hydrogen peroxide, singlet oxygen, nitric oxide, and peroxynitrite 56. Oxygen free radicals and mitochondrial DNA mutations have become associated with tissue injury, aging, and accumulated toxicity for an organism 1,57,58. Most ROS are produced at low levels during normal physiological conditions and are scavenged by endogenous antioxidant systems that include superoxide dismutase, glutathione peroxidase, catalase, and small molecule substances, such as vitamins C, E, D3 59 and nicotinamide, the amide form of niacin or vitamin B3 60-63.
Genes involved in apoptosis have recently been found to be involved in processes of cell replication and transcription, suggesting that apoptotic pathways may be involved in multiple cellular events that do not necessarily lead to cell death 64. However, during conditions that inadequately control the production of ROS and lead to oxidative stress, cell apoptotic injury can ensue and contribute to disease pathology in disorders such as diabetes, Alzheimer's disease, and cardiovascular injury 1,57,65,66. Apoptotic cell death is a dynamic process that entails both early and late events. Membrane phosphatidylserine (PS) externalization is an early event during cell apoptosis 67,68 that assists microglia to target cells for phagocytosis 32,33,62,69,70. This process occurs with the expression of the phosphatidylserine receptor (PSR) on microglia during oxidative stress 71-73, since blockade of PSR function in microglia prevents the activation of microglia 23,33. As an example, externalization of membrane PS residues occur in neurons during anoxia 74-76, nitric oxide exposure 77,78, and during the administration of agents that induce the production of ROS, such as 6-hydroxydopamine 79. The cleavage of genomic DNA into fragments 80-82 is considered to be a later event during apoptotic injury 41. Several enzymes responsible for DNA degradation have been identified and include the acidic, cation independent endonuclease (DNase II), cyclophilins, and the 97 kDa magnesium - dependent endonuclease 1,56. Three separate endonuclease activities are present in neurons that include a constitutive acidic cation-independent endonuclease, a constitutive calcium/magnesium-dependent endonuclease, and an inducible magnesium dependent endonuclease 83,84.
Cell culture and animal studies that examine the effects of oxidative stress illustrate that FoxO proteins are closely tied to apoptotic injury (Table 1). It appears that FoxO1 and FoxO3a must be present for oxidative stress to result in apoptotic cell injury 85 and that the conditional deletion of FoxO1, FoxO3a, and FoxO4 can lead to an increase in ROS 86. Furthermore, FoxO3a in conjunction with JNK has been shown to modulate an apoptotic ligand activating a Fas-mediated death pathway in cultured motoneurons 87, to lead to apoptosis through tumor-necrosis-factor-related apoptosis-induced ligand (TRAIL) and BH3-only proteins Noxa and Bim in neuroblastoma cells 47, and to promote pro-apoptotic activity of p53 88. In addition, loss of FoxO protein activity can result in cytoprotection. Protein inhibition or gene knockdown of FoxO proteins, such as FoxO1 or FoxO3a, increases neuronal survival through NAD+ precursors 46, leads to stroke reduction by estradiol 89, mediates the protective effects of metabotropic glutamate receptors 45, and provides trophic factor protection with erythropoietin (EPO) 35 and neurotrophins 90. This cytoprotection, such as with EPO, involves both inhibition of nuclear shuttling (Figure 1) as well as phosphorylation by Akt 35.
Table 1. FoxO signaling and related pathways in disease.
Pathological modalities | FoxO proteins expression and post-translational effects | Selected references |
---|---|---|
Alzheimer's disease and aging | Amyloid leads to phopshorylarion of FoxO1 and FoxO3a | 95 |
β-catenin can modulate FoxO transcriptional activity | 96 | |
Wnt signaling relies upon β-catenin to prevent neuronal amyloid toxicity | 94 | |
Benefits of physical cardiovascular activity may be tied to FoxO protein expression | 148 | |
Prevention of FoxO protein translational activity can reduce cell loss during oxidative stress and neurodegeneration | 35, 45, 89-90 | |
Cancer | FoxO proteins are attractive strategies against cancer cell proliferation through cell cycle regulation | 161-164 |
Absence of functional FoxO proteins leads to breast, thymic, and prostate tumors | 17, 167-171 | |
FoxO3a transcriptional activity may be required to prevent B-chronic lymphocytic leukemia and chronic myelogenous leukemia | 162. 173 | |
FoxO proteins may have a complex role in cancers during modulation of androgen receptor expression and chemotherapy sensitivity | 175-176 | |
Diabetes mellitus | FoxO1 and FoxO3a associated with increased HbA1c and increased mortality | 128 |
FoxO1 and FoxO3a linked to increased risk of diabetes and stroke | 129 | |
In some animal models of diabetes, loss of FoxO3a activity may contribute to disease complications | 131 | |
NAD+ precursors protect β-cell function and reduce clinical disease during diabetes that may require removal of FoxO3a activity | 38, 46, 62, 111 | |
Immune system dysfunction | Post-translational phosphorylation of FoxO proteins may promote inflammatory cell activation during rheumatoid arthritis and osteoarthritis | 154-156 |
Animal studies suggest that Fox proteins are required to modulate T-cell activity, neutrophil activity, NF-κB activity, and interferon-gamma secretion | 157-159 | |
Premature ovarian failure | FoxO3a can prevent oocyte growth and follicular development | 13, 102 |
In some women who experience premature ovarian failure, mutations in FOXO1a and FOXO3a have been reported | 103 |
Nomenclature: “FoxO” refers to cell culture, tissue, or animal studies; “FOXO” refers to human studies
3.3 FoxO proteins and integration with novel cellular pathways
FoxO proteins are integrated with multiple signal transduction pathways which regulate cellular apoptosis and longevity during oxidative stress. One pathway in particular involves proteins derived from the Drosophila Wingless (Wg) and the mouse Int-1 genes. The Wnt proteins are secreted cysteine-rich glycosylated proteins that can control cell proliferation, differentiation, survival, and tumorigenesis 91,92. More than eighty target genes of Wnt signaling pathways have been demonstrated in human, mouse, Drosophila, Xenopus, and zebrafish. These genes are present in several cellular populations, such as neurons, cardiomyocytes, endothelial cells, cancer cells, and pre-adipocytes 61. At least nineteen of twenty-four Wnt genes that express Wnt proteins have been identified in the human 91-93.
Wnt proteins are generally divided into functional classes based on their ability to induce a secondary body axis in Xenopus embryos and to activate certain signaling cascades that consist of the Wnt1 class and the Wnt5a class 61,92. These involve intracellular signaling pathways that are critical for Wnt signal transduction. One Wnt pathway involves intracellular calcium release and is termed the non-canonical or Wnt/calcium pathway consisting primarily of Wnt4, Wnt5a, and Wnt11. The non-canonical system functions through non-β-catenin-dependent pathways and also includes the planar cell polarity (PCP) pathway or the Wnt-calcium-dependent pathways 91-93. A second pathway controls target gene transcription through β-catenin, generally referred to as the canonical pathway that involves Wnt1, Wnt3a, and Wnt8. It is the β-catenin pathway that appears to tie FoxO proteins and Wnt signaling together. For example, in relation to Alzheimer's disease, amyloid is toxic in cell culture 28,94 and is associated with the phosphorylation of FoxO1 and FoxO3a that can be blocked with ROS scavengers 95 (Table 1). Interestingly, a common denominator in the pathways linked to amyloid toxicity involves Wnt signaling through β-catenin. β-catenin may increase FoxO transcriptional activity and competitively limit β-catenin interaction with members of the lymphoid enhancer factor/T cell factor family 96 and β-catenin also has been demonstrated to be necessary for protection against amyloid toxicity in neuronal cells 94.
In addition to shared signal transduction pathways between Wnt and FoxO proteins that involve β-catenin, Akt is intimately tied to both Wnt and FoxO signaling. As previously described, Akt phosphorylates and blocks the activity of the FoxO proteins FoxO1, FoxO3a, and FoxO4 2,22. In relation to Wnt signaling, Wnt relies upon Akt for cell differentiation and cytoprotection. For example, neuronal cell differentiation that is dependent upon Wnt signaling and trophic factor induction is blocked during the repression of Akt activity 97. Furthermore, Wnt differentiation of cardiomyocytes does not proceed without Akt activation 98 while reduction in tissue injury during pressure overload cardiac hypertrophy 99 and the benefits of cardiac ischemic preconditioning also appear to rely upon Akt 100. In addition, Wnt over-expression can independently increase the phosphorylation and the activation of Akt to promote neuronal protection and control microglial activation 94.
4. FoxO, stem cells, and cellular development
As our knowledge of FoxO proteins continues to grow, new work provides evidence for the role of FoxO proteins in vascular system development, fertility, and progenitor cell differentiation. Studies have shown that Foxo3a -/- and Foxo4 -/- mice develop without incidence and are indistinguishable from control littermates, but mice singly deficient in Foxo1 die by embryonic day eleven and lack development of the vasculature 101. Furthermore, Foxo3a -/- mice become infertile and experience follicular activation to the extent that ovarian follicles are depleted of oocytes 13. Other work using a mouse model of Foxo3a over-expression in oocytes further suggests that Foxo3a retards oocyte growth and follicular development and leads to anovulation and luteinization of unruptured follicles 102. The studies with Foxo3a null mice may suggest a role for FoxO3a in addition to FoxO1 in relation to oocyte and follicular development. For example, in a small percentage of women who suffer from premature ovarian failure, mutations in FOXO3a and FOXO1a have been observed 103 (Table 1). If one examines hematopoietic stem cell development, studies suggest that FoxO3a alone may play a role in maintaining hematopoietic stem cells, since these cells are significantly decreased in aged Foxo3-/- mice compared to the littermate controls 104. Yet, other work indicates that the combined loss of Foxo1, Foxo3a, and Foxo4 in mice is required to lead to defective repopulation of hematopoietic stem cells with resultant apoptosis 86.
A number of cellular agents, such as the growth factor and cytokine EPO 36,105, also may determine whether FoxO proteins function in concert or independently to progenitor cell growth. In cell culture and animal studies, EPO is cytoprotective in neuronal and vascular cells and can stimulate postnatal neovascularization by increasing endothelial progenitor cell mobilization from the bone marrow 44,106,107. Interestingly, the ability of EPO to foster eythroid progenitor cell development is dependent upon the inhibition of FoxO3a activity 36,108, but also may require regulation of specific gene expression through an EPO-FoxO3a association to promote erythropoiesis in cultured cells 109. In addition, rat enteric nervous system precursor development that occurs in the presence of the growth factor glial cell line-derived neurotrophic factor appears to require the inactivation of FoxO1 and FoxO3a 110.
5. FoxO proteins, diabetes, and cellular metabolism
Both clinical and experimental studies exemplify the role of FoxO proteins in cellular metabolism and disorders such as diabetes mellitus (DM). DM is a significant health concern for both young and older populations 111,112. Approximately 16 million individuals in the United States and more than 165 million individuals worldwide suffer from DM. By the year 2030, it is predicted that more than 360 million individuals will be afflicted with DM and its debilitating conditions 113. Type 2 DM 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 114. Type 1 insulin-dependent DM is present in 5-10 percent of all diabetics 112, but is increasing in adolescent minority groups 115. Furthermore, the incidence of undiagnosed diabetes, impaired glucose tolerance, and fluctuations in serum glucose in the young raises additional concerns 116.
Patients with DM can develop severe neurological and vascular disease 117,118 that can lead to an increased risk for cognitive decline 1,38,119. Interestingly, the development of insulin resistance and the complications of DM in the nervous and vascular systems can be the result of cellular oxidative stress 111,112. Hyperglycemia can lead to increased production of ROS in endothelial cells, liver and pancreatic β-cells 120-123. Recent clinical correlates support these experimental studies to show that elevated levels of ceruloplasmin are suggestive of increased ROS 124. Furthermore, acute glucose swings in addition to chronic hyperglycemia can trigger oxidative stress mechanisms, illustrating the importance for therapeutic interventions during acute and sustained hyperglycemic episodes 125.
Both clinical and experimental studies exemplify the role of FoxO proteins during cellular metabolism and DM. FoxO proteins can stimulate the insulin-like growth factor binding protein-1 (IGFBP1) promoter by binding to the insulin-responsive sequence (IRS) 126. Insulin and insulin-like growth factor-1 (IGF-1) can suppress FoxO protein activity through activation of Akt 126,127. In a clinical study of 734 individuals, the c. −343–1582C>T polymorphism of FOXO3a displayed a significant association with body mass index such that the highest body mass index was present in individuals homozygous for this allele 128. Analysis of the genetic variance in FOXO1a and FOXO3a on metabolic profiles, age-related diseases, fertility, fecundity, and mortality illustrated higher HbA1c levels and increased mortality risk associated with specific haplotypes of FOXO1a. There also was an increased risk of stroke in two haplotypes of FOXO3a block-A, suggesting an association with cerebral oxidative stress disorders such as diabetes and stroke with FOXO1a and FOXO3a 129 (Table 1).
In some animal and cell culture studies, modulation of forkhead transcription factors, such as FoxO3a, may counteract the detrimental effects of high serum glucose levels. For example, interferon-gamma driven expression of tryptophan catabolism by cytotoxic T lymphocyte antigen 4 may activate Foxo3a to protect dendritic cells from injury in nonobese diabetic mice 130. Additional investigations have associated diabetic nephropathy to post-translational changes in FoxO3a by demonstrating that inhibitory phosphorylation of FoxO3a increases in rat and mouse renal cortical tissues two weeks after the induction of diabetes by streptozotocin 131, suggesting that the loss of FoxO3a activity can lead to renal disease. The human immunodeficiency virus (HIV) -1 accessory protein Vpr also has been reported in human hepatoma cells to contribute to insulin resistance by interfering with FoxO3a signaling with protein 14-3-3 132. Yet, other work suggests that inactivation of FoxO proteins may foster cytoprotection. For example, enteric neurons can be protected from hyperglycemia by glial cell line-derived neurotrophic factor that can affect Akt signaling and prevent FoxO3a activation 25.
The preservation of cellular energy reserves and mitochondrial function also may be a critical factor for FoxO proteins to regulate cellular metabolism during DM. Chronic exposure to elevated levels of free fatty acids can increase ROS production in cells and has been shown to lead to mitochondrial DNA damage and impaired pancreatic β-cell function 133. In patients with type 2 DM, skeletal muscle mitochondria have been described to be smaller than those in control subjects 134. In addition, a decrease in the levels of mitochondrial proteins and mitochondrial DNA in adipocytes has been correlated with the development of type 2 DM 135. Insulin resistance in the elderly also has been associated with elevation in fat accumulation and altered mitochondrial oxidative and phosphorylation activity 136,137. In caloric restricted mice that have decreased energy reserves, mRNA expression is progressively increased for Foxo1, Foxo3a, and Foxo4 over a two year course 15. This work is complementary to investigations in Drosophila and mammalian cells that show up-regulation of FoxO1 expression leads to increased insulin signaling to regulate cellular metabolism 138. Yet, other studies such as with Foxo1 have shown that over-expression of this transcription factor in skeletal muscles of mice can lead to reduced skeletal muscle mass and poor glycemic control 139, illustrating that activation of FoxO proteins may also impair cellular energy reserves. As a result, one potential agent to consider for the maintenance of cellular metabolism in DM is nicotinamide 61,140,141, an agent that also can inhibit FoxO protein activity 46. In patients with DM, oral nicotinamide protects β-cell function, prevents clinical disease in islet-cell antibody-positive first-degree relatives of type-1 DM, and can reduce HbA1c levels 38,62,111. It is of interest to note that nicotinamide may derive its protective capacity through two separate mechanisms of post-translational modification of FoxO3a. Nicotinamide not only can maintain phosphorylation of FoxO3a and inhibit its activity, but also can preserve the integrity of the FoxO3a protein to block FoxO3a proteolysis that can yield pro-apoptotic amino-terminal fragments 46 (Table 1).
6. FoxO proteins, cellular longevity, and immune system function
As an extension of the studies examining apoptotic cell injury, FoxO proteins also have been tied to cell longevity and aging as shown by early studies linking DAF-16 in Caenorhabditis elegans to increased longevity 38,142. However, the relationship between FoxO transcription factors and proteins that increased cellular lifespan has been met with controversy. SIRT1 is an NAD+-dependent deacetylase and the mammalian ortholog of the silent information regulator 2 (Sir2) protein associated with increased lifespan in yeast. Some studies suggest that stimulation of SIRT1 during starvation is dependent upon FoxO3a activity as well as p53 143. In contrast, other work has shown in cell culture that SIRT1 may repress the activity of FoxO1, FoxO3a, and FoxO4, suggesting that cellular longevity may benefit from reduction in FoxO protein generated apoptosis 144. Additional studies offer alternative views to illustrate that SIRT1 binds to FoxO proteins, such as FoxO4, to catalyze its deacetylation and enhance FoxO4 activity while acetylation of FoxO4 by cyclic-AMP responsive element binding (CREB)-binding protein serves to inhibit FoxO4 transcriptional activity 145,146.
FoxO proteins also have been linked to cell aging and senescence. In cultured human dermal fibroblasts, gene silencing of FoxO3a protein results in cell morphology consistent with cell senescence, cell population doubling times, and the generation of ROS, suggesting that FoxO protein activity may be required to extend cell longevity and limit oxidative stress 147. Additional work in animal models of aging demonstrates a reduction in SIRT1 in the heart, but no significant change in FoxO3a expression with advanced age. However, during exercise training, an up-regulation of FoxO3a and SIRT1 activity is observed in the heart 148, suggesting that the benefits of physical activity for the cardiovascular system may be associated with FoxO proteins. In addition, FoxO proteins may be protective during aging, since loss of FoxO3a activity in explanted vascular smooth muscle of aged animals may limit tissue antioxidant properties through decreased manganese-superoxide dismutase and lead to enhanced cell injury with aging 149. Extension of cellular lifespan that depends upon the prevention of cell senescence at least in primary human cultured vascular cells also may require the negative regulation of Akt to allow for the activation of FoxO3a 150.
Given that inflammatory cell modulation has a significant impact upon cellular apoptosis, FoxO proteins also function as critical components for modulation of immune cell function. The ability to regulate early apoptotic membrane PS externalization and subsequent inflammatory cell activity can ultimately impact upon cell survival and longevity since activated immune cells can lead to the phagocytic removal of both neurons and vascular cells 56,71. Inflammatory cells, such as macrophages or microglia, require the activation of intracellular cytoprotective pathways to proliferate and remove injured cells 72,151. Inflammatory cells can be beneficial and form a barrier for the removal of foreign microorganisms and promote tissue repair during cell injury 39,152. Yet, inflammatory cells also may lead to cellular damage through the generation of ROS and through the production of cytokines 41,71,153.
Clinical studies have demonstrated in synovial biopsy tissue of patients with rheumatoid arthritis and osteoarthritis the phosphorylation of FOXO1 and FOXO4 in macrophages and the phosphorylation of FOXO3a in T lymphocytes, suggesting that inhibitory post-translational phosphorylation of these FOXO family members may lead to inflammatory cell activation 154 (Table 1). Additional work has shown that FOXO1 gene transcript levels are down-regulated in peripheral blood mononuclear cell of patients with systemic lupus erythematosus and rheumatoid arthritis 155, illustrating a potential etiology through the loss of functional FOXO proteins for these disorders and possibly providing a biomarker of disease activity. Clinical work also suggests a relationship between the regulation of immune system activity and the induction of apoptotic pathways that are dependent upon FoxO proteins. FoxO proteins may work in concert with Fas signaling to clear activated T cells following a decrease in cytokine stimulation in patients with autoimmune lymphoproliferative syndromes, suggesting that specific FoxO proteins may be targeted for treatment of autoimmune disorders 156. In mice deficient for Foxo3a, lymphoproliferation, organ inflammation of the salivary glands, lung, and kidney, and increased activity of helper T cells results, supporting an important role for FoxO3a in preventing T cell hyperactivity 157. FoxO3a also appears to be necessary for neutrophil activity, since Foxo3a -/- mice are resistant to models of neutrophilic inflammation that involve immune complex-mediated inflammatory arthritis 158. Prevention of inflammatory activation and apoptosis in the nervous system such as in systemic lupus erythematosus in animal models may require the up-regulation of different Fox proteins, such as FoxJ1 and FoxO3a, that can block nuclear factor-κB (NF-κB) activation and interferon-gamma secretion 159 (Table 1). Animal studies using experimental autoimmune encephalomyelitis to mimic multiple sclerosis and myelin injury also have shown that osteopontin, a protein expressed in multiple sclerosis lesions, leads to the prolonged survival of myelin-reactive T cells and disease progression through a combination of events that involve FoxO3a inhibition, NF-κB activation, and modulating the expression of the pro-apoptotic proteins Bim, Bak, and Bax 160.
7. FoxO proteins and tumorigenesis
One of the most interesting therapeutic applications for FoxO proteins involves strategies directed against human cancer, since the pro-apoptotic effects of FoxO proteins and their ability to block cell cycle progression make these transcription factors almost ideal therapeutic targets to control tumorigenesis. For example, Foxo3a and Foxo4 can promote cell cycle arrest in mouse myoblastic cell lines through modulation of growth-arrest and DNA-damage-response protein 45 161. Treatment of chronic myelogenous leukemia cell lines with the Bcr-Abl tyrosine kinase inhibitor imatinib requires FoxO3a activation to antagonize cell proliferation and promote apoptotic cell death through increased TRAIL production 162. In addition, the transcription factor E2F-1 that controls the induction of the cell cycle has been reported in cell lines to increase the endogenous expression of FoxO1 and FoxO3a to lead to cell cycle arrest 163 (Table 1). However, the loss of FoxO3a activity in association with c-myc, p27, and NF-κB can result in cell cycle induction and malignant transformation of mouse cells in the presence of oncogene activation 164. It should be noted that FoxO protein inhibition of cell cycle progression may not consistently lead to apoptotic cell death. Some investigations suggest that during oxidative stress, FoxO3a activation in association with the Sir2 homolog SIRT1 can lead to cell cycle arrest, but not result in apoptosis 165.
Early clinical studies of breast cancer in relation to FOXO3a suggested that activation of FOXO3a was associated with lymph nodal metastasis and a poor prognosis 166. However, other studies reported that FOXO3a was confined to the cytoplasm of human tumor cells, inactivated by IKK, and that this inactivation of FOXO3a was associated with a poor prognosis in breast cancer 50, suggesting that FOXO3a sub-cellular localization and pathways that enhance its activity could be used not only as prognostic assays but also as therapeutic targets. In animal studies, somatic deletion in mice of Foxo1, Foxo3a, and Foxo4 results in the growth of thymic lymphomas and hemangiomas, further illustrating the potential of FoxO proteins to function as redundant repressors of tumor growth 167. Studies in breast cancer cells parallel this work and show that increased activity of FoxO3a in association with JNK in breast cancer cell lines 168 or in association with cyclin-dependent kinase inhibitor p27 in isolated human breast cancer cells can suppress breast cancer progression 169.
Studies with prostate cancer have shown that the tumor suppressor phosphatase and tensin homolog deleted on chromosome ten (PTEN) was mutated in almost eighty percent of tumors with the loss of FOXO1 and FOXO3a activity. In cell culture work, over-expression of FoxO1 and FoxO3a in prostrate tumor cell lines could result in apoptosis, suggesting that FoxO1 and FoxO3a were necessary for limiting prostate cell tumor growth 17. In further support of this work, inhibition of FoxO3a activity can result in enhanced prostate tumor cell growth 170 while agents that increase FoxO3a activity in both androgen sensitive and androgen insensitive prostate cell lines prevent prostate cancer cell progression 171 (Table 1).
In addition to neoplasms in breast and prostate, FoxO proteins also may represent a viable option to control tumor growth in tissues throughout the body. FoxO3a activation in colon carcinoma cell lines prevents tumor proliferation through Myc target genes that involve the Mad/Mxd family of transcriptional repressors 172. Other investigations illustrate that the loss of FoxO3a activity may participate in oncogenic transformation in B-chronic lymphocytic leukemia 173 and in the progression of chronic myelogenous leukemia cell lines 162. Furthermore, studies suggest that some proteins, such as the Kaposi's sarcoma-associated herpesvirus latent protein LANA2, may specifically block the transcriptional activity of FoxO3a to lead to tumor growth 174. Yet, FoxO proteins may have a complex role during tumor growth. FoxO3a is a positive regulator of androgen receptor expression and therefore may also assist with prostate cancer cell proliferation 175. In addition, loss of functional FoxO3a in human ovarian cancer cell lines can limit the sensitivity of ovarian cancer cells to chemotherapy 176, suggesting that FoxO proteins may be responsible for altered treatment outcomes in the presence of combined therapeutic approaches (Table 1).
8. Perspectives and future considerations: Targeting FoxO proteins
In light of the robust ability of FoxO proteins to oversee cell proliferation, cell metabolism, cell survival, and immune system function, these transcription factors may be enthusiastically considered for the treatment of a wide variety of disorders. For example, the known mutations in FoxO proteins that exist in several disease entities may provide novel insights for therapeutic strategies that can address a broad range of disorders. Although theses mutations are considered to represent one of multiple factors responsible for disorders such as premature ovarian failure, further analysis in larger populations of patients with premature ovarian failure could enhance our understanding of the role of FoxO proteins in disorders of human fertility. In relation to the immune system, recent work has suggested that FoxO proteins may function as biomarkers of disease activity and also offer a potential target for the treatment of autoimmune disorders. Furthermore, the ability of FoxO proteins to control cell cycle progression and promote apoptotic cell death suggests that FoxO transcription factors may be developed for new advances against tumorigenesis. As an example, triple mutant FoxO3a expression that cannot be inhibited through phosphorylation has been proposed as a potential therapeutic target against melanoma tumors 177.
Yet, it must be realized that the causal relationships between FoxO proteins and cellular metabolism, apoptotic injury, immune system function, and cancer are not well defined and that protocols to modulate FoxO proteins may yield a double-edge sword for both beneficial and detrimental clinical results. For example, the common pathways shared between Wnt and forkhead proteins may have another side that relates to tumorigenesis 91,178. Fox transcription factors can activate the Wnt/β-catenin pathway to increase extracellular proteoglycans and promote gastrointestinal cell proliferation 179. In the presence of Wnt deregulation and increased β-catenin activity, tumorigenesis may ensue 92. Deregulation in the Wnt pathway that promotes activation of β-catenin and cell survival also has been associated with the proliferation of medulloblastoma tumors 180. In addition, Wnt expression has been correlated with advanced gastric cancer stages and a poor prognosis 181 while experimental activation of the β-catenin pathway leads to the development of gastric tumors 182. Irrespective of the role of Wnt signaling, conditions also can exist that allow FoxO proteins to prevent cell cycle progression in cells without leading to apoptotic injury. Although this result may be considered beneficial to block degenerative disorders, in the setting of cancer, these results would severely limit clinical utility. Furthermore, FoxO transcription factors can foster apoptosis in prostate cancer cells, but also may contribute to androgen receptor expression and potentially diminish any clinical benefits.
As a result, it becomes essential to promote both basic as well as clinical research with FoxO proteins to comprehend the complex role played by these transcription factors. For example, FoxO proteins are expressed throughout the brain and particularly in sensitive cognitive and motor regions, but it is unclear how FoxO transcription factors may influence neuronal plasticity, angiogenesis, and immune system function that alter the progression of dementias or behavioral abnormalities. Investigations also are required to clarify both independent and shared signal transduction pathways of FoxO proteins, such as with Wnt signaling, to further understand the cellular processes that can influence gene transcription and intracellular trafficking of these pathways. In some scenarios, Fox proteins may prevent tumorigenesis during Wnt deregulation, but in other examples, Fox proteins may assist with β-catenin activation and potentially tumor cell proliferation. A number of currently unknown parameters may account for these discrepancies, such as the role of FoxO post-translational state, factors that control FoxO protein DNA binding, intracellular signaling and tissue specifics during oxidative stress, age of the organism, and cellular metabolic state. As new studies continue to unfold, the clinical applications for FoxO proteins should develop at a surprisingly fast pace to not only foster health maintenance, but also limit disease progression.
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.
Footnotes
Conflicts: There are no conflicts to disclose.
References
- 1.Chong ZZ, Li F, Maiese K. Oxidative stress in the brain: Novel cellular targets that govern survival during neurodegenerative disease. Prog Neurobiol. 2005;75(3):207–246. doi: 10.1016/j.pneurobio.2005.02.004. [DOI] [PubMed] [Google Scholar]
- 2.Maiese K, Chong ZZ, Shang YC. “Sly as a FOXO”: New paths with Forkhead signaling in the brain. Curr Neurovasc Res. 2007;4(4):295–302. doi: 10.2174/156720207782446306. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Maiese K, Chong ZZ, Shang YC. OutFOXOing disease and disability: the therapeutic potential of targeting FoxO proteins. Trends Mol Med. 2008;14(5):219–227. doi: 10.1016/j.molmed.2008.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Jagani Z, Singh A, Khosravi-Far R. FoxO tumor suppressors and BCR-ABL-induced leukemia: a matter of evasion of apoptosis. Biochim Biophys Acta. 2008;1785(1):63–84. doi: 10.1016/j.bbcan.2007.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.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(6436):412–420. doi: 10.1038/364412a0. [DOI] [PubMed] [Google Scholar]
- 6.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(17):6179–6187. doi: 10.1021/bi980031v. [DOI] [PubMed] [Google Scholar]
- 7.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;368(2):249–261. doi: 10.1016/j.virol.2007.06.040. [DOI] [PubMed] [Google Scholar]
- 8.Biggs WH, 3rd, Cavenee WK, Arden KC. Identification and characterization of members of the FKHR (FOX O) subclass of winged-helix transcription factors in the mouse. Mamm Genome. 2001;12(6):416–425. doi: 10.1007/s003350020002. [DOI] [PubMed] [Google Scholar]
- 9.Huang H, Tindall DJ. Dynamic FoxO transcription factors. J Cell Sci. 2007;120(Pt 15):2479–2487. doi: 10.1242/jcs.001222. [DOI] [PubMed] [Google Scholar]
- 10.Tsai KL, Sun YJ, Huang CY, Yang JY, Hung MC, Hsiao CD. Crystal structure of the human FOXO3a-DBD/DNA complex suggests the effects of post-translational modification. Nucleic Acids Res. 2007;35(20):6984–6994. doi: 10.1093/nar/gkm703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Weigelt J, Climent I, Dahlman-Wright K, Wikstrom M. Solution structure of the DNA binding domain of the human forkhead transcription factor AFX (FOXO4) Biochemistry. 2001;40(20):5861–5869. doi: 10.1021/bi001663w. [DOI] [PubMed] [Google Scholar]
- 12.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(Pt 2):297–309. doi: 10.1042/BJ20040167. [DOI] [PMC free article] [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(5630):215–218. doi: 10.1126/science.1086336. [DOI] [PubMed] [Google Scholar]
- 14.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(Pt 2):629–634. doi: 10.1042/0264-6021:3490629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Furuyama T, Yamashita H, Kitayama K, Higami Y, Shimokawa I, Mori N. Effects of aging and caloric restriction on the gene expression of Foxo1, 3, and 4 (FKHR, FKHRL1, and AFX) in the rat skeletal muscles. Microscopy research and technique. 2002;59(4):331–334. doi: 10.1002/jemt.10213. [DOI] [PubMed] [Google Scholar]
- 16.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(2):134–140. doi: 10.1016/j.modgep.2005.07.003. [DOI] [PubMed] [Google Scholar]
- 17.Modur V, Nagarajan R, Evers BM, Milbrandt J. FOXO proteins regulate tumor necrosis factor-related apoptosis inducing ligand expression. Implications for PTEN mutation in prostate cancer. J Biol Chem. 2002;277(49):47928–47937. doi: 10.1074/jbc.M207509200. [DOI] [PubMed] [Google Scholar]
- 18.Parry P, Wei Y, Evans G. Cloning and characterization of the t(X;11) breakpoint from a leukemic cell line identify a new member of the forkhead gene family. Genes Chromosomes Cancer. 1994;11(2):79–84. doi: 10.1002/gcc.2870110203. [DOI] [PubMed] [Google Scholar]
- 19.Hillion J, Le Coniat M, Jonveaux P, Berger R, Bernard OA. AF6q21, a novel partner of the MLL gene in t(6;11)(q21;q23), defines a forkhead transcriptional factor subfamily. Blood. 1997;90(9):3714–3719. [PubMed] [Google Scholar]
- 20.Myatt SS, Lam EW. The emerging roles of forkhead box (Fox) proteins in cancer. Nat Rev Cancer. 2007;7(11):847–859. doi: 10.1038/nrc2223. [DOI] [PubMed] [Google Scholar]
- 21.van der Horst A, Burgering BM. Stressing the role of FoxO proteins in lifespan and disease. Nat Rev Mol Cell Biol. 2007;8(6):440–450. doi: 10.1038/nrm2190. [DOI] [PubMed] [Google Scholar]
- 22.Chong ZZ, Li F, Maiese K. Activating Akt and the brain's resources to drive cellular survival and prevent inflammatory injury. Histol Histopathol. 2005;20(1):299–315. doi: 10.14670/hh-20.299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.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. 2003;138(6):1107–1118. doi: 10.1038/sj.bjp.0705161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.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(5):2037–2046. [PubMed] [Google Scholar]
- 25.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(2):344–356. doi: 10.1172/JCI26295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Chong ZZ, Kang JQ, Maiese K. Erythropoietin is a novel vascular protectant through activation of Akt1 and mitochondrial modulation of cysteine proteases. Circulation. 2002;106(23):2973–2979. doi: 10.1161/01.cir.0000039103.58920.1f. [DOI] [PubMed] [Google Scholar]
- 27.Zhang Y, Park TS, Gidday JM. Hypoxic preconditioning protects human brain endothelium from ischemic apoptosis by Akt-dependent survivin activation. Am J Physiol Heart Circ Physiol. 2007;292(6):H2573–2581. doi: 10.1152/ajpheart.01098.2006. [DOI] [PubMed] [Google Scholar]
- 28.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. 2005;2(5):387–399. doi: 10.2174/156720205774962683. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Du B, Ohmichi M, Takahashi K, Kawagoe J, Ohshima C, Igarashi H, Mori-Abe A, Saitoh M, Ohta T, Ohishi A, Doshida M, Tezuka N, Takahashi T, Kurachi H. Both estrogen and raloxifene protect against beta-amyloid-induced neurotoxicity in estrogen receptor alpha-transfected PC12 cells by activation of telomerase activity via Akt cascade. J Endocrinol. 2004;183(3):605–615. doi: 10.1677/joe.1.05775. [DOI] [PubMed] [Google Scholar]
- 30.Nakagami Y, Nishimura S, Murasugi T, Kubo T, Kaneko I, Meguro M, Marumoto S, Kogen H, Koyama K, Oda T. A novel compound RS-0466 reverses beta-amyloid-induced cytotoxicity through the Akt signaling pathway in vitro. Eur J Pharmacol. 2002;457(1):11–17. doi: 10.1016/s0014-2999(02)02657-2. [DOI] [PubMed] [Google Scholar]
- 31.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. 2004;296(2):196–207. doi: 10.1016/j.yexcr.2004.01.021. [DOI] [PubMed] [Google Scholar]
- 32.Kang JQ, Chong ZZ, Maiese K. Critical role for Akt1 in the modulation of apoptotic phosphatidylserine exposure and microglial activation. Mol Pharmacol. 2003;64(3):557–569. doi: 10.1124/mol.64.3.557. [DOI] [PubMed] [Google Scholar]
- 33.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. 2003;74(1):37–51. doi: 10.1002/jnr.10740. [DOI] [PubMed] [Google Scholar]
- 34.Maiese K, Chong ZZ, Hou J, Shang YC. Erythropoietin and oxidative stress. Curr Neurovasc Res. 2008;5(2):125–142. doi: 10.2174/156720208784310231. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.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. 2007;150(7):839–850. doi: 10.1038/sj.bjp.0707161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Maiese K, Li F, Chong ZZ. New avenues of exploration for erythropoietin. JAMA. 2005;293(1):90–95. doi: 10.1001/jama.293.1.90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.van der Heide LP, Jacobs FM, Burbach JP, Hoekman MF, Smidt MP. FoxO6 transcriptional activity is regulated by Thr26 and Ser184, independent of nucleo-cytoplasmic shuttling. Biochem J. 2005;391(Pt 3):623–629. doi: 10.1042/BJ20050525. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Li F, Chong ZZ, Maiese K. Cell Life Versus Cell Longevity: The Mysteries Surrounding the NAD(+) Precursor Nicotinamide. Curr Med Chem. 2006;13(8):883–895. doi: 10.2174/092986706776361058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Maiese K, Chong ZZ, Li F. Driving cellular plasticity and survival through the signal transduction pathways of metabotropic glutamate receptors. Curr Neurovasc Res. 2005;2(5):425–446. doi: 10.2174/156720205774962692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Chong ZZ, Kang JQ, Maiese K. Hematopoietic factor erythropoietin fosters neuroprotection through novel signal transduction cascades. J Cereb Blood Flow Metab. 2002;22(5):503–514. doi: 10.1097/00004647-200205000-00001. [DOI] [PubMed] [Google Scholar]
- 41.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. 2004;6(2):277–287. doi: 10.1089/152308604322899341. [DOI] [PubMed] [Google Scholar]
- 42.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. 2003;71(5):659–669. doi: 10.1002/jnr.10528. [DOI] [PubMed] [Google Scholar]
- 43.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. 2003;23(3):320–330. doi: 10.1097/01.WCB.0000050061.57184.AE. [DOI] [PubMed] [Google Scholar]
- 44.Maiese K, Chong ZZ, Li F, Shang YC. Erythropoietin: Elucidating new cellular targets that broaden therapeutic strategies. Prog Neurobiol. 2008;85:194–213. doi: 10.1016/j.pneurobio.2008.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.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(2):107–117. doi: 10.2174/156720206776875830. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.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. 2004;24(7):728–743. doi: 10.1097/01.WCB.0000122746.72175.0E. [DOI] [PubMed] [Google Scholar]
- 47.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(3):534–547. doi: 10.1038/sj.cdd.4402017. [DOI] [PubMed] [Google Scholar]
- 48.Matsuzaki H, Daitoku H, Hatta M, Tanaka K, Fukamizu A. Insulin-induced phosphorylation of FKHR (Foxo1) targets to proteasomal degradation. Proc Natl Acad Sci U S A. 2003;100(20):11285–11290. doi: 10.1073/pnas.1934283100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Plas DR, Thompson CB. Akt activation promotes degradation of tuberin and FOXO3a via the proteasome. J Biol Chem. 2003;278(14):12361–12366. doi: 10.1074/jbc.M213069200. [DOI] [PubMed] [Google Scholar]
- 50.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(2):225–237. doi: 10.1016/s0092-8674(04)00302-2. [DOI] [PubMed] [Google Scholar]
- 51.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(8):5871–5882. doi: 10.1074/jbc.M211649200. [DOI] [PubMed] [Google Scholar]
- 52.Sunayama J, Tsuruta F, Masuyama N, Gotoh Y. JNK antagonizes Akt-mediated survival signals by phosphorylating 14-3-3. J Cell Biol. 2005;170(2):295–304. doi: 10.1083/jcb.200409117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Lehtinen MK, Yuan Z, Boag PR, Yang Y, Villen J, Becker EB, DiBacco 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(5):987–1001. doi: 10.1016/j.cell.2006.03.046. [DOI] [PubMed] [Google Scholar]
- 54.Song JJ, Lee YJ. Differential cleavage of Mst1 by caspase-7/-3 is responsible for TRAIL-induced activation of the MAPK superfamily. Cell Signal. 2008;20(5):892–906. doi: 10.1016/j.cellsig.2008.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Matsuzaki H, Daitoku H, Hatta M, Aoyama H, Yoshimochi K, Fukamizu A. Acetylation of Foxo1 alters its DNA-binding ability and sensitivity to phosphorylation. Proc Natl Acad Sci U S A. 2005;102(32):11278–11283. doi: 10.1073/pnas.0502738102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.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. 2007;22(11):1251–1267. doi: 10.14670/hh-22.1251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.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. 2005;49(1):1–21. doi: 10.1016/j.brainresrev.2004.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Yui R, Matsuura ET. Detection of deletions flanked by short direct repeats in mitochondrial DNA of aging Drosophila. Mutat Res. 2006;594(1-2):155–161. doi: 10.1016/j.mrfmmm.2005.08.003. [DOI] [PubMed] [Google Scholar]
- 59.Regulska M, Leskiewicz M, Budziszewska B, Kutner A, Jantas D, Basta-Kaim A, Kubera M, Jaworska-Feil L, Lason W. Inhibitory effects of 1,25-dihydroxyvitamin D(3) and its low-calcemic analogues on staurosporine-induced apoptosis. Pharmacol Rep. 2007;59(4):393–401. [PubMed] [Google Scholar]
- 60.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(9):1380–1391. doi: 10.1097/00004647-200009000-00013. [DOI] [PubMed] [Google Scholar]
- 61.Maiese K. Triple play: Promoting neurovascular longevity with nicotinamide, WNT, and erythropoietin in diabetes mellitus. Biomed Pharmacother. 2008;62(4):218–232. doi: 10.1016/j.biopha.2008.01.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Maiese K, Chong ZZ. Nicotinamide: necessary nutrient emerges as a novel cytoprotectant for the brain. Trends Pharmacol Sci. 2003;24(5):228–232. doi: 10.1016/S0165-6147(03)00078-6. [DOI] [PubMed] [Google Scholar]
- 63.Ying W. NAD(+)/NADH and NADP(+)/NADPH in Cellular Functions and Cell Death: Regulation and Biological Consequences. Antioxid Redox Signal. 2008;10(2):179–206. doi: 10.1089/ars.2007.1672. [DOI] [PubMed] [Google Scholar]
- 64.Cohen SM, Cordeiro-Stone M, Kaufman DG. Early replication and the apoptotic pathway. J Cell Physiol. 2007;213(2):434–439. doi: 10.1002/jcp.21156. [DOI] [PubMed] [Google Scholar]
- 65.Harris SE, Fox H, Wright AF, Hayward C, Starr JM, Whalley LJ, Deary IJ. A genetic association analysis of cognitive ability and cognitive ageing using 325 markers for 109 genes associated with oxidative stress or cognition. BMC Genet. 2007;8:43. doi: 10.1186/1471-2156-8-43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Leuner K, Hauptmann S, Abdel-Kader R, Scherping I, Keil U, Strosznajder JB, Eckert A, Muller WE. Mitochondrial dysfunction: the first domino in brain aging and Alzheimer's disease? Antioxid Redox Signal. 2007;9(10):1659–1675. doi: 10.1089/ars.2007.1763. [DOI] [PubMed] [Google Scholar]
- 67.Maiese K, Vincent A, Lin SH, Shaw T. Group I and Group III metabotropic glutamate receptor subtypes provide enhanced neuroprotection. J Neurosci Res. 2000;62(2):257–272. doi: 10.1002/1097-4547(20001015)62:2<257::AID-JNR10>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
- 68.Mari C, Karabiyikoglu M, Goris ML, Tait JF, Yenari MA, Blankenberg FG. Detection of focal hypoxic-ischemic injury and neuronal stress in a rodent model of unilateral MCA occlusion/reperfusion using radiolabeled annexin V. Eur J Nucl Med Mol Imaging. 2004;31(5):733–739. doi: 10.1007/s00259-004-1473-5. [DOI] [PubMed] [Google Scholar]
- 69.Chong ZZ, Kang JQ, Maiese K. Metabotropic glutamate receptors promote neuronal and vascular plasticity through novel intracellular pathways. Histol Histopathol. 2003;18(1):173–189. doi: 10.14670/HH-18.173. [DOI] [PubMed] [Google Scholar]
- 70.Mallat M, Marin-Teva JL, Cheret C. Phagocytosis in the developing CNS: more than clearing the corpses. Curr Opin Neurobiol. 2005;15(1):101–107. doi: 10.1016/j.conb.2005.01.006. [DOI] [PubMed] [Google Scholar]
- 71.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. 2005;2(3):197–211. doi: 10.2174/1567202054368317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.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. 2006;3(3):187–201. doi: 10.2174/156720206778018758. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.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(3):262–275. doi: 10.1097/00004647-200103000-00010. [DOI] [PubMed] [Google Scholar]
- 74.Maiese K. The dynamics of cellular injury: transformation into neuronal and vascular protection. Histol Histopathol. 2001;16(2):633–644. doi: 10.14670/HH-16.633. [DOI] [PubMed] [Google Scholar]
- 75.Maiese K, Boccone L. Neuroprotection by peptide growth factors against anoxia and nitric oxide toxicity requires modulation of protein kinase C. J Cereb Blood Flow Metab. 1995;15(3):440–449. doi: 10.1038/jcbfm.1995.55. [DOI] [PubMed] [Google Scholar]
- 76.Vincent AM, Maiese K. Direct temporal analysis of apoptosis induction in living adherent neurons. J Histochem Cytochem. 1999;47(5):661–672. doi: 10.1177/002215549904700508. [DOI] [PubMed] [Google Scholar]
- 77.Chong ZZ, Lin SH, Kang JQ, Maiese K. The tyrosine phosphatase SHP2 modulates MAP kinase p38 and caspase 1 and 3 to foster neuronal survival. Cell Mol Neurobiol. 2003;23(4-5):561–578. doi: 10.1023/A:1025158314016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Maiese K, TenBroeke M, Kue I. Neuroprotection of lubeluzole is mediated through the signal transduction pathways of nitric oxide. J Neurochem. 1997;68(2):710–714. doi: 10.1046/j.1471-4159.1997.68020710.x. [DOI] [PubMed] [Google Scholar]
- 79.Salinas M, Diaz R, Abraham NG, Ruiz de Galarreta CM, Cuadrado A. Nerve growth factor protects against 6-hydroxydopamine-induced oxidative stress by increasing expression of heme oxygenase-1 in a phosphatidylinositol 3-kinase-dependent manner. J Biol Chem. 2003;278(16):13898–13904. doi: 10.1074/jbc.M209164200. [DOI] [PubMed] [Google Scholar]
- 80.Maiese K, Ahmad I, TenBroeke M, Gallant J. Metabotropic glutamate receptor subtypes independently modulate neuronal intracellular calcium. J Neurosci Res. 1999;55:472–485. doi: 10.1002/(SICI)1097-4547(19990215)55:4<472::AID-JNR7>3.0.CO;2-Q. [DOI] [PubMed] [Google Scholar]
- 81.Maiese K, Vincent AM. Critical temporal modulation of neuronal programmed cell injury. Cell Mol Neurobiol. 2000;20(3):383–400. doi: 10.1023/a:1007070311203. [DOI] [PubMed] [Google Scholar]
- 82.Maiese K, Vincent AM. Membrane asymmetry and DNA degradation: functionally distinct determinants of neuronal programmed cell death. J Neurosci Res. 2000;59(4):568–580. doi: 10.1002/(SICI)1097-4547(20000215)59:4<568::AID-JNR13>3.0.CO;2-R. [DOI] [PubMed] [Google Scholar]
- 83.Vincent AM, Maiese K. Nitric oxide induction of neuronal endonuclease activity in programmed cell death. Exp Cell Res. 1999;246(2):290–300. doi: 10.1006/excr.1998.4282. [DOI] [PubMed] [Google Scholar]
- 84.Vincent AM, TenBroeke M, Maiese K. Metabotropic glutamate receptors prevent programmed cell death through the modulation of neuronal endonuclease activity and intracellular pH. Exp Neurol. 1999;155(1):79–94. doi: 10.1006/exnr.1998.6966. [DOI] [PubMed] [Google Scholar]
- 85.Nakamura T, Sakamoto K. Forkhead transcription factor FOXO subfamily is essential for reactive oxygen species-induced apoptosis. Mol Cell Endocrinol. 2007;281(1-2):47–55. doi: 10.1016/j.mce.2007.10.007. [DOI] [PubMed] [Google Scholar]
- 86.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(2):325–339. doi: 10.1016/j.cell.2007.01.003. [DOI] [PubMed] [Google Scholar]
- 87.Barthelemy C, Henderson CE, Pettmann B. Foxo3a induces motoneuron death through the Fas pathway in cooperation with JNK. BMC Neurosci. 2004;5(1):48. doi: 10.1186/1471-2202-5-48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.You H, Yamamoto K, Mak TW. Regulation of transactivation-independent proapoptotic activity of p53 by FOXO3a. Proc Natl Acad Sci U S A. 2006;103(24):9051–9056. doi: 10.1073/pnas.0600889103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Won CK, Ji HH, Koh PO. Estradiol prevents the focal cerebral ischemic injury-induced decrease of forkhead transcription factors phosphorylation. Neurosci Lett. 2006;398(1-2):39–43. doi: 10.1016/j.neulet.2005.12.060. [DOI] [PubMed] [Google Scholar]
- 90.Zheng WH, Kar S, Quirion R. FKHRL1 and its homologs are new targets of nerve growth factor Trk receptor signaling. J Neurochem. 2002;80(6):1049–1061. doi: 10.1046/j.0022-3042.2002.00783.x. [DOI] [PubMed] [Google Scholar]
- 91.Li F, Chong ZZ, Maiese K. Winding through the WNT pathway during cellular development and demise. Histol Histopathol. 2006;21(1):103–124. doi: 10.14670/hh-21.103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Maiese K, Li F, Chong ZZ, Shang YC. The Wnt signaling pathway: Aging gracefully as a protectionist? Pharmacol Ther. 2008;118(1):58–81. doi: 10.1016/j.pharmthera.2008.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Li F, Chong ZZ, Maiese K. Vital elements of the wnt-frizzled signaling pathway in the nervous system. Curr Neurovasc Res. 2005;2(4):331–340. doi: 10.2174/156720205774322557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.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. 2007;19(6):1150–1162. doi: 10.1016/j.cellsig.2006.12.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Smith WW, Norton DD, Gorospe M, Jiang H, Nemoto S, Holbrook NJ, Finkel T, Kusiak JW. Phosphorylation of p66Shc and forkhead proteins mediates Abeta toxicity. J Cell Biol. 2005;169(2):331–339. doi: 10.1083/jcb.200410041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Hoogeboom D, Essers MA, Polderman PE, Voets E, Smits LM, Burgering BM. Interaction of FOXO with {beta}-Catenin Inhibits {beta}-Catenin/T Cell Factor Activity. J Biol Chem. 2008;283(14):9224–9230. doi: 10.1074/jbc.M706638200. [DOI] [PubMed] [Google Scholar]
- 97.Fukumoto S, Hsieh CM, Maemura K, Layne MD, Yet SF, Lee KH, Matsui T, Rosenzweig A, Taylor WG, Rubin JS, Perrella MA, Lee ME. Akt participation in the Wnt signaling pathway through Dishevelled. J Biol Chem. 2001;276(20):17479–17483. doi: 10.1074/jbc.C000880200. [DOI] [PubMed] [Google Scholar]
- 98.Naito AT, Akazawa H, Takano H, Minamino T, Nagai T, Aburatani H, Komuro I. Phosphatidylinositol 3-kinase-Akt pathway plays a critical role in early cardiomyogenesis by regulating canonical Wnt signaling. Circ Res. 2005;97(2):144–151. doi: 10.1161/01.RES.0000175241.92285.f8. [DOI] [PubMed] [Google Scholar]
- 99.van de Schans VA, van den Borne SW, Strzelecka AE, Janssen BJ, van der Velden JL, Langen RC, Wynshaw-Boris A, Smits JF, Blankesteijn WM. Interruption of Wnt signaling attenuates the onset of pressure overload-induced cardiac hypertrophy. Hypertension. 2007;49(3):473–480. doi: 10.1161/01.HYP.0000255946.55091.24. [DOI] [PubMed] [Google Scholar]
- 100.Barandon L, Dufourcq P, Costet P, Moreau C, Allieres C, Daret D, Dos Santos P, Daniel Lamaziere JM, Couffinhal T, Duplaa C. Involvement of FrzA/sFRP-1 and the Wnt/frizzled pathway in ischemic preconditioning. Circ Res. 2005;96(12):1299–1306. doi: 10.1161/01.RES.0000171895.06914.2c. [DOI] [PubMed] [Google Scholar]
- 101.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(9):2975–2980. doi: 10.1073/pnas.0400093101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Liu L, Rajareddy S, Reddy P, Du C, Jagarlamudi K, Shen Y, Gunnarsson D, Selstam G, Boman K, Liu K. Infertility caused by retardation of follicular development in mice with oocyte-specific expression of Foxo3a. Development. 2007;134(1):199–209. doi: 10.1242/dev.02667. [DOI] [PubMed] [Google Scholar]
- 103.Watkins WJ, Umbers AJ, Woad KJ, Harris SE, Winship IM, Gersak K, Shelling AN. Mutational screening of FOXO3A and FOXO1A in women with premature ovarian failure. Fertility and sterility. 2006;86(5):1518–1521. doi: 10.1016/j.fertnstert.2006.03.054. [DOI] [PubMed] [Google Scholar]
- 104.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]
- 105.Maiese K, Li F, Chong ZZ. Erythropoietin in the brain: can the promise to protect be fulfilled? Trends Pharmacol Sci. 2004;25(11):577–583. doi: 10.1016/j.tips.2004.09.006. [DOI] [PubMed] [Google Scholar]
- 106.Chong ZZ, Shang YC, Maiese K. Vascular injury during elevated glucose can be mitigated by erythropoietin and Wnt signaling. Curr Neurovasc Res. 2007;4(3):194–204. doi: 10.2174/156720207781387150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Maiese K, Chong ZZ, Shang YC. Raves and risks for erythropoietin. Cytokine Growth Factor Rev. 2008;19(2):145–155. doi: 10.1016/j.cytogfr.2008.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Mahmud DL, M GA, 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(10):1556–1562. doi: 10.1038/sj.onc.1205230. [DOI] [PubMed] [Google Scholar]
- 109.Bakker WJ, van Dijk TB, Parren-van Amelsvoort M, Kolbus A, Yamamoto K, Steinlein P, Verhaak RG, Mak TW, Beug H, Lowenberg B, von Lindern M. Differential regulation of Foxo3a target genes in erythropoiesis. Mol Cell Biol. 2007;27(10):3839–3854. doi: 10.1128/MCB.01662-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Srinivasan S, Anitha M, Mwangi S, Heuckeroth RO. Enteric neuroblasts require the phosphatidylinositol 3-kinase/Akt/Forkhead pathway for GDNF-stimulated survival. Mol Cell Neurosci. 2005;29(1):107–119. doi: 10.1016/j.mcn.2005.02.005. [DOI] [PubMed] [Google Scholar]
- 111.Maiese K, Chong ZZ, Shang YC. Mechanistic insights into diabetes mellitus and oxidative stress. Curr Med Chem. 2007;14(16):1729–1738. doi: 10.2174/092986707781058968. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Maiese K, Morhan SD, Chong ZZ. Oxidative stress biology and cell injury during type 1 and type 2 diabetes mellitus. Curr Neurovasc Res. 2007;4(1):63–71. doi: 10.2174/156720207779940653. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.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(5):1047–1053. doi: 10.2337/diacare.27.5.1047. [DOI] [PubMed] [Google Scholar]
- 114.Laakso M. Cardiovascular disease in type 2 diabetes: challenge for treatment and prevention. J Intern Med. 2001;249(3):225–235. doi: 10.1046/j.1365-2796.2001.00789.x. [DOI] [PubMed] [Google Scholar]
- 115.Dabelea D, Bell RA, D'Agostino RB, Jr, Imperatore G, Johansen JM, Linder B, Liu LL, Loots B, Marcovina S, Mayer-Davis EJ, Pettitt DJ, Waitzfelder B. Incidence of diabetes in youth in the United States. JAMA. 2007;297(24):2716–2724. doi: 10.1001/jama.297.24.2716. [DOI] [PubMed] [Google Scholar]
- 116.Jacobson AM, Musen G, Ryan CM, Silvers N, Cleary P, Waberski B, Burwood A, Weinger K, Bayless M, Dahms W, Harth J. Long-term effect of diabetes and its treatment on cognitive function. N Engl J Med. 2007;356(18):1842–1852. doi: 10.1056/NEJMoa066397. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Donahoe SM, Stewart GC, McCabe CH, Mohanavelu S, Murphy SA, Cannon CP, Antman EM. Diabetes and mortality following acute coronary syndromes. JAMA. 2007;298(7):765–775. doi: 10.1001/jama.298.7.765. [DOI] [PubMed] [Google Scholar]
- 118.Maiese K. Diabetic stress: new triumphs and challenges to maintain vascular longevity. Expert Rev Cardiovasc Ther. 2008;6(3):281–284. doi: 10.1586/14779072.6.3.281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Schnaider Beeri M, Goldbourt U, Silverman JM, Noy S, Schmeidler J, Ravona-Springer R, Sverdlick A, Davidson M. Diabetes mellitus in midlife and the risk of dementia three decades later. Neurology. 2004;63(10):1902–1907. doi: 10.1212/01.wnl.0000144278.79488.dd. [DOI] [PubMed] [Google Scholar]
- 120.Ceriello A, dello Russo P, Amstad P, Cerutti P. High glucose induces antioxidant enzymes in human endothelial cells in culture. Evidence linking hyperglycemia and oxidative stress. Diabetes. 1996;45(4):471–477. doi: 10.2337/diab.45.4.471. [DOI] [PubMed] [Google Scholar]
- 121.Ihara Y, Toyokuni S, Uchida K, Odaka H, Tanaka T, Ikeda H, Hiai H, Seino Y, Yamada Y. Hyperglycemia causes oxidative stress in pancreatic beta-cells of GK rats, a model of type 2 diabetes. Diabetes. 1999;48(4):927–932. doi: 10.2337/diabetes.48.4.927. [DOI] [PubMed] [Google Scholar]
- 122.Ling PR, Mueller C, Smith RJ, Bistrian BR. Hyperglycemia induced by glucose infusion causes hepatic oxidative stress and systemic inflammation, but not STAT3 or MAP kinase activation in liver in rats. Metabolism. 2003;52(7):868–874. doi: 10.1016/s0026-0495(03)00057-x. [DOI] [PubMed] [Google Scholar]
- 123.Yano M, Hasegawa G, Ishii M, Yamasaki M, Fukui M, Nakamura N, Yoshikawa T. Short-term exposure of high glucose concentration induces generation of reactive oxygen species in endothelial cells: implication for the oxidative stress associated with postprandial hyperglycemia. Redox Rep. 2004;9(2):111–116. doi: 10.1179/135100004225004779. [DOI] [PubMed] [Google Scholar]
- 124.Memisogullari R, Bakan E. Levels of ceruloplasmin, transferrin, and lipid peroxidation in the serum of patients with Type 2 diabetes mellitus. J Diabetes Complications. 2004;18(4):193–197. doi: 10.1016/S1056-8727(03)00032-1. [DOI] [PubMed] [Google Scholar]
- 125.Monnier L, Mas E, Ginet C, Michel F, Villon L, Cristol JP, Colette C. Activation of oxidative stress by acute glucose fluctuations compared with sustained chronic hyperglycemia in patients with type 2 diabetes. JAMA. 2006;295(14):1681–1687. doi: 10.1001/jama.295.14.1681. [DOI] [PubMed] [Google Scholar]
- 126.Guo S, Rena G, Cichy S, He X, Cohen P, Unterman T. Phosphorylation of serine 256 by protein kinase B disrupts transactivation by FKHR and mediates effects of insulin on insulin-like growth factor-binding protein-1 promoter activity through a conserved insulin response sequence. J Biol Chem. 1999;274(24):17184–17192. doi: 10.1074/jbc.274.24.17184. [DOI] [PubMed] [Google Scholar]
- 127.Nakae J, Park BC, Accili D. Insulin stimulates phosphorylation of the forkhead transcription factor FKHR on serine 253 through a Wortmannin-sensitive pathway. J Biol Chem. 1999;274(23):15982–15985. doi: 10.1074/jbc.274.23.15982. [DOI] [PubMed] [Google Scholar]
- 128.Kim JR, Jung HS, Bae SW, Kim JH, Park BL, Choi YH, Cho HY, Cheong HS, Shin HD. Obesity. 2. Vol. 14. Silver Spring, Md: 2006. Polymorphisms in FOXO gene family and association analysis with BMI; pp. 188–193. [DOI] [PubMed] [Google Scholar]
- 129.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(3):294–301. doi: 10.1038/sj.ejhg.5201766. [DOI] [PubMed] [Google Scholar]
- 130.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(8):1051–1062. doi: 10.1084/jem.20040942. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.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(12):3325–3335. doi: 10.1681/ASN.2006070754. [DOI] [PubMed] [Google Scholar]
- 132.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(1):23–31. doi: 10.2337/diabetes.54.1.23. [DOI] [PubMed] [Google Scholar]
- 133.Rachek LI, Thornley NP, Grishko VI, LeDoux SP, Wilson GL. Protection of INS-1 cells from free fatty acid-induced apoptosis by targeting hOGG1 to mitochondria. Diabetes. 2006;55(4):1022–1028. doi: 10.2337/diabetes.55.04.06.db05-0865. [DOI] [PubMed] [Google Scholar]
- 134.Kelley DE, He J, Menshikova EV, Ritov VB. Dysfunction of mitochondria in human skeletal muscle in type 2 diabetes. Diabetes. 2002;51(10):2944–2950. doi: 10.2337/diabetes.51.10.2944. [DOI] [PubMed] [Google Scholar]
- 135.Choo HJ, Kim JH, Kwon OB, Lee CS, Mun JY, Han SS, Yoon YS, Yoon G, Choi KM, Ko YG. Mitochondria are impaired in the adipocytes of type 2 diabetic mice. Diabetologia. 2006;49(4):784–791. doi: 10.1007/s00125-006-0170-2. [DOI] [PubMed] [Google Scholar]
- 136.Petersen KF, Befroy D, Dufour S, Dziura J, Ariyan C, Rothman DL, DiPietro L, Cline GW, Shulman GI. Mitochondrial dysfunction in the elderly: possible role in insulin resistance. Science. 2003;300(5622):1140–1142. doi: 10.1126/science.1082889. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Pospisilik JA, Knauf C, Joza N, Benit P, Orthofer M, Cani PD, Ebersberger I, Nakashima T, Sarao R, Neely G, Esterbauer H, Kozlov A, Kahn CR, Kroemer G, Rustin P, Burcelin R, Penninger JM. Targeted deletion of AIF decreases mitochondrial oxidative phosphorylation and protects from obesity and diabetes. Cell. 2007;131(3):476–491. doi: 10.1016/j.cell.2007.08.047. [DOI] [PubMed] [Google Scholar]
- 138.Puig O, Tjian R. Transcriptional feedback control of insulin receptor by dFOXO/FOXO1. Genes Dev. 2005;19(20):2435–2446. doi: 10.1101/gad.1340505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.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(39):41114–41123. doi: 10.1074/jbc.M400674200. [DOI] [PubMed] [Google Scholar]
- 140.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. 2005;2(4):271–285. doi: 10.2174/156720205774322584. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Chong ZZ, Lin SH, Maiese K. Nicotinamide Modulates Mitochondrial Membrane Potential and Cysteine Protease Activity during Cerebral Vascular Endothelial Cell Injury. J Vasc Res. 2002;39(2):131–147. doi: 10.1159/000057762. [DOI] [PubMed] [Google Scholar]
- 142.Tissenbaum HA, Guarente L. Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature. 2001;410(6825):227–230. doi: 10.1038/35065638. [DOI] [PubMed] [Google Scholar]
- 143.Nemoto S, Fergusson MM, Finkel T. Nutrient availability regulates SIRT1 through a forkhead-dependent pathway. Science. 2004;306(5704):2105–2108. doi: 10.1126/science.1101731. [DOI] [PubMed] [Google Scholar]
- 144.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(4):551–563. doi: 10.1016/s0092-8674(04)00126-6. [DOI] [PubMed] [Google Scholar]
- 145.Kobayashi Y, Furukawa-Hibi Y, Chen C, Horio Y, Isobe K, Ikeda K, Motoyama N. SIRT1 is critical regulator of FOXO-mediated transcription in response to oxidative stress. Int J Mol Med. 2005;16(2):237–243. [PubMed] [Google Scholar]
- 146.van der Horst A, Tertoolen LG, de Vries-Smits LM, Frye RA, Medema RH, Burgering BM. FOXO4 is acetylated upon peroxide stress and deacetylated by the longevity protein hSir2(SIRT1) J Biol Chem. 2004;279(28):28873–28879. doi: 10.1074/jbc.M401138200. [DOI] [PubMed] [Google Scholar]
- 147.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(1):4–9. doi: 10.1093/gerona/60.1.4. [DOI] [PubMed] [Google Scholar]
- 148.Ferrara N, Rinaldi B, Corbi G, Conti V, Stiuso P, Boccuti S, Rengo G, Rossi F, Filippelli A. Exercise Training Promotes SIRT1 Activity in Aged Rats. Rejuvenation Res. 11(1):139–150. doi: 10.1089/rej.2007.0576. [DOI] [PubMed] [Google Scholar]
- 149.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. 2006;281(52):40429–40439. doi: 10.1074/jbc.M606596200. [DOI] [PubMed] [Google Scholar]
- 150.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(1):212–220. doi: 10.1038/sj.emboj.7600045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.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. 2007;19(2):263–272. [PMC free article] [PubMed] [Google Scholar]
- 152.Maiese K, Chong ZZ. Insights into oxidative stress and potential novel therapeutic targets for Alzheimer disease. Restor Neurol Neurosci. 2004;22(2):87–104. [PubMed] [Google Scholar]
- 153.Dringen R. Oxidative and antioxidative potential of brain microglial cells. Antioxid Redox Signal. 2005;7(9-10):1223–1233. doi: 10.1089/ars.2005.7.1223. [DOI] [PubMed] [Google Scholar]
- 154.Ludikhuize J, de Launay D, Groot D, Smeets TJ, Vinkenoog M, Sanders ME, Tas SW, Tak PP, Reedquist KA. Inhibition of forkhead box class O family member transcription factors in rheumatoid synovial tissue. Arthritis Rheum. 2007;56(7):2180–2191. doi: 10.1002/art.22653. [DOI] [PubMed] [Google Scholar]
- 155.Kuo CC, Lin SC. Altered FOXO1 transcript levels in peripheral blood mononuclear cells of systemic lupus erythematosus and rheumatoid arthritis patients. Mol Med. 2007;13(11-12):561–566. doi: 10.2119/2007-00021.Kuo. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Bosque A, Aguilo JI, Alava MA, Paz-Artal E, Naval J, Allende LM, Anel A. The induction of Bim expression in human T-cell blasts is dependent on nonapoptotic Fas/CD95 signaling. Blood. 2007;109(4):1627–1635. doi: 10.1182/blood-2006-05-022319. [DOI] [PubMed] [Google Scholar]
- 157.Lin L, Hron JD, Peng SL. Regulation of NF-kappaB, Th activation, and autoinflammation by the forkhead transcription factor Foxo3a. Immunity. 2004;21(2):203–213. doi: 10.1016/j.immuni.2004.06.016. [DOI] [PubMed] [Google Scholar]
- 158.Jonsson H, Allen P, Peng SL. Inflammatory arthritis requires Foxo3a to prevent Fas ligand-induced neutrophil apoptosis. Nat Med. 2005;11(6):666–671. doi: 10.1038/nm1248. [DOI] [PubMed] [Google Scholar]
- 159.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(11):2971–2980. doi: 10.1002/eji.200636137. [DOI] [PubMed] [Google Scholar]
- 160.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(1):74–83. doi: 10.1038/ni1415. [DOI] [PubMed] [Google Scholar]
- 161.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(30):26729–26732. doi: 10.1074/jbc.C200256200. [DOI] [PubMed] [Google Scholar]
- 162.Kikuchi S, Nagai T, Kunitama M, Kirito K, Ozawa K, Komatsu N. Active FKHRL1 overcomes imatinib resistance in chronic myelogenous leukemia-derived cell lines via the production of tumor necrosis factor-related apoptosis-inducing ligand. Cancer Sci. 2007;98(12):1949–1958. doi: 10.1111/j.1349-7006.2007.00623.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Nowak K, Killmer K, Gessner C, Lutz W. E2F-1 regulates expression of FOXO1 and FOXO3a. Biochim Biophys Acta. 2007;1769(4):244–252. doi: 10.1016/j.bbaexp.2007.04.001. [DOI] [PubMed] [Google Scholar]
- 164.Jacobsen EA, Ananieva O, Brown ML, Chang Y. Growth, differentiation, and malignant transformation of pre-B cells mediated by inducible activation of v-Abl oncogene. J Immunol. 2006;176(11):6831–6838. doi: 10.4049/jimmunol.176.11.6831. [DOI] [PubMed] [Google Scholar]
- 165.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(5666):2011–2015. doi: 10.1126/science.1094637. [DOI] [PubMed] [Google Scholar]
- 166.Jin GS, Kondo E, Miyake T, Shibata M, Takashima T, Liu YX, Hayashi K, Akagi T, Yoshino T. Expression and intracellular localization of FKHRL1 in mammary gland neoplasms. Acta medica Okayama. 2004;58(4):197–205. doi: 10.18926/AMO/32088. [DOI] [PubMed] [Google Scholar]
- 167.Paik JH, Kollipara R, Chu G, Ji H, Xiao Y, Ding Z, Miao L, Tothova Z, Horner JW, Carrasco DR, Jiang S, Gilliland DG, Chin L, Wong WH, Castrillon DH, DePinho RA. FoxOs are lineage-restricted redundant tumor suppressors and regulate endothelial cell homeostasis. Cell. 2007;128(2):309–323. doi: 10.1016/j.cell.2006.12.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Sunters A, Madureira PA, Pomeranz KM, Aubert M, Brosens JJ, Cook SJ, Burgering BM, Coombes RC, Lam EW. Paclitaxel-induced nuclear translocation of FOXO3a in breast cancer cells is mediated by c-Jun NH2-terminal kinase and Akt. Cancer Res. 2006;66(1):212–220. doi: 10.1158/0008-5472.CAN-05-1997. [DOI] [PubMed] [Google Scholar]
- 169.Eddy SF, Kane SE, Sonenshein GE. Trastuzumab-resistant HER2-driven breast cancer cells are sensitive to epigallocatechin-3 gallate. Cancer Res. 2007;67(19):9018–9023. doi: 10.1158/0008-5472.CAN-07-1691. [DOI] [PubMed] [Google Scholar]
- 170.Lynch RL, Konicek BW, McNulty AM, Hanna KR, Lewis JE, Neubauer BL, Graff JR. The progression of LNCaP human prostate cancer cells to androgen independence involves decreased FOXO3a expression and reduced p27KIP1 promoter transactivation. Mol Cancer Res. 2005;3(3):163–169. doi: 10.1158/1541-7786.MCR-04-0163. [DOI] [PubMed] [Google Scholar]
- 171.Li Y, Wang Z, Kong D, Murthy S, Dou QP, Sheng S, Reddy GP, Sarkar FH. Regulation of FOXO3a/beta-catenin/GSK-3beta signaling by 3,3′-diindolylmethane contributes to inhibition of cell proliferation and induction of apoptosis in prostate cancer cells. J Biol Chem. 2007;282(29):21542–21550. doi: 10.1074/jbc.M701978200. [DOI] [PubMed] [Google Scholar]
- 172.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(13):4917–4930. doi: 10.1128/MCB.01789-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Ticchioni M, Essafi M, Jeandel PY, Davi F, Cassuto JP, Deckert M, Bernard A. Homeostatic chemokines increase survival of B-chronic lymphocytic leukemia cells through inactivation of transcription factor FOXO3a. Oncogene. 2007;26(50):7081–7091. doi: 10.1038/sj.onc.1210519. [DOI] [PubMed] [Google Scholar]
- 174.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(3):1511–1516. doi: 10.1128/JVI.01816-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Yang L, Xie S, Jamaluddin MS, Altuwaijri S, Ni J, Kim E, Chen YT, Hu YC, Wang L, Chuang KH, Wu CT, Chang C. Induction of androgen receptor expression by phosphatidylinositol 3-kinase/Akt downstream substrate, FOXO3a, and their roles in apoptosis of LNCaP prostate cancer cells. J Biol Chem. 2005;280(39):33558–33565. doi: 10.1074/jbc.M504461200. [DOI] [PubMed] [Google Scholar]
- 176.Arimoto-Ishida E, Ohmichi M, Mabuchi S, Takahashi T, Ohshima C, Hayakawa J, Kimura A, Takahashi K, Nishio Y, Sakata M, Kurachi H, Tasaka K, Murata Y. Inhibition of phosphorylation of a forkhead transcription factor sensitizes human ovarian cancer cells to cisplatin. Endocrinology. 2004;145(4):2014–2022. doi: 10.1210/en.2003-1199. [DOI] [PubMed] [Google Scholar]
- 177.Gomez-Gutierrez JG, Souza V, Hao HY, Montes de Oca-Luna R, Dong YB, Zhou HS, McMasters KM. Adenovirus-mediated gene transfer of FKHRL1 triple mutant efficiently induces apoptosis in melanoma cells. Cancer biology & therapy. 2006;5(7):875–883. doi: 10.4161/cbt.5.7.2911. [DOI] [PubMed] [Google Scholar]
- 178.Emami KH, Corey E. When prostate cancer meets bone: control by wnts. Cancer Lett. 2007;253(2):170–179. doi: 10.1016/j.canlet.2006.12.040. [DOI] [PubMed] [Google Scholar]
- 179.Perreault N, Katz JP, Sackett SD, Kaestner KH. Foxl1 controls the Wnt/beta-catenin pathway by modulating the expression of proteoglycans in the gut. J Biol Chem. 2001;276(46):43328–43333. doi: 10.1074/jbc.M104366200. [DOI] [PubMed] [Google Scholar]
- 180.Sauvageot CM, Kesari S, Stiles CD. Molecular pathogenesis of adult brain tumors and the role of stem cells. Neurol Clin. 2007;25(4):891–924. vii. doi: 10.1016/j.ncl.2007.07.014. [DOI] [PubMed] [Google Scholar]
- 181.Kurayoshi M, Oue N, Yamamoto H, Kishida M, Inoue A, Asahara T, Yasui W, Kikuchi A. Expression of Wnt-5a is correlated with aggressiveness of gastric cancer by stimulating cell migration and invasion. Cancer Res. 2006;66(21):10439–10448. doi: 10.1158/0008-5472.CAN-06-2359. [DOI] [PubMed] [Google Scholar]
- 182.Tomita H, Yamada Y, Oyama T, Hata K, Hirose Y, Hara A, Kunisada T, Sugiyama Y, Adachi Y, Linhart H, Mori H. Development of gastric tumors in Apc(Min/+) mice by the activation of the beta-catenin/Tcf signaling pathway. Cancer Res. 2007;67(9):4079–4087. doi: 10.1158/0008-5472.CAN-06-4025. [DOI] [PubMed] [Google Scholar]