Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 May 13.
Published in final edited form as: Rev Neurosci. 2008;19(4-5):341–361. doi: 10.1515/revneuro.2008.19.4-5.341

Role of Signal Transducer and Activator of Transcription 3 in Neuronal Survival and Regeneration

Suzan Dziennis 1, Nabil J Alkayed 1,2
PMCID: PMC2681485  NIHMSID: NIHMS97971  PMID: 19145989

Synopsis

Signal Transducers and Activators of Transcription (STATs) comprise a family of transcription factors that mediate a wide variety of biological functions in the central and peripheral nervous systems. Injury to neural tissue induces STAT activation, and STATs are increasingly recognized for their role in neuronal survival. In this review, we discuss the role of STAT3 during neural development and following ischemic and traumatic injury in brain, spinal cord and peripheral nerves. We focus on STAT3 because of the expanding body of literature that investigates protective and regenerative effects of growth factors, hormones and cytokines that use STAT3 to mediate their effect, in part through transcriptional upregulation of neuroprotective and neurotrophic genes. Defining the endogenous molecular mechanisms that lead to neuroprotection by STAT3 after injury might identify novel therapeutic targets against acute neural tissue damage as well as chronic neurodegenerative disorders.

Keywords: STAT3, ischemia, neuroregeneration, neuroprotection, axotomy, spinal cord injury, neurodevelopment

Introduction

STATs are activated in the central nervous system (CNS) following injury in response to multiple signaling pathways (Fig. 1). The primary mechanism for STAT activation in brain and spinal cord is believed to be in response to the multiple cytokines and growth factors that are released after injury. However, STATs can also be activated by free radicals, excitatory neurotransmitters and other inflammatory mediators that are also produced in damaged neural tissue [3]. STATs comprise a family of seven transcription factors: STAT1, STAT2, STAT3, STAT4, STAT5a, STAT5b and STAT6 [119]. STATs are expressed during development in both neurons and glia. In adult brain, STATs are normally quiescent, but as summarized in Tables 1 and 2, several STAT family members, including STAT1, STAT3 and STAT5, have been shown to be activated after injury. Interestingly, however, different STATs seem to play different roles in injured tissue. For example, whereas STAT3 and STAT5 have been linked to neuroprotection by trophic factors and cytokines after ischemic brain or nerve injury, STAT1 activation has been associated with neuronal cell death. STAT3 and STAT5 promote neuronal survival by inducing neuroprotective genes, whereas STAT1 promotes neurodegeneration by inducing apoptotic and other cell death promoting genes. Finally, although several STAT family members influence neuronal survival during development and in response to trophic factors and cytokines, STAT3 emerges as a key effector of neuronal survival after injury (Table 3).

Fig. 1.

Fig. 1

Schematic diagram summarizing mechanisms of STAT3 neuroprotection. Cytokines, growth factors and the hormone estradiol (E2) activate STAT3 by phosphorylation after ischemic or traumatic injury. STAT3 activation is linked to the neuroprotective role of these factors after injury. STAT3 activation results in the upregulation of genes that promote neuroprotection, neuroregeneration or neurodevelopment. RTK = receptor tyrosine kinase; P = phosphorylation; SIE = sis inducible element. See text for other abbreviations.

TABLE 1.

Temporal activation and cell type specificity of STATs following ischemic injury

STAT Type of ischemic injury Species Cell type Time after reperfusion Region Reference
STAT1 MCAO Rat Not determined 6-24 h
4 d, 7 d, 15 d
Cerebral cortex - decrease
Cerebral cortex - increase
[100]
STAT1 MCAO Mouse Neuron at 24 h .5 h, 2 h, 24 h Lateral striatum [145]
P-STAT1 MCAO Rat Not determined 0 h, 3 h Lesioned core and periphery [160]
STAT3 MCAO Rat Microglia 4 d, 7 d, 15 d Areas undergoing cell death
Cortex and striatum
[101]
STAT3 MCAO Rat Neuron
Neuron
Astrocyte
First hours
12 h
1-12 h
Nuclear expression in cortex and striatum
Boundaries of ischemic core of cortex and striatum
Increased nuclear translocation
[60]
P-STAT3 MCAO Rat Neuron
Neuron
Endothelial
Neuron
Microglia-monocyte/macrophage
3.5 h-168 h
Peak 3.5 h, decreased at 24 h
48 h
3.5 h, 24 h, 48 h
.5 h-168 h
Peri-ischemic area-cortex
Ischemic core-cortex
Ischemic core-cortex and striatum
Contralateral cortex and striatum
No phosphorylation detected in any region
[140]
P-STAT3 pMCAO Mouse Neuron 6-24 h Periphery of ischemic area [159]
STAT3 Ischemic preconditioning Rat Astrocyte 3 d, 7 d Hippocampus-dendritic CA1 molecular layer [64]
P-STAT3 Bilateral carotid global forebrain ischemia Rat Astrocyte 4 h, 1 d, 3 d Hippocampus-dentate hilar [25]
P-STAT3 MCAO Rat Activated monocyte/ macrophage predominant cell type, neuron, astrocyte 24 h, 72 h Core striatum and cortex
Peri-infarct striatum and cortex
No phosphorylation in contralateral side
[112]
P-STAT3 MCAO Female rat Undetermined
Neuron
3 h
22 h
Cortex
Peri-infarct region of cortex
[36]
P-STAT5 Hypoxic-ischemic neonatal model Rat-PND7 Not determined 1 d, 3 d Cortex [131]
P-STAT5 Transient global ischemia Rat Not determined 1 h, 3 h Hippocampus-CA1 [170]

MCAO = middle cerebral artery occlusion; pMCAO = permanent MCAO; PND = postnatal day.

TABLE 2.

Temporal activation and cell type specificity of STATs following nerve injury

STAT Type of injury Species Cell type Time after reperfusion Comments Reference
P-STAT3 Axotomy Right facial nerve Rat Motoneuron 24 h
5 d
Right facial nerve
Facial nuclei-no change in total protein levels of STAT3
[45]
P-STAT3 Axotomy Facial and hypoglossal nerve Rat Neuron
Astrocyte
3 h-42 d
1 d-133 d
P-STAT3 was not found in non-regenerating Clarks nucleus following axotomy [122]
STAT3 N-Methyl-D-aspartate induced excitotoxic cell death Rat PND9 Astrocyte, endothelial cells
Microglia/macrophages
Endothelial cells
Neurons
0 h, 2 h, 4 h, 10 h, 24 h, 3 d, 7d
10 h
0 h, 2 h, 4 h, 10 h, 24 h, 3 d, 7d
Not detected
Present solely in the nucleus of astrocytes at 2 h, 4 h, 10 h
All other times present both in nucleus and cytosol
Neocortex neurons showed STAT3 immunoreactivity prior to injury but not after
[2]
P-STAT3 Axotomy sciatic nerve Rat Neuron 15 min, 1 h, 6 h, 24 h, 1 wk, 6 wk In the DRG nuclei and axons at 6 h and 24 h
In motorneurons nuclei and axons at 24 h-1 wk, but gone at 6 wk
[126]
P-STAT3 Endorhinal cortex lesion Rat Reactive astrocytes
Neuron (GABAergic)
24 h
24 h strong, 3 d reduced
7 d undetectable
Fascia dentate
Medial septum
[163]
P-STAT3 Axotomy Sciatic nerve Rat Sensory neuron 6 h P-STAT3 observed in axons before nuclei [77]
P-STAT3 Neurochemical lesion with capsaicin Rat Sciatic nerve
DRG
6 h
30 min-6 h
Peak at 6 h then decreased to baseline
Peak 3-6 h then decreased to below baseline
[34]
P-STAT3 Axotomy Sciatic nerve
Dorsal column crush
Rat Neuron 6 h peak, present at 2 d, 7 d and 30 d Observed in the DRG after sciatic nerve axotomy but not after dorsal column crush [105]
P-STAT3 Spinal cord compression model Female mouse Neuron
Reactive astrocyte/microglia
Not determined
12 h strong, 48 h weak, 12 h weak, 48 h strong
Peak 12 h, decreasing but present at 168 h
Observed in the anterior horn [167]
P-STAT3 Spinal cord injury Female rat Not determined 4 h P-Ser 727 observed, but not P-Tyr 705 [152]

PND = postnatal day; DRG = dorsal root ganglia.

TABLE 3.

Effect of STAT inhibition/deletion on neuronal survival

STAT Injury Type of inhibition Species Phenotype Reference
STAT1 MCAO Gene deletion STAT1 Mouse Deletion of STAT1 decreased infarct volume [145]
STAT1 Gene deletion STAT1 bred with GFAP-IFNα Mouse Earlier onset and exacerbation of neurodegeneration [157]
STAT2 Gene deletion STAT2 bred with GFAP-IFNα Mouse Cerebellar dysplasia and medulloblastoma [156]
STAT3 Ball-Cre –STAT3 gene Deletion sensory neurons Mouse Enhanced neuronal apoptosis [8]
STAT3 Axotomy of motoneuron Neuron specific -STAT3 gene deletion (NFL-CRE) Mouse STAT3 deletion decreased motoneuron survival [123]
STAT3 Contusive spinal cord injury Conditional STAT3 gene deletion (Nestin-CRE) Mouse STAT3 inhibition decreased reactive astrocyte migration to injury site, accompanied by decreased motor function [98]
STAT3 MCAO si-RNA Spontaneously hypertensive rat STAT3 inhibition decreased infarct size [112]
STAT3 MCAO Pharmacological inhibition of P-STAT3 Wistar rat STAT3 inhibition increased infarct size in estradiol-replaced female rats [36]

MCAO = middle cerebral artery occlusion; GFAP = glial fibrillary acidic protein; IFNα = interferon-alpha.

Mechanisms of Stat Activation

The cellular and molecular mechanisms of STAT activation have been well characterized in the context of growth factor and cytokine signaling. Growth factors are signaling proteins that stimulate cellular proliferation and differentiation. They include cytokines and hormones that act on multiple cell types in the nervous system during development and following injury. Figure 1 demonstrates that several of those growth factors, which are protective against cell death and promote repair after injury, converge on STAT3.

As mentioned above, in the adult CNS, STATs reside in the cytosol in an inactivated state. The Janus kinase (JAK) family of tyrosine kinases, which are closely associated with the cognate receptors for multiple cytokines and growth factors, including interleukin-6 (IL-6), leukemia inhibitory factor (LIF), ciliary neurotrophic factor (CNTF), granulocyte-colony stimulating factor (G-CSF), erythropoietin (EPO) and insulin-like growth factor-I (IGF-I) [86,136,144] activate and work through STATs. Binding of these growth factors or cytokines to their cognate receptors activates JAK kinases, which phosphorylate tyrosine residues on the cytoplasmic portion of the receptor complex. These receptor-phosphorylated tyrosine residues then become docking sites for the STATs, which get phosphorylated by Jaks on a tyrosine residue into phosphorylated STAT (P-STAT). STATs can also be activated by growth factor receptors with intrinsic tyrosine kinase activity, such as the cognate receptors for epidermal growth factor (EGF) [5], platelet-derived growth factor (PDGF) [154], nerve growth factor (NGF) [34,95] and brain-derived neurotrophic factor (BDNF) [80]. In addition, STATs can also be activated by non-receptor associated tyrosine kinases belonging to the Src family of kinases.

Upon activation by tyrosine phosphorylation, members of the STAT family either heterodimerize or homodimerize and translocate to the nucleus to induce gene transcription (Fig. 1). STATs recognize and bind to a known consensus DNA sequence consisting of the 8-10 base pair (bp) inverted repeat 5′-TTn4-6AA-3′. In particular, consensus sites for STAT3 are evident in several neuroprotective gene promoters (see Fig. 1). The STAT binding consensus sequences are sometimes referred to as SIE (sis-inducible element), GAS (gamma activated sequence) or ISRE (interferon stimulated regulating element). STAT actions are terminated by deactivation in the nucleus by phosphatases [149] or by targeted degradation [65]. The unphosphorylated STATs are then recycled back to the cytosol [109]. There are further complexities of STAT deactivation as STAT activation can also be abrogated in the cytosol [116,161]. For example, activated STAT3 upregulates the suppressor of cytokine signaling 3 (SOCS3) gene, that inhibits STAT3 activation by suppressing JAK activity [70,79].

Stat Structure-Function Relationship

STAT family members share several common domains. STATs are comprised of an amino terminal, coiled-coil, DNA binding domain (except STAT2), linker, SH2 and transcriptional activation domain (TAD). The amino terminal domain is required for STAT dimerization after activation [59]. The coiled-coil domain allows interaction with other transcription factors and regulatory proteins [19,155]. The DNA binding domain, which varies slightly among individual STATs, recognizes the consensus binding regions (SIEs) on gene promoters. STAT3 favors the canonical binding site 5′-TTn4AA-3′ [125], whereas STAT1 favors the sequence 5′-TTn3AA-3′ [120]. As mentioned above, STAT1 binding sites are present in the pro-apoptotic genes caspase-1, caspase-2 and caspase-3 [22,24,72], whereas STAT3 binding sites are present in the anti-apoptotic genes bcl-2 and bcl-xL [137]. The linker region is involved in transcriptional activation [124] and protein-protein interaction [84]. Adjacent to the linker region is the SH2 domain that recognizes phosphotyrosine residues and docks the STAT protein to tyrosine phosphorylated cytokine and growth factor receptors. The SH2 domain also contains the highly conserved tyrosine residue that is required for STAT activation. In close proximity to the tyrosine residue is a serine residue, located in the transcriptional activation domain (TAD). Serine phosphorylation at this site can either enhance or suppress gene transcription (P-Ser-STAT) [19,30, 57,69]. This site is a putative mitogen-activated protein kinase (MAPK) phosphorylation motif, but the identity of the kinase(s) responsible for these modifications is not entirely elucidated [19]. Therefore other signaling pathways activated by ischemia, such as the MAPK and the AKT pathway [66], may directly or indirectly alter STAT function. While the other domains are highly conserved among STAT family members, the TADs are variable, which may contribute to their specificity of function. This may explain why STAT3, in addition to preferring known promoter sequences, is uniquely able to induce protective genes, in contrast to STAT1, which induces apoptotic genes (see the section below on STAT induction of neuroprotective genes). Although little is known about the structure-function relationship of the TAD in STAT3, STAT3 may partner with transcriptional co-activators and the combination of STAT3 with other transcription factors may increase the transcription of neuroprotective genes.

Stat Expression in the Developing and Mature Nervous System

In rodents, protein expression of STAT3 is detected in the cerebral cortex and striatum starting at embryonic day 14 (E14), and in the hippocampus at E18. STAT3 protein [29] and mRNA [171] are expressed continually from E14 to 3 months. During postnatal development, STAT3 expression increases from postnatal day 3 (P3) through P21 [40]. Immunohistochemistry has been employed to visualize activity, by localization in cytoplasm versus nucleus, and cell type specificity. STAT3 is localized to the nucleus in the developing rat brain at P3 with increasing intensity at P10 in the corpus callosum, cortex, cerebellar white matter, and stellate cells, suggesting that STAT3 is transcriptionally active during neurodevelopment. Total STAT3 is expressed in both neuronal and glial cell types throughout the rat brain during development [91] and in the adult [99]. Gautron et al. [40] found that during development, glial cells posses more nuclear reactivity than neurons, with the exception of robust immunoreactivity in brainstem cranial nerves (facial, hypoglossal) at P3 and P10. Of particular interest to nerve injury, brainstem cranial nerves have regenerative capabilities in the adult after nerve transection.

STAT3 is implicated in developmental processes that are later needed for support and repair of neurons in the adult after injury. STAT3 is required for vertebrate motor neuron pathfinding and guidance cues [16,26] and neurite outgrowth [50,55,162]. Glia provide structural and trophic support for neurons, and STAT3 is an important factor for the development of glial cell phenotype [17,107,148,153,168]. STAT3 induces the glia-specific protein glial fibrillary acidic protein (GFAP) [146] and is required for glial cell differentiation from cortical precursors cells in vitro [17]. In vivo, robust STAT3 activity precedes the onset of astrogliogenesis [49], suggesting that STAT3 is an integral component for differentiation of glial cells. Interestingly, neurons and glia arise from the same precursor cell population [9]. Although the precise mechanisms of neural development remain unclear, it has been proposed that failure of STAT3 to initiate glial gene expression in neuronal precursors is attributed to STAT-independent mechanisms, such as the requirement of additional transcription factors for neuronal development.

By P21, STAT3 activity declines in the brain with a few exceptions. One exception is in hypothalamic leptin-responsive neurons, where STAT3 phosphorylation is constitutive, suggesting that in addition to a developmental role, STAT3 is required for maintenance of the hypothalamic neuroendocrine axis in the adult [40,138]. Another exception is that activated STAT3 (P-STAT3) is detected in a small population of cells in the adult brain that continually undergo self-renewal [33,38, 83]. This ongoing postnatal neurogenesis occurs throughout life in the subventricular zone (SVZ) of the forebrain and in the subgranular zone (SGZ) of the hippocampus. STAT3 activation is ongoing in precursor cells within these regions throughout development and in the adult. Since endogenous neuronal precursor cells may be a source to replace dying cells after ischemic brain injury, STAT3 activation may contribute to the repair of neurons and glia after stroke.

The continual expression and nuclear localization of STAT3 during development suggests that STAT3 is an important mediator of the development and maintenance of the nervous system. These developmental processes involving STAT3 activation are important to consider because many of these developmental paradigms become reinstated after injury [49,51,71,81,82,87,140,142, 143]. Thus, STAT3 is poised to play a functional role in neuroprotection and repair after injury.

The importance of STAT3 in neurodevelopment and neuronal survival is highlighted by the fact that total STAT3 knockout animals are embryonic lethal [147]. These animals die prior to gastrulation (E6.5). In order to specifically determine the role of STAT3 in neuronal survival, several groups have circumvented embryonic lethality by generating tissue-, cell-type and time-specific knockout animals (Table 3). In particular, Cre-Lox technology has greatly improved our ability to study the effects of tissue specific and/or time specific gene deletion of STAT3 on neuronal survival (for review see [4]).

Alonzi et al. used the balancer strain (bal1) promoter to delete the STAT3 gene after E6.5 [8]. Bal1-cre is expressed in many tissues, including the brain, shortly after gastrulation (E6.5). The bal1-cre STAT3 knockout animals exhibit impaired survival of cranial sensory neurons from the nodose ganglia in vivo at El8. In vitro, cultured sensory neurons from bal1-STAT3 knock out animals show impaired responses to the neurotrophic cytokines CNTF and LIF. Interestingly, these knockout animals are deficient in their ability to activate signal transduction of another survival pathway, the PI3K/Akt pathway, in response to CNTF [8], suggesting a novel function of STAT3 in cytosolic signaling, apart from its role as a nuclear transcription factor. Bal1-Cre mediated STAT3 deletion, however, results in death within 5-8 hours after birth.

Neuron-specific STAT3 knockout mice are viable. However, mice harboring neural tissue-specific deletion of STAT3 during early development are obese, infertile, diabetic and are unable to maintain body temperature when challenged with cold exposure [39]. This phenotype is consistent with the strong STAT3 expression in hypothalamic leptin-responsive neurons [40,138], and suggest a role for STAT3 in the development of neuronal circuitry in the neuroendocrine axis, which controls feeding, appetite, thermoregulation and reproduction.

In the aging rodent brain, STAT3 expression progressively decreases between 3 and 26 months. In contrast, STAT1, which plays a deleterious role in stroke [145], remains unchanged in the cortex, striatum and hippocampus [28]. It is enticing, therefore, to speculate that these changes may be linked to age-related susceptibility and risk of stroke, especially, as mentioned above, that the loss of STAT3 is associated with the development of stroke risk factors, such as weight gain and diabetes mellitus. Whether age-related decline in STAT3 correlates with obesity, diabetes mellitus and increased risk and sensitivity to stroke in reproductively senescent animals, however, remains to be determined.

STAT3 Contributes to Neuroprotection After Ischemic Brain Injury

STAT3 activation following ischemic injury

STAT3 is activated in a time- and cell type-specific manner in ischemic brain (Table 1). STAT3 phosphorylation is observed in neurons [36,60,112,140], astrocytes, microglia [25,60,101, 140] and endothelial cells [112,140] in cortex, striatum and hippocampus in multiple models of ischemic brain injury both in vivo and in vitro (Table 1). In general, STAT3 phosphorylation is present in neurons during the acute phase of neural tissue injury, approximately from 0.5-24 h, and in astrocytes, microglia and endothelial cells during the chronic phases of injury, starting from 24 h to several days.

In transient cerebral ischemia models, neuronal STAT3 activation is slightly induced as early as 30 min after reperfusion [140] in the peri-ischemic area of the cerebral cortex, and increases at approximately 3 hours of reperfusion [36,140]. Robust nuclear P-STAT3 immunoreactivity is observed at 24 h, suggesting that STAT3 is functionally active and engaged in inducing gene transcription [36]. The majority of the data show that onset of glial STAT3 activation is most prevalent after 24 h of injury onset [64,112]. In some cases, however, STAT3 activation is observed in astrocytes at an earlier time. For example, cytosolic to nuclear translocation of STAT3 is observed in astrocytes as early as 1 h following focal cerebral ischemia [60]. STAT3 activation in glia remains high after several days, which may be linked to the role of STAT3 in the delayed inflammatory response to cerebral ischemia [169]. STAT3 activation is also apparent in endothelial cells of blood vessels at 24 h [112] and 48 h after experimental stroke induced in the rat by middle cerebral artery occlusion (MCAO) [140].

Role in neuroprotection

STAT3 activation in neurons correlates with survival in animal models of transient focal ischemia (MCAO). P-STAT3 co-localizes with the neuronal marker MAP-2 [36,140] and the anti-apoptotic protein bcl-2 [36,68], but not with cleaved-caspase 3 [166], suggesting that STAT3 is correlated with neuronal survival. In contrast, using a permanent occlusion MCAO model, Wen et al. [159] observed that P-STAT3-immunoreactive neurons were TUNEL-positive, leading the authors to conclude that STAT3 activation is associated with ischemic cell death. However, as seen in Table 3, the majority of functional, rather than correlative studies, which used STAT3 inhibitors and gene deletion approaches, support a role of STAT3 as a mediator of neuronal survival (Table 3).

Neurons depend on glia to provide structural, trophic and metabolic support after ischemia [10, 46,94]. Therefore, STAT3 activation in astrocytes may indirectly be linked to neuronal survival. In support of this idea, ischemic preconditioning increases STAT3 expression in reactive astrocytes in the hippocampus [64], and STAT3 phosphorylation correlates with increased expression of cIAP2, a member of the inhibitor of apoptosis protein family, in glial cells in the penumbra [134]. Interestingly, STAT3 phosphorylation correlates with decreased gliosis in the most severely damaged part of the brain, the striatal core of the infarct. In addition, Satriotomo and colleagues found that STAT3 knockdown by siRNA, rather than inhibition of STAT3 activity, reduced brain damage after MCAO in spontaneously hypertensive rats [112]. Furthermore, in that study, STAT3 phosphorylation was predominantly localized in microglia/macrophages, the main source of inflammatory cytokines in the ischemic brain. Because STAT3 is activated in both neurons and glia, it is possible that severe ischemic damage, such as observed in spontaneously hypertensive rats or in the core of the infarct, and in inflammatory cells [169], may override the neuronal protection.

Endothelial cell STAT3 is protective against ischemic injury in the heart [52,158] and will be an important avenue for further study in the brain. However, little is currently known about the role of STAT3 in endothelial cells after ischemic brain injury. For example, it has not been determined whether STAT3 phosphorylation observed within blood vessels originates in endothelial cells or is also due to the presence of inflammatory cells within the vasculature.

Role of STAT3 in Regeneration After Spinal Cord and Peripheral Nerve Injury

STAT3 may act as an injury-induced signaling molecule and transcription factor that promotes regeneration, and may serve as a new target for therapeutic intervention after traumatic injury to the spinal column or peripheral nerves. Neuronal regeneration after spinal cord injury depends on the survival of axotomized neurons. Therefore, experimental studies have focused on investigating the effect of STAT3 activation/phosphorylation on regeneration after injury induced by nerve crush, axotomy or neurochemical lesion.

Spinal cord compression injury induces phosphorylation of STAT3 in neurons in a pattern that resembles that seen after ischemic damage (Table 2). P-STAT3 is induced immediately following compression, and activation is sustained for up to 7 days, the latest time point analyzed [167]. STAT3 phosphorylation is first observed in motor neurons around 12 h, and later in reactivate astrocytes/ microglia at approximately 48 h after injury. AG-490, an inhibitor of JAK/STAT signaling, decreases hind limb motor function recovery after spinal cord compression injury, suggesting that the JAK/STAT pathway mediates motor neuron recovery. However, in this study, JAK phosphorylation was inhibited to a greater extent than STAT3 phosphorylation, suggesting that additional signaling cascades contribute to STAT3 activation following injury [167].

Conditional ablation of STAT3 in astrocytes revealed a surprising dual role for STAT3 after spinal cord injury [98]. Reactive astrocytes are known to migrate to the site of injury and form a physical barrier against inflammatory cells and neurotrophic factors. This barrier is commonly referred to as the glial scar and is believed to inhibit neuronal repair and axon regeneration through the scar. Since STAT3 is important for glial development, and therefore may contribute to the glial scar, it seems plausible that STAT3 deletion in astrocytes would be protective. Conditional ablation of STAT3 in astrocytes decreased migration of reactive astrocytes, but was accompanied by increased infiltration of CD11b+ (monocytes, macrophages and neutrophils; monocytic lineage) inflammatory cells, demyelination and more severe functional motor deficits, after spinal cord injury. When the SOCS3 gene was conditionally ablated, leading to prolonged activation of STAT3, increased reactive astrocyte migration to the site of injury was observed, and mice showed improvement of motor function compared to wild-type littermates, suggesting that P-STAT3-dependent recruitment of reactive astrocytes is beneficial, and that STAT3 activation dampens the influx of inflammatory cells, which may otherwise be harmful to the insult.

In general, axotomy increases P-STAT3 in the sciatic nerve [77,105,126], and STAT3 promotes axon regeneration after injury to the peripheral nervous system. Sensory neurons can regenerate if transected close to the nerve (sciatic), but not close to the spinal column. Qui and colleagues observed a positive correlation between sensory neurons' ability to regenerate and STAT3 phosphorylation [105]. STAT3 phosphorylation is observed in the dorsal root ganglia (DRG) sensory neurons following peripheral nerve injury to the sciatic nerve, but not in the spinal cord, when injury is induced by dorsal column crush. Neurochemical lesion with capsaicin, which destroys sensory neuron fibers but leaves their cell bodies intact [53], increases STAT3 phosphorylation in the DRG at 30 min, and in regenerating sensory fibers within the sciatic nerve at 6 h [34]. These findings are in agreement with the observation that mechanical lesion to the sciatic nerve increases P-STAT3 in spinal motor neurons and DRG during the regeneration period [77], and that P-STAT3 immunoreactivity decreases when regeneration is complete at 6 weeks [126], again implicating STAT3 in the regenerative process.

In vitro, STAT3 induces neurite outgrowth in the DRG. In an in vitro model of nerve injury-induced stress and overexcitation, application of Cortisol and kanaic acid to cultured DRG neurons induces STAT3 phosphorylation on Ser-727 and promotes neurite outgrowth [152], suggesting that STAT3 phosphorylation may be involved in neuroregeneration and repair after injury. STAT3 also correlates with the expression of an important protein for growth and repair of neurons, growth-associated protein 43 (GAP-43). GAP-43, which is used as a marker for nerve regeneration after nerve injury [97,150], is upregulated in STAT3 expressing cells [105] on growth cones. Furthermore, STAT3 induces GAP-43 transcriptional activity [162], and GAP-43 expression is inhibited with the JAK inhibitor AG-490, suggesting a role for the JAK/STAT3 pathway in GAP-43 expression and neurite outgrowth and regeneration.

STAT3 activation and correlation with regeneration has also been described in cranial motor neurons following axotomy. For example, axotomy induces STAT3 phosphorylation in hypoglossal motor neurons [122], and activated STAT3 is observed in astrocytes and sprouting septal neurons after endorhinal cortical lesion [163]. Schwaiger et al. found that STAT3 phosphorylation is increased in facial and hypoglossal neurons, which can regenerate, but not in neurons from the Clark's nucleus, which cannot regenerate [122]. In that study, increased STAT3 phosphorylation was observed acutely in neurons at 3 h with later activation in astrocytes at 24 h. Haas and colleagues [45] also observed STAT3 activation at 24 h and 5 d after facial motor neuron axotomy. These observations suggest that STAT3 plays an important role in neuronal repair after injury. In contrast, Luo et al. [85] found that STAT3 is activated following optic nerve transection and that inhibition of JAK/STAT promoted, rather than inhibited, retinal ganglion cell survival and axon regeneration. Ganglion cell survival was accompanied by inhibition of macrophage recruitment into the eye, suggesting that JAK/STAT recruits macrophages to the site of injury, which exacerbates retinal injury, and that inhibition of JAK/STAT may promote regeneration by inhibiting macrophage recruitment.

Although most nerve and spinal cord injury studies examined the role of STAT3 in the context of regeneration, Schweizer and colleagues [123] employed a tissue-specific gene ablation of STAT3 in motoneurons to study the role of neural STAT3 in neuronal survival following nerve injury in the adult. Cre recombinase was expressed under the neurofilament light chain promoter (NF-L), which is not expressed in neural tissue until E12. The animals survived into adulthood with no apparent gross phenotypic abnormalities. However, facial motor neuron survival was significantly reduced in these mice after nerve lesion in the adult. Moreover, the upregulation of two important survival genes, Reg-2 and bcl-xL, was reduced in STAT3 conditional knockout animals. The results from the NF-L/Cre STAT3 knockout study suggest an important role for STAT3 in promoting neuronal survival after injury, in part though the upregulation of neuroprotective genes.

Interestingly, STAT3 has been shown to be activated at the site of injury within nerve terminals, and then translocates retrogradely to the nucleus. Using compartmentalized cultures, addition of the neuropoeitic cytokine LIF to neurites of cultured sympathetic neurons induces P-STAT3 activation and subsequent translocation to nucleus [96]. Furthermore, STAT3 has been shown to be involved in retrograde signaling of Trk receptors [95]. A similar process seems to take place after injury in vivo, since STAT3 is localized to postsynaptic density (PSD) in cortical neurons, and STAT3 is phosphorylated by endogenous PSD tyrosine kinases [91]. Finally, P-STAT3 is observed in axons prior to its appearance in nuclei of sensory and motor neurons after injury [77,126].

Role of STAT3 in Neuroprotection by Growth Factors, Hormones and Cytokines

STAT3 is activated in response to growth factors, cytokines and hormones that are known to play a protective role after cerebral ischemia and nerve injury (Table 4). The neuropoeitic cytokines IL-6, LIF and CNTF, which are known to be induced after ischemia and exert neuroprotective effects [51,81,82,87,142], share the signaling receptor gp130. Cytokine-binding to this receptor complex induces either heterodimerization or homodimerization of receptor subunits, with IL-6 using a gp130 homodimer and all other family members a gp130/ LIF receptor heterodimer [11,27,41,56,135]. CNTF and IL-6 first bind to a non-signaling glycosylphosphatidylinositol-linked alpha receptor, CNTF-RA or IL6-RA, respectively, which then complexes with the LIFR and/or gp130 receptors to form a functional receptor signaling complex.

TABLE 4.

Protective factors utilizing STAT family members after injury

STAT Factor Method of admin. Species Injury model Comments Reference
STAT1 Theaflavin Intravenous Rat Transient focal ischemia Reduced infarct volume and reduced phosphorylation of STAT1 [20]
STAT3 G-CSF Injection of recombinant G-CSF Rat Transient focal ischemia Decreased infarct size [114]
STAT3 G-CSF Addition to culture media Rat Glutamate-induced excitotoxicity Protection [114]
STAT3 G-CSF Injection of recombinant G-CSF Mouse Transient focal ischemia Decreased infarct size [68]
STAT3 NGF Injection of recombinant NGF Rat Neurochemical lesion Maintained P-STAT3 and increased regeneration [34]
STAT3 IL-6 Blockade of IL-6RA Mouse Transient focal ischemia Increased cerebral damage associated with loss of P-STAT3 [166]
STAT3 IGF-I Addition to culture medium Rat Baseline survival IGF-I promotes survival of primary cortical neurons and activation of STAT3 [164]
STAT3 EPO Addition to culture medium Rat Optic nerve lesion Regeneration of CNS neurons associated with P-STAT3 [71]
STAT3 Estradiol Subcutaneous pellet Female rat MCAO Increased P-STAT3 after MCAO [36]
STAT5 EPO Single IP injection Rat PND7 MCAO Neuroprotective [131]
STAT5 EPO ICV infusion of EPO Rat Transient global ischemia Pro-survival of hippocampal neurons [170]

MCAO = middle cerebral artery occlusion; PND = postnatal day; G-CSF = granulocyte-colony stimulating factor: NGF = nerve growth factor; IL-6RA = interleukin-6 receptor alpha; IGF-I = insulin-like growth factor-I; EPO = erythropoietin.

LIF is released after nerve injury [141,172] and activates STAT3 [108]. Exogenous LIF reduces infarct volume after stroke [143], which correlates with increased P-STAT3 and the presence of LIFR/gp130, suggesting that LIF attenuates ischemic brain injury by activating STAT3 downstream of the LIFR/gp130 receptor complex.

IL-6 has also been shown to be neuroprotective. Intraventricular injection of IL-6 decreases infarct size in animal models of cerebral ischemia [51,82]. In vitro, IL-6 improves neuronal survival and promotes axon outgrowth and neuronal differentiation [47,74]. Although IL-6, LIF and CNTF share receptor components, Yamashita and colleagues have shown that IL-6 is specifically neuroprotective in cerebral ischemia. Endogenous IL-6 is increased after cerebral ischemia [139,142], and blockade of IL-6 signaling with an IL-6RA antibody ïncreases infarct size, number of apoptotic cells in the peri-infarct and neurological deficits after MCAO [166]. STAT3 phosphorylation is also reduced in IL-6RA-blocked animals [166], suggesting that the mechanism of neuroprotection by endogenous IL-6 may be mediated via STAT3.

The hematopoeitic cytokines G-CSF and EPO were primarily known as factors that promote blood cell proliferation and differentiation [92,102,128] until their expression was discovered in brain. G-CSF is a well-characterized neuroprotectant [115, 132], and is being evaluated for the treatment of stroke. G-CSF and G-CSFR are upregulated in neurons following ischemia [67,121], and recombinant human G-CSF is protective in both in vitro and in vivo models of cerebral ischemia [114]. The protection by G-CSF correlates with STAT3 phosphorylation, increased total STAT3 and increased anti-apoptotic proteins bcl-2 and Pim-1 in neurons [68,134], and with increased cellular inhibitor of apoptosis protein 2 (cIAP2) in glia [134]. The promoters of bcl-2, Pim-1 and cIAP2 are known transcriptional targets of STAT3 [13,106,127,137].

EPO and its receptor EPOR are also present in brain tissue [32,103,130], and EPO has been shown to be protective against hypoxia/ischemia [12,23, 111], excitotoxicity [61,90], chemical toxicity [42] and oxidative stress [21]. EPO has also been shown to be protective against ischemia-induced apoptosis [37,133] and traumatic brain [18,43] and spinal cord [23,43] injury. Interestingly, EPO promotes regeneration of CNS neurons via STAT3 [71], and EPO also maintains bcl-2 expression, which correlates with cell survival after hypoxia/ischemia [73]. Like G-CSF, EPO is also being evaluated in clinical trials for the treatment of stroke-related brain damage [37,58,115].

EGF reduces infarct size [58], induces neurogenesis after cerebral ischemia [93], and activates STAT3 in vascular smooth muscle cell [76]. Since STAT3 activation is also associated with neuroprotection and stem cell proliferation, it seems likely that STAT3 activation would underlie the protective effects of EGF after cerebral ischemia. However, whether EGF elicits neurogenesis and protects brain after stroke specifically via the STAT3 pathway remains to be studied.

IGF-I has also been shown to increase neuronal survival [62], and inhibition of the JAK/STAT pathway with AG-490 inhibits the ability of IGF-I to promote cell survival in primary cortical neurons. IGF-I increases neuritic length in cortical cultures [164], and concomitantly induces STAT3 phosphorylation, suggesting that STAT3 is involved in neurite outgrowth in response to IGF-I. One limitation of the use of the pharmacological inhibitor AG-490 is that it inhibits both JAK and STAT activities independently. Because JAKs activate other STAT family members in addition to STAT3, the effect of inhibiting other STATs cannot be separated in studies using AG-490.

Neurotrophins such as NGF and BDNF are critical for neuronal survival and neurodevelopment. In vitro, addition of NGF induces differentiation of neuronal phenotype in the rat adrenal pheochromocytoma cell line, which is a well-characterized model for studying NGF-induced neuronal differentiation. Recently, Ng et al. have shown that NGF activates STAT3 in these cells and results in STAT3 binding and upregulation of gene transcription [95], suggesting that STAT3 may play a role in neuronal differentiation in response to NGF. Interestingly, in this study STAT3 was activated by serine-727, rather than the classical tyrosine-705 phosphorylation. In the in vivo rodent model of capsaicin-induced sensory neuron lesion, regeneration was markedly enhanced with the addition of NGF [35,118]. Interestingly, addition of NGF maintains P-STAT3, when it would otherwise decrease during regeneration, suggesting that NGF may improve regeneration through P-STAT3 [34]. NGF also increases the STAT3 inducible GAP-43 gene in small sensory neurons after neurochemical lesion [117], further suggesting a potential role for STAT3 in regeneration by NGF after injury.

BDNF increases neurite outgrowth and arborization in hippocampal neurons in vitro [75] and in vivo [151]. Inhibition of STAT3 by siRNA abolishes BDNF-induced neurite extension in cultured hippocampal neurons, suggesting that BDNF promotes neurite outgrowth via STAT3. Lin et al. also demonstrated that BDNF promoted neurite extension through the JAK/STAT3 pathway in pelvic ganglia [80]. In one study, administration of BDNF after permanent MCAO reduced infarct volume at 24 h [165], whereas another study found no effect of BDNF on infarct volume, although neurological outcome was improved in that study at 6 weeks [113]. BDNF has not been linked to STAT3 in terms of its role in promoting functional recovery after stroke, and further studies are needed to determine whether BDNF plays a role in promoting axonal growth after either ischemic or nerve injury, and whether such an effect is mediated through STAT3.

PDGF or CNTF enhance oligodendrocyte progenitor cells survival upon trophic factor withdrawal, and both factors maintain oligodendrocyte survival in cultured cells in the presence of the cytotoxic cytokine tumor necrosis factor-α (TNF-α) [31]. In oligodendrocyte progenitor cells, addition of either PDGF or CNTF rapidly activates and induced nuclear translocation of STAT3. However, not all oligodendrocytes responded with robust STAT activation, which may be due to differences in the developmental stage of oligodendrocyte progenitor cells in vitro. Interestingly, PDGF resulted in strong activation of both STAT1 and STAT3, suggesting that the cascades of survival and differentiation are complex and may involve multiple STATs.

The female sex hormone estradiol reduces lesion size and neuronal death after experimental cerebral ischemia when administered at physiological concentrations (for review see [54]). Young adult female rats sustain smaller infarcts after MCAO compared to age-matched males [1 10]. This sex difference in ischemic brain injury in young adult rats disappears after ovariectomy [6], and is absent in middle-aged, reproductively senescent female rats in which ovarian function has naturally abated [7]. These observations suggest that ovarian hormones are protective against ischemic brain injury in young adult females.

Although ligand-bound estrogen receptor (ER) functions as a nuclear transcription factor, estrogen also induces rapid non-genomic cytosolic effects, such as changes in protein phosphorylation /78A Specifically, estradiol has been shown to rapidly activate STAT1, STAT3 and STAT5 in multiple cell types [14,15,63]. The role of STAT3 activation by estradiol has recently been linked to estrogen's neuroprotective effects against cerebral ischemia. We have recently reported that estradiol replacement increases STAT3 phosphorylation after transient focal ischemia relative to ovariectomized female rats. The protective effect of estradiol to reduce infarct volume appeared to be mediated via P-STAT3, because pharmacological inhibition of P-STAT3 abolished the protective effect of estradiol on infarct size [36]. Furthermore, we found that P-STAT3 was primarily activated in neurons, and that P-STAT3 co-localized in cells expressing the survival marker bcl-2. These findings suggest that STAT3 is an important player in mediating the protective effects of estradiol to reduce infarct size after MCAO in ovariectomized female rats.

Unlike the well-characterized activation of STAT3 by JAKs in response to growth factors and cytokines, the mechanism of ischemia-induced activation of STAT3 by estradiol is not known. However, recent studies suggest that a membrane-associated ER can interact with G-proteins [104] to activate Src [129], raising the possibility that estrogen may activate STAT3 after cerebral ischemia via Src.

STATs and Neuroprotective Genes

Several genes that promote survival and regeneration are activated in response to growth factors and cytokines, in part through STAT3 (Fig. 1). The bcl-2 family of proteins, including bcl-2 and bcl-xL, are critical in promoting neuronal survival after injury (reviewed in [88]). Bcl-2 is not expressed at high levels in mature neurons [1,89], but it is induced after ischemia. STAT3 expression directly induces bcl-2 and bcl-xL gene expression in cell lines [137]. Conversely, STAT1 activity suppresses bcl-2 and bcl-xL gene promoter activity [137] and induces pro-apoptotic genes such as caspase-1, -2 and -3 [24,72,145]. Interestingly, conditional knockout of the STAT3 gene in neurons led to decreased expression of survival genes Reg-2 and bcl-xL, but had no effect on the expression of bcl-2 [123], suggesting that additional factors may play a role in the expression of bcl-2, or that STAT3-induced expression of bcl-2 may differ among various types of neurons.

STAT3 binds and activates the expression of members of the inhibitor of apoptosis protein family, including survivin [44] and cIAP2 [48]. Survivin was identified as a potential STAT3-regulated gene by microarray analysis, which was confirmed by additional studies, including chromatin immunoprecipitation (ChIP), showing that STAT3 directly binds and regulates the survivin promoter in cancer cells. STAT3 is associated with the expression of GAP-43 during neurite outgrowth, indicating its importance in axonal repair [105]. Although STAT3 has been linked to the upregulation of multiple neuroprotective and neuro-repair genes after injury, few studies confirmed direct binding of STAT3 to these promoters. This may be due to technical difficulties in studying transcription factor/promoter interactions in vivo, since most studies characterizing STAT3 interaction with neuroprotective genes have been performed in cell lines. Whether STAT3 binds to and directly regulates bcl-2, bcl-xL, survivin, cIAP2 or GAP-43 expression in vivo after injury remains to be determined.

Summary

In this review we summarized accumulating evidence that STAT3 is an important player in neuroprotection and neuroregeneration, and a key factor for the development, differentiation, maintenance and survival of neurons and glia in the central and peripheral nervous systems. STATs are activated following ischemic injury and axotomy in a time- and cell type-specific manner. Of all STAT family members, STAT3 seems to be specifically linked to neuronal survival during development and after injury. It co-localizes with survival markers in neurons after ischemic injury, and is associated with survival and regeneration of neurons after axotomy. Data from knockout animals suggest that loss of STAT3 leads to deficits in neuronal function and survival after injury. Multiple growth factors, hormones and cytokines, including IL-6, LIF, G-CSF, EPO, IGF-I, NGF and estrogen, utilize STAT3 to protect the brain against cell death by inducing pro-survival and repair genes, including Reg-2, bcl-2, bcl-xL, survivin, Pirn-1, cIAP2 and GAP-43. The fact that signaling pathways utilized by growth factors, cytokines and hormones converge on STAT3 strongly suggests that STAT3 is a key and universal factor for promoting neuronal survival. Continuing to elucidate the molecular mechanisms of neuroprotection via the STAT pathway will, therefore, be crucial to our understanding of how the brain protects itself against insult, and may provide us with novel therapeutic approaches for neuroprotection in chronic neurodegenerative diseases as well as acute ischemic and traumatic brain and nerve injuries.

Acknowledgments

We would like to acknowledge Robin Feidelson for her assistance in preparing Figure 1, and John Todd and Kathy Gage for their assistance in preparing the manuscript. The authors would like to acknowledge support by National Institutes of Health grants F32 NS053065 (SD), NS044313 (NJA), and PO1 NS049210 (NJA).

References

  • 1.Abe-Dohmae S, Harada N, Yamada K, Tanaka R. Bcl-2 gene is highly expressed during neurogenesis in the central nervous system. Biochem Biophys Res Commun. 1993;191:915–921. doi: 10.1006/bbrc.1993.1304. [DOI] [PubMed] [Google Scholar]
  • 2.Acarin L, Gonzalez B, Castellano B. STAT3 and NFκB activation precedes glial reactivity in the excitotoxically injured young cortex but not in the corresponding distal thalamic nuclei. J Neuropathol Exp Neurol. 2000;59:151–163. doi: 10.1093/jnen/59.2.151. [DOI] [PubMed] [Google Scholar]
  • 3.Ahn YH, Lee G, Kang SK. Molecular insights of the injured lesions of rat spinal cords: inflammation, apoptosis, and cell survival. Biochem Biophys Res Commun. 2006;348:560–570. doi: 10.1016/j.bbrc.2006.07.105. [DOI] [PubMed] [Google Scholar]
  • 4.Akira S. Roles of STAT3 defined by tissue-specific gene targeting. Oncogene. 2000;19:2607–2611. doi: 10.1038/sj.onc.1203478. [DOI] [PubMed] [Google Scholar]
  • 5.Akira S. IL-6-regulated transcription factors. Int J Biochem Cell Biol. 1997;29:1401–1418. doi: 10.1016/s1357-2725(97)00063-0. [DOI] [PubMed] [Google Scholar]
  • 6.Alkayed NJ, Harukuni I, Kimes AS, London ED, Traystman RJ, Hurn PD. Gender-linked brain injury in experimental stroke. Stroke. 1998;29:159–165. doi: 10.1161/01.str.29.1.159. [DOI] [PubMed] [Google Scholar]
  • 7.Alkayed NJ, Murphy SJ, Traystman RJ, Hurn PD, Miller VM. Neuroprotective effects of female gonadal steroids in reproductively senescent female rats. Stroke. 2000;31:161–168. doi: 10.1161/01.str.31.1.161. [DOI] [PubMed] [Google Scholar]
  • 8.Alonzi T, Middleton G, Wyatt S, Buchman V, Betz UA, Muller W, Musiani P, Poli V, Davies AM. Role of STAT3 and PI 3-kinase/Akt in mediating the survival actions of cytokines on sensory neurons. Mol Cell Neurosci. 2001;18:270–282. doi: 10.1006/mcne.2001.1018. [DOI] [PubMed] [Google Scholar]
  • 9.Anderson DJ, Guo B, Xu Y, Ng LM, Kricka LJ, Skogerboe KJ, Hage DS, Schoeff L, Wang J, Sokoll LJ, Chan DW, Ward KM, Davis KA. Clinical chemistry. Anal Chem. 1997;69:165R–229R. doi: 10.1021/a1970008p. [DOI] [PubMed] [Google Scholar]
  • 10.Bambrick L, Kristian T, Fiskum G. Astrocyte mitochondrial mechanisms of ischemic brain injury and neuroprotection. Neurochem Res. 2004;29:601–608. doi: 10.1023/b:nere.0000014830.06376.e6. [DOI] [PubMed] [Google Scholar]
  • 11.Baumann H, Ziegler SF, Mosley B, Morella KK, Pajovic S, Gearing DP. Reconstitution of the response to leukemia inhibitory factor, oncostatin M, and ciliary neurotrophic factor in hepatoma cells. J Biol Chem. 1993;268:8414–8417. [PubMed] [Google Scholar]
  • 12.Bernaudin M, Marti HH, Roussel S, Divoux D, Nouvelot A, MacKenzie ET, Petit E. A potential role for erythropoietin in focal permanent cerebral ischemia in mice. J Cereb Blood Flow Metab. 1999;19:643–651. doi: 10.1097/00004647-199906000-00007. [DOI] [PubMed] [Google Scholar]
  • 13.Bhattacharya S, Ray RM, Johnson LR. STAT3-mediated transcription of Bcl-2, Mcl-1 and c-IAP2 prevents apoptosis in polyamine-depleted cells. Biochem J. 2005;392:335–344. doi: 10.1042/BJ20050465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Bjornstrom L, Sjoberg M. Mechanisms of estrogen receptor signaling: convergence of genomic and non-genomic actions on target genes. Mol Endocrinol. 2005;19:833–842. doi: 10.1210/me.2004-0486. [DOI] [PubMed] [Google Scholar]
  • 15.Bjornstrom L, Sjoberg M. Signal transducers and activators of transcription as downstream targets of nongenomic estrogen receptor actions. Mol Endocrinol. 2002;16:2202–2214. doi: 10.1210/me.2002-0072. [DOI] [PubMed] [Google Scholar]
  • 16.Boccaccio C, Ando M, Tamagnone L, Bardelli A, Michieli P, Battistini C, Comoglio PM. Induction of epithelial tubules by growth factor HGF depends on the STAT pathway. Nature. 1998;391:285–288. doi: 10.1038/34657. [DOI] [PubMed] [Google Scholar]
  • 17.Bonni A, Sun Y, Nadal-Vicens M, Bhatt A, Frank DA, Rozovsky I, Stahl N, Yancopoulos GD, Greenberg ME. Regulation of gliogenesis in the central nervous system by the JAK-STAT signaling pathway. Science. 1997;278:477–483. doi: 10.1126/science.278.5337.477. [DOI] [PubMed] [Google Scholar]
  • 18.Brines ML, Ghezzi P, Keenan S, Agnello D, de Lanerolle NC, Cerami C, Itri LM, Cerami A. Erythropoietin crosses the blood-brain barrier to protect against experimental brain injury. Proc Natl Acad Sci USA. 2000;97:10526–10531. doi: 10.1073/pnas.97.19.10526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Bromberg J, Darnell JE., Jr The role of STATs in transcriptional control and their impact on cellular function. Oncogene. 2000;19:2468–2473. doi: 10.1038/sj.onc.1203476. [DOI] [PubMed] [Google Scholar]
  • 20.Cai F, Li CR, Wu JL, Chen JG, Liu C, Min Q, Yu W, Ouyang CH, Chen JH. Theaflavin ameliorates cerebral ischemia-reperfusion injury in rats through its antiinflammatory effect and modulation of STAT-1. Mediators Inflamm. 2006;2006:30490. doi: 10.1155/MI/2006/30490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Calapai G, Marciano MC, Corica F, Allegra A, Parisi A, Frisina N, Caputi AP, Buemi M. Erythropoietin protects against brain ischemic injury by inhibition of nitric oxide formation. Eur J Pharmacol. 2000;401:349–356. doi: 10.1016/s0014-2999(00)00466-0. [DOI] [PubMed] [Google Scholar]
  • 22.Cattaneo E, Conti L, De-Fraja C. Signalling through the JAK-STAT pathway in the developing brain. Trends Neurosci. 1999;22:365–369. doi: 10.1016/s0166-2236(98)01378-2. [DOI] [PubMed] [Google Scholar]
  • 23.Celik M, Gokmen N, Erbayraktar S, Akhisaroglu M, Konakc S, Ulukus C, Genc S, Genc K, Sagiroglu E, Cerami A, Brines M. Erythropoietin prevents motor neuron apoptosis and neurologic disability in experimental spinal cord ischemic injury. Proc Natl Acad Sci USA. 2002;99:2258–2263. doi: 10.1073/pnas.042693799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Chin YE, Kitagawa M, Kuida K, Flavell RA, Fu XY. Activation of the STAT signaling pathway can cause expression of caspase I and apoptosis. Mol Cell Biol. 1997;17:5328–5337. doi: 10.1128/mcb.17.9.5328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Choi JS, Kim SY, Cha JH, Choi YS, Sung KW, Oh ST, Kim ON, Chung JW, Chun MH, Lee SB, Lee MY. Upregulation of gp130 and STAT3 activation in the rat hippocampus following transient forebrain ischemia. Glia. 2003;41:237–246. doi: 10.1002/glia.10186. [DOI] [PubMed] [Google Scholar]
  • 26.Conway G. STAT3-dependent pathfinding and control of axonal branching and target selection. Dev Biol. 2006;296:119–136. doi: 10.1016/j.ydbio.2006.04.444. [DOI] [PubMed] [Google Scholar]
  • 27.Davis S, Aldrich TH, Stahl N, Pan L, Taga T, Kishimoto T, Ip NY, Yancopoulos GD. LIFR beta and gp130 as heterodimerizing signal transducers of the tripartite CNTF receptor. Science. 1993;260:1805–1808. doi: 10.1126/science.8390097. [DOI] [PubMed] [Google Scholar]
  • 28.De-Fraja C, Conti L, Govoni S, Battaini F, Cattaneo E. STAT signalling in the mature and aging brain. Int J Dev Neurosci. 2000;18:439–446. doi: 10.1016/s0736-5748(00)00007-1. [DOI] [PubMed] [Google Scholar]
  • 29.De-Fraja C, Conti L, Magrassi L, Govoni S, Cattaneo E. Members of the JAK/STAT proteins are expressed and regulated during development in the mammalian forebrain. J Neurosci Res. 1998;54:320–330. doi: 10.1002/(SICI)1097-4547(19981101)54:3<320::AID-JNR3>3.0.CO;2-R. [DOI] [PubMed] [Google Scholar]
  • 30.Decker T, Kovarik P. Serine phosphorylation of STATs. Oncogene. 2000;19:2628–2637. doi: 10.1038/sj.onc.1203481. [DOI] [PubMed] [Google Scholar]
  • 31.Dell' Albani P, Kahn MA, Cole R, Condorelli DF, Giuffrida-Stella AM, de Vellis J. Oligodendroglial survival factors, PDGF-AA and CNTF, activate similar JAK/STAT signaling pathways. J Neurosci Res. 1998;54:191–205. doi: 10.1002/(SICI)1097-4547(19981015)54:2<191::AID-JNR7>3.0.CO;2-9. [DOI] [PubMed] [Google Scholar]
  • 32.Digicaylioglu M, Bichet S, Marti HH, Wenger RH, Rivas LA, Bauer C, Gassmann M. Localization of specific erythropoietin binding sites in defined areas of the mouse brain. Proc Natl Acad Sci USA. 1995;92:3717–3720. doi: 10.1073/pnas.92.9.3717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Doetsch F, Garcia-Verdugo JM, Alvarez-Buylla A. Regeneration of a germinal layer in the adult mammalian brain. Proc Natl Acad Sci USA. 1999;96:11619–11624. doi: 10.1073/pnas.96.20.11619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Donnerer J, Liebmann I, Schicho R. ERK and STAT3 phosphorylation in sensory neurons during capsaicin-induced impairment and nerve growth factor treatment. Pharmacology. 2005;75:116–121. doi: 10.1159/000088015. [DOI] [PubMed] [Google Scholar]
  • 35.Donnerer J, Liebmann I, Schicho R. Differential regulation of 3-beta-hydroxysteroid dehydrogenase and vanilloid receptor TRPV1 mRNA in sensory neurons by capsaicin and NGF. Pharmacology. 2005;73:97–101. doi: 10.1159/000081625. [DOI] [PubMed] [Google Scholar]
  • 36.Dziennis S, Jia T, Ronnekleiv OK, Hum PD, Alkayed NJ. Role of signal transducer and activator of transcription-3 in estradiol-mediated neuroprotection. J Neurosci. 2007;27:7268–7274. doi: 10.1523/JNEUROSCI.1558-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Ehrenreich H, Hasselblatt M, Dembowski C, Cepek L, Lewczuk P, Stiefel M, Rustenbeck HH, Breiter N, Jacob S, Knerlich F, Bohn M, Poser W, Ruther E, Kochen M, Gefeller O, Gleiter C, Wessel TC, De Ryck M, Itri L, Prange H, Cerami A, Brines M, Siren AL. Erythropoietin therapy for acute stroke is both safe and beneficial. Mol Med. 2002;8:495–505. [PMC free article] [PubMed] [Google Scholar]
  • 38.Eriksson PS, Perfilieva E, Bjork-Eriksson T, Alborn AM, Nordborg C, Peterson DA, Gage FH. Neurogenesis in the adult human hippocampus. Nat Med. 1998;4:1313–1317. doi: 10.1038/3305. [DOI] [PubMed] [Google Scholar]
  • 39.Gao Q, Wolfgang MJ, Neschen S, Morino K, Horvath TL, Shulman GI, Fu XY. Disruption of neural signal transducer and activator of transcription 3 causes obesity, diabetes, infertility, and thermal dysregulation. Proc Natl Acad Sci USA. 2004;101:4661–4666. doi: 10.1073/pnas.0303992101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Gautron L, De Smedt-Peyrusse V, Laye S. Characterization of STAT3-expressing cells in the postnatal rat brain. Brain Res. 2006;1098:26–32. doi: 10.1016/j.brainres.2006.04.115. [DOI] [PubMed] [Google Scholar]
  • 41.Gearing DP, Comeau MR, Friend DJ, Gimpel SD, Thut CJ, McGourty J, Brasher KK, King JA, Gillis S, Mosley B. The IL-6 signal transducer, gp130: an onco-statin M receptor and affinity converter for the LIF receptor. Science. 1992;255:1434–1437. doi: 10.1126/science.1542794. [DOI] [PubMed] [Google Scholar]
  • 42.Genc S, Kuralay F, Genc K, Akhisaroglu M, Fadiloglu S, Yorukoglu K, Fadiloglu M, Gure A. Erythropoietin exerts neuroprotection in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-treated C57/BL mice via increasing nitric oxide production. Neurosci Lett. 2001;298:139–141. doi: 10.1016/s0304-3940(00)01716-x. [DOI] [PubMed] [Google Scholar]
  • 43.Gorio A, Gokmen N, Erbayraktar S, Yilmaz O, Madaschi L, Cichetti C, Di Giulio AM, Vardar E, Cerami A, Brines M. Recombinant human erythropoietin counteracts secondary injury and markedly enhances neurological recovery from experimental spinal cord trauma. Proc Natl Acad Sci USA. 2002;99:9450–9455. doi: 10.1073/pnas.142287899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Gritsko T, Williams A, Turkson J, Kaneko S, Bowman T, Huang M, Nam S, Eweis I, Diaz N, Sullivan D, Yoder S, Enkemann S, Eschrich S, Lee JH, Beam CA, Cheng J, Minton S, Muro-Cacho CA, Jove R. Persistent activation of stat3 signaling induces survivin gene expression and confers resistance to apoptosis in human breast cancer cells. Clin Cancer Res. 2006;12:11–19. doi: 10.1158/1078-0432.CCR-04-1752. [DOI] [PubMed] [Google Scholar]
  • 45.Haas CA, Hofmann HD, Kirsch M. Expression of CNTF/LIF-receptor components and activation of STAT3 signaling in axotomized facial motoneurons: evidence for a sequential postlesional function of the cytokines. J Neurobiol. 1999;41:559–571. doi: 10.1002/(sici)1097-4695(199912)41:4<559::aid-neu11>3.0.co;2-a. [DOI] [PubMed] [Google Scholar]
  • 46.Haber M, Zhou L, Murai KK. Cooperative astrocyte and dendritic spine dynamics at hippocampal excitatory synapses. J Neurosci. 2006;26:8881–8891. doi: 10.1523/JNEUROSCI.1302-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Hama T, Kushima Y, Miyamoto M, Kubota M, Takei N, Hatanaka H. Interleukin-6 improves the survival of mesencephalic catecholaminergic and septal cholinergic neurons from postnatal, two-week-old rats in cultures. Neuroscience. 1991;40:445–452. doi: 10.1016/0306-4522(91)90132-8. [DOI] [PubMed] [Google Scholar]
  • 48.Hasegawa T, Suzuki K, Sakamoto C, Ohta K, Nishiki S, Hino M, Tatsumi N, Kitagawa S. Expression of the inhibitor of apoptosis (IAP) family members in human neutrophils: up-regulation of cIAP2 by granulocyte colony-stimulating factor and overexpression of cIAP2 in chronic neutrophilic leukemia. Blood. 2003;101:1164–1171. doi: 10.1182/blood-2002-05-1505. [DOI] [PubMed] [Google Scholar]
  • 49.He F, Ge W, Martinowich K, Becker-Catania S, Coskun V, Zhu W, Wu H, Castro D, Guillemot F, Fan G, de Vellis J, Sun YE. A positive autoregulatory loop of Jak-STAT signaling controls the onset of astrogliogenesis. Nat Neurosci. 2005;8:616–625. doi: 10.1038/nn1440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.He JC, Gomes I, Nguyen T, Jayaram G, Ram PT, Devi LA, Iyengar R. The G alpha(o/i)-coupled cannabinoid receptor-mediated neurite outgrowth involves Rap regulation of Src and Stat3. J Biol Chem. 2005;280:33426–33434. doi: 10.1074/jbc.M502812200. [DOI] [PubMed] [Google Scholar]
  • 51.Herrmann O, Tarabin V, Suzuki S, Attigah N, Coserea I, Schneider A, Vogel J, Prinz S, Schwab S, Monyer H, Brombacher F, Schwaninger M. Regulation of body temperature and neuroprotection by endogenous interleukin-6 in cerebral ischemia. J Cereb Blood Flow Metab. 2003;23:406–415. doi: 10.1097/01.WCB.0000055177.50448.FA. [DOI] [PubMed] [Google Scholar]
  • 52.Hilfiker-Kleiner D, Hilfiker A, Fuchs M, Kaminski K, Schaefer A, Schieffer B, Hillmer A, Schmiedl A, Ding Z, Podewski E, Podewski E, Poli V, Schneider MD, Schulz R, Park JK, Wollert KC, Drexler H. Signal transducer and activator of transcription 3 is required for myocardial capillary growth, control of interstitial matrix deposition, and heart protection from ischemic injury. Circ Res. 2004;95:187–195. doi: 10.1161/01.RES.0000134921.50377.61. [DOI] [PubMed] [Google Scholar]
  • 53.Holzer P. Capsaicin as a tool for studying sensory neuron functions. Adv Exp Med Biol. 1991;298:3–16. doi: 10.1007/978-1-4899-0744-8_1. [DOI] [PubMed] [Google Scholar]
  • 54.Hurn PD, Macrae IM. Estrogen as a neuroprotectant in stroke. J Cereb Blood Flow Metab. 2000;20:631–652. doi: 10.1097/00004647-200004000-00001. [DOI] [PubMed] [Google Scholar]
  • 55.Ihara S, Nakajima K, Fukada T, Hibi M, Nagata S, Hirano T, Fukui Y. Dual control of neurite outgrowth by STAT3 and MAP kinase in PC12 cells stimulated with interleukin-6. EMBO J. 1997;16:5345–5352. doi: 10.1093/emboj/16.17.5345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Ip NY, Nye SH, Boulton TG, Davis S, Taga T, Li Y, Birren SJ, Yasukawa K, Kishimoto T, Anderson DJ. CNTF and LIF act on neuronal cells via shared signaling pathways that involve the IL-6 signal transducing receptor component gp 130. Cell. 1992;69:1121–1132. doi: 10.1016/0092-8674(92)90634-o. [DOI] [PubMed] [Google Scholar]
  • 57.Jain N, Zhang T, Fong SL, Lim CP, Cao X. Repression of Stat3 activity by activation of mitogen-activated protein kinase (MAPK) Oncogene. 1998;17:3157–3167. doi: 10.1038/sj.onc.1202238. [DOI] [PubMed] [Google Scholar]
  • 58.Jin K, Sun Y, Xie L, Childs J, Mao XO, Greenberg DA. Post-ischemic administration of heparin-binding epidermal growth factor-like growth factor (HB-EGF) reduces infarct size and modifies neurogenesis after focal cerebral ischemia in the rat. J Cereb Blood Flow Metab. 2004;24:399–408. doi: 10.1097/00004647-200404000-00005. [DOI] [PubMed] [Google Scholar]
  • 59.John S, Vinkemeier U, Soldaini E, Darnell JE, Jr, Leonard WJ. The significance of tetramerization in promoter recruitment by Stat5. Mol Cell Biol. 1999;19:1910–1918. doi: 10.1128/mcb.19.3.1910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Justicia C, Gabriel C, Planas AM. Activation of the JAK/STAT pathway following transient focal cerebral ischemia: signaling through Jak1 and Stat3 in astrocytes. Glia. 2000;30:253–270. doi: 10.1002/(sici)1098-1136(200005)30:3<253::aid-glia5>3.0.co;2-o. [DOI] [PubMed] [Google Scholar]
  • 61.Kawakami M, Sekiguchi M, Sato K, Kozaki S, Takahashi M. Erythropoietin receptor-mediated inhibition of exocytotic glutamate release confers neuroprotection during chemical ischemia. J Biol Chem. 2001;276:39469–39475. doi: 10.1074/jbc.M105832200. [DOI] [PubMed] [Google Scholar]
  • 62.Kenchappa RS, Diwakar L, Annepu J, Ravindranath V. Estrogen and neuroprotection: higher constitutive expression of glutaredoxin in female mice offers protection against MPTP-mediated neurodegeneration. FASEB J. 2004;18:1102–1104. doi: 10.1096/fj.03-1075fje. [DOI] [PubMed] [Google Scholar]
  • 63.Kennedy AM, Shogren KL, Zhang M, Turner RT, Spelsberg TC, Maran A. 17β-Estradiol-dependent activation of signal transducer and activator of transcription-1 in human fetal osteoblasts is dependent on Src kinase activity. Endocrinology. 2005;146:201–207. doi: 10.1210/en.2004-0486. [DOI] [PubMed] [Google Scholar]
  • 64.Kim SY, Park HJ, Choi JS, Lee JE, Cha JH, Choi YS, Cho KO, Chun MH, Lee MY. Ischemic preconditioning-induced expression of gpl30 and STAT3 in astrocytes of the rat hippocampus. Brain Res Mol Brain Res. 2004;129:96–103. doi: 10.1016/j.molbrainres.2004.06.025. [DOI] [PubMed] [Google Scholar]
  • 65.Kim TK, Maniatis T. Regulation of interferon-gamma-activated STAT1 by the ubiquitin-proteasome pathway. Science. 1996;273:1717–1719. doi: 10.1126/science.273.5282.1717. [DOI] [PubMed] [Google Scholar]
  • 66.Kitagawa H, Warita H, Sasaki C, Zhang WR, Sakai K, Shiro Y, Mitsumoto Y, Mori T, Abe K. Immuno-reactive Akt, PI3-K and ERK protein kinase expression in ischemic rat brain. Neurosci Lett. 1999;274:45–48. doi: 10.1016/s0304-3940(99)00676-x. [DOI] [PubMed] [Google Scholar]
  • 67.Kleinschnitz C, Schroeter M, Jander S, Stoll G. Induction of granulocyte colony-stimulating factor mRNA by focal cerebral ischemia and cortical spreading depression. Brain Res Mol Brain Res. 2004;131:73–78. doi: 10.1016/j.molbrainres.2004.08.011. [DOI] [PubMed] [Google Scholar]
  • 68.Komine-Kobayashi M, Zhang N, Liu M, Tanaka R, Hara H, Osaka A, Mochizuki H, Mizuno Y, Urabe T. Neuroprotective effect of recombinant human granulocyte colony-stimulating factor in transient focal ischemia of mice. J Cereb Blood Flow Metab. 2006;26:402–413. doi: 10.1038/sj.jcbfm.9600195. [DOI] [PubMed] [Google Scholar]
  • 69.Kovarik P, Mangold M, Ramsauer K, Heidari H, Steinborn R, Zotter A, Levy DE, Muller M, Decker T. Specificity of signaling by STAT1 depends on SH2 and C-terminal domains that regulate Ser727 phosphorylation, differentially affecting specific target gene expression. EMBO J. 2001;20:91–100. doi: 10.1093/emboj/20.1.91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Krebs DL, Hilton DJ. SOCS: physiological suppressors of cytokine signaling. J Cell Sci. 2000;113:2813–2819. doi: 10.1242/jcs.113.16.2813. [DOI] [PubMed] [Google Scholar]
  • 71.Kretz A, Happold CJ, Marticke JK, Isenmann S. Erythropoietin promotes regeneration of adult CNS neurons via Jak2/Stat3 and PI3K/AKT pathway activation. Mol Cell Neurosci. 2005;29:569–579. doi: 10.1016/j.mcn.2005.04.009. [DOI] [PubMed] [Google Scholar]
  • 72.Kumar A, Commane M, Flickinger TW, Horvath CM, Stark GR. Defective TNF-alpha-induced apoptosis in STAT1-null cells due to low constitutive levels of caspases. Science. 1997;278:1630–1632. doi: 10.1126/science.278.5343.1630. [DOI] [PubMed] [Google Scholar]
  • 73.Kumral A, Genc S, Ozer E, Yilmaz O, Gokmen N, Koroglu TF, Duman N, Genc K, Ozkan H. Erythropoietin downregulates bax and DP5 proapoptotic gene expression in neonatal hypoxic-ischemic brain injury. Biol Neonate. 2006;89:205–210. doi: 10.1159/000089951. [DOI] [PubMed] [Google Scholar]
  • 74.Kushima Y, Hatanaka H. Interleukin-6 and leukemia inhibitory factor promote the survival of acetylcholinesterase-positive neurons in culture from embryonic rat spinal cord. Neurosci Lett. 1992;143:110–114. doi: 10.1016/0304-3940(92)90244-2. [DOI] [PubMed] [Google Scholar]
  • 75.Labelle C, Leclerc N. Exogenous BDNF, NT-3 and NT-4 differentially regulate neurite outgrowth in cultured hippocampal neurons. Brain Res Dev Brain Res. 2000;123:1–11. doi: 10.1016/s0165-3806(00)00069-9. [DOI] [PubMed] [Google Scholar]
  • 76.Lee KS, Park JH, Lee S, Lim HJ, Choi HE, Park HY. HB-EGF induces delayed STAT3 activation via NF-kappaB mediated IL-6 secretion in vascular smooth muscle cell. Biochim Biophys Acta. 2007;1773:1637–1644. doi: 10.1016/j.bbamcr.2007.07.001. [DOI] [PubMed] [Google Scholar]
  • 77.Lee N, Neitzel KL, Devlin BK, MacLennan AJ. STAT3 phosphorylation in injured axons before sensory and motor neuron nuclei: potential role for STAT3 as a retrograde signaling transcription factor. J Comp Neurol. 2004;474:535–545. doi: 10.1002/cne.20140. [DOI] [PubMed] [Google Scholar]
  • 78.Levin ER. Cellular functions of plasma membrane estrogen receptors. Steroids. 2002;67:471–475. doi: 10.1016/s0039-128x(01)00179-9. [DOI] [PubMed] [Google Scholar]
  • 79.Lim CP, Cao X. Structure, function, and regulation of STAT proteins. Mol Biosyst. 2006;2:536–550. doi: 10.1039/b606246f. [DOI] [PubMed] [Google Scholar]
  • 80.Lin G, Bella AJ, Lue TF, Lin CS. Brain-derived neurotrophic factor (BDNF) acts primarily via the JAK/STAT pathway to promote neurite growth in the major pelvic ganglion of the rat: part 2. J Sex Med. 2006;3:821–827. doi: 10.1111/j.1743-6109.2006.00292.x. [DOI] [PubMed] [Google Scholar]
  • 81.Lin TN, Wang PY, Chi SI, Kuo JS. Differential regulation of ciliary neurotrophic factor (CNTF) and CNTF receptor alpha (CNTFR alpha) expression following focal cerebral ischemia. Brain Res Mol Brain Res. 1998;55:71–80. doi: 10.1016/s0169-328x(97)00367-7. [DOI] [PubMed] [Google Scholar]
  • 82.Loddick SA, Turnbull AV, Rothwell NJ. Cerebral interleukin-6 is neuroprotective during permanent focal cerebral ischemia in the rat. J Cereb Blood Flow Metab. 1998;18:176–179. doi: 10.1097/00004647-199802000-00008. [DOI] [PubMed] [Google Scholar]
  • 83.Lois C, Alvarez-Buylla A. Proliferating subventricular zone cells in the adult mammalian forebrain can differentiate into neurons and glia. Proc Natl Acad Sci U S A. 1993;90:2074–2077. doi: 10.1073/pnas.90.5.2074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Lufei C, Ma J, Huang G, Zhang T, Novotny-Diermayr V, Ong CT, Cao X. GRIM-19, a death-regulatory gene product, suppresses Stat3 activity via functional interaction. EMBO J. 2003;22:1325–1335. doi: 10.1093/emboj/cdg135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Luo JM, Cen LP, Zhang XM, Chiang SW, Huang Y, Lin D, Fan YM, van RN, Lam DS, Pang CP, Cui Q. PI3K/akt, JAK/STAT and MEK/ERK pathway inhibition protects retinal ganglion cells via different mechanisms after optic nerve injury. Eur J Neurosci. 2007;26:828–842. doi: 10.1111/j.1460-9568.2007.05718.x. [DOI] [PubMed] [Google Scholar]
  • 86.Lutticken C, Wegenka UM, Yuan J, Buschmann J, Schindler C, Ziemiecki A, Harpur AG, Wilks AF, Yasukawa K, Taga T. Association of transcription factor APRF and protein kinase Jakl with the interleukin-6 signal transducer gpl30. Science. 1994;263:89–92. doi: 10.1126/science.8272872. [DOI] [PubMed] [Google Scholar]
  • 87.MacLaren RE, Buch PK, Smith AJ, Balaggan KS, MacNeil A, Taylor JS, Osborne NN, Ali RR. CNTF gene transfer protects ganglion cells in rat retinae undergoing focal injury and branch vessel occlusion. Exp Eye Res. 2006;83:1118–1127. doi: 10.1016/j.exer.2006.05.019. [DOI] [PubMed] [Google Scholar]
  • 88.Merry DE, Korsmeyer SJ. Bcl-2 gene family in the nervous system. Annu Rev Neurosci. 1997;20:245–267. doi: 10.1146/annurev.neuro.20.1.245. [DOI] [PubMed] [Google Scholar]
  • 89.Merry DE, Veis DJ, Hickey WF, Korsmeyer SJ. bcl-2 protein expression is widespread in the developing nervous system and retained in the adult PNS. Development. 1994;120:301–311. doi: 10.1242/dev.120.2.301. [DOI] [PubMed] [Google Scholar]
  • 90.Morishita E, Masuda S, Nagao M, Yasuda Y, Sasaki R. Erythropoietin receptor is expressed in rat hippocampal and cerebral cortical neurons, and erythropoietin prevents in vitro glutamate-induced neuronal death. Neuroscience. 1997;76:105–116. doi: 10.1016/s0306-4522(96)00306-5. [DOI] [PubMed] [Google Scholar]
  • 91.Murata S, Usuda N, Okano A, Kobayashi S, Suzuki T. Occurrence of a transcription factor, signal transducer and activators of transcription 3 (Stat3), in the postsynaptic density of the rat brain. Brain Res Mol Brain Res. 2000;78:80–90. doi: 10.1016/s0169-328x(00)00077-2. [DOI] [PubMed] [Google Scholar]
  • 92.Muta K, Krantz SB, Bondurant MC, Wickrema A. Distinct roles of erythropoietin, insulin-like growth factor I, and stem cell factor in the development of erythroid progenitor cells. J Clin Invest. 1994;94:34–43. doi: 10.1172/JCI117327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Nakatomi H, Kuriu T, Okabe S, Yamamoto S, Hatano O, Kawahara N, Tamura A, Kirino T, Nakafuku M. Regeneration of hippocampal pyramidal neurons after ischemic brain injury by recruitment of endogenous neural progenitors. Cell. 2002;110:429–441. doi: 10.1016/s0092-8674(02)00862-0. [DOI] [PubMed] [Google Scholar]
  • 94.Nedergaard M, Dirnagl U. Role of glial cells in cerebral ischemia. Glia. 2005;50:281–286. doi: 10.1002/glia.20205. [DOI] [PubMed] [Google Scholar]
  • 95.Ng YP, Cheung ZH, Ip NY. STAT3 as a downstream mediator of Trk signaling and functions. J Biol Chem. 2006;281:15636–15644. doi: 10.1074/jbc.M601863200. [DOI] [PubMed] [Google Scholar]
  • 96.O'Brien JJ, Nathanson NM. Retrograde activation of STAT3 by leukemia inhibitory factor in sympathetic neurons. J Neurochem. 2007;103:288–302. doi: 10.1111/j.1471-4159.2007.04736.x. [DOI] [PubMed] [Google Scholar]
  • 97.Oestreicher AB, de Graan PN, Gispen WH, Verhaagen J, Schrama LH. B-50, the growth associated protein-43: modulation of cell morphology and communication in the nervous system. Prog Neurobiol. 1997;53:627–686. doi: 10.1016/s0301-0082(97)00043-9. [DOI] [PubMed] [Google Scholar]
  • 98.Okada S, Nakamura M, Katoh H, Miyao T, Shimazaki T, Ishii K, Yamane J, Yoshimura A, Iwamoto Y, Toyama Y, Okano H. Conditional ablation of Stat3 or Socs3 discloses a dual role for reactive astrocytes after spinal cord injury. Nat Med. 2006;12:829–834. doi: 10.1038/nm1425. [DOI] [PubMed] [Google Scholar]
  • 99.Planas AM, Berruezo M, Justicia C, Barron S, Ferrer I. Stat3 is present in the developing and adult rat cerebellum and participates in the formation of transcription complexes binding DNA at the sis-inducible element. J Neurochem. 1997;68:1345–1351. doi: 10.1046/j.1471-4159.1997.68041345.x. [DOI] [PubMed] [Google Scholar]
  • 100.Planas AM, Justicia C, Ferrer I. Stat1 in developing and adult rat brain. Induction after transient focal ischemia. NeuroReport. 1997;8:1359–1362. doi: 10.1097/00001756-199704140-00008. [DOI] [PubMed] [Google Scholar]
  • 101.Planas AM, Soriano MA, Berruezo M, Justicia C, Estrada A, Pitarch S, Ferrer I. Induction of Stat3, a signal transducer and transcription factor, in reactive microglia following transient focal cerebral ischaemia. Eur J Neurosci. 1996;8:2612–2618. doi: 10.1111/j.1460-9568.1996.tb01556.x. [DOI] [PubMed] [Google Scholar]
  • 102.Platzer E, Simon S, Kalden JR. Human granulocyte colony stimulating factor: effects on human long-term bone marrow cultures. Blood Cells. 1988;14:463–469. [PubMed] [Google Scholar]
  • 103.Poveshchenko AF, Filimonov PN, Abramov VV, Korotkova NA, Iakushenko EV, Kozlov VA. Expression of erythropoietin receptor mRNA in mouse brain hemispheres. Tsitologiia. 2001;43:279–283. [PubMed] [Google Scholar]
  • 104.Qiu J, Bosch MA, Tobias SC, Grandy DK, Scanlan TS, Ronnekleiv OK, Kelly MJ. Rapid signaling of estrogen in hypothalamic neurons involves a novel G-protein-coupled estrogen receptor that activates protein kinase C. J Neurosci. 2003;23:9529–9540. doi: 10.1523/JNEUROSCI.23-29-09529.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Qiu J, Cafferty WB, McMahon SB, Thompson SW. Conditioning injury-induced spinal axon regeneration requires signal transducer and activator of transcription 3 activation. J Neurosci. 2005;25:1645–1653. doi: 10.1523/JNEUROSCI.3269-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Rahaman SO, Vogelbaum MA, Haque SJ. Aberrant Stat3 signaling by interleukin-4 in malignant glioma cells: involvement of IL-13Rα2. Cancer Res. 2005;65:2956–2963. doi: 10.1158/0008-5472.CAN-04-3592. [DOI] [PubMed] [Google Scholar]
  • 107.Rajan P, McKay RD. Multiple routes to astrocytic differentiation in the CNS. J Neurosci. 1998;18:3620–3629. doi: 10.1523/JNEUROSCI.18-10-03620.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Rajan P, Stewart CL, Fink JS. LIF-mediated activation of STAT proteins after neuronal injury in vivo. NeuroReport. 1995;6:2240–2244. doi: 10.1097/00001756-199511000-00033. [DOI] [PubMed] [Google Scholar]
  • 109.Reich NC, Liu L. Tracking STAT nuclear traffic. Nat Rev Immunol. 2006;6:602–612. doi: 10.1038/nri1885. [DOI] [PubMed] [Google Scholar]
  • 110.Rusa R, Alkayed NJ, Crain BJ, Traystman RJ, Kimes AS, London ED, Klaus JA, Hurn PD. 17β-Estradiol reduces stroke injury in estrogen-deficient female animals. Stroke. 1999;30:1665–1670. doi: 10.1161/01.str.30.8.1665. [DOI] [PubMed] [Google Scholar]
  • 111.Sadamoto Y, Igase K, Sakanaka M, Sato K, Otsuka H, Sakaki S, Masuda S, Sasaki R. Erythropoietin prevents place navigation disability and cortical infarction in rats with permanent occlusion of the middle cerebral artery. Biochem Biophys Res Commun. 1998;253:26–32. doi: 10.1006/bbrc.1998.9748. [DOI] [PubMed] [Google Scholar]
  • 112.Satriotomo I, Bowen KK, Vemuganti R. JAK2 and STAT3 activation contributes to neuronal damage following transient focal cerebral ischemia. J Neurochem. 2006;98:1353–1368. doi: 10.1111/j.1471-4159.2006.04051.x. [DOI] [PubMed] [Google Scholar]
  • 113.Schabitz WR, Berger C, Kollmar R, Seitz M, Tanay E, Kiessling M, Schwab S, Sommer C. Effect of brain-derived neurotrophic factor treatment and forced arm use on functional motor recovery after small cortical ischemia. Stroke. 2004;35:992–997. doi: 10.1161/01.STR.0000119754.85848.0D. [DOI] [PubMed] [Google Scholar]
  • 114.Schabitz WR, Kollmar R, Schwaninger M, Juettler E, Bardutzky J, Scholzke MN, Sommer C, Schwab S. Neuroprotective effect of granulocyte colony-stimulating factor after focal cerebral ischemia. Stroke. 2003;34:745–751. doi: 10.1161/01.STR.0000057814.70180.17. [DOI] [PubMed] [Google Scholar]
  • 115.Schabitz WR, Schneider A. Developing granulocyte-colony stimulating factor for the treatment of stroke: current status of clinical trials. Stroke. 2006;37:1654. doi: 10.1161/01.STR.0000227299.62106.0e. [DOI] [PubMed] [Google Scholar]
  • 116.Schaper F, Gendo C, Eck M, Schmitz J, Grimm C, Anhuf D, Kerr IM, Heinrich PC. Activation of the protein tyrosine phosphatase SHP2 via the interleukin-6 signal transducing receptor protein gp130 requires tyrosine kinase Jak1 and limits acute-phase protein expression. Biochem J. 1998;335:557–565. doi: 10.1042/bj3350557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Schicho R, Schuligoi R, Sirinathsinghji DJ, Donnerer J. Increased expression of GAP-43 in small sensory neurons after stimulation by NGF indicative of neuroregeneration in capsaicin-treated rats. Regul Pept. 1999;83:87–95. doi: 10.1016/s0167-0115(99)00051-8. [DOI] [PubMed] [Google Scholar]
  • 118.Schicho R, Skofitsch G, Donnerer J. Regenerative effect of human recombinant NGF on capsaicin-lesioned sensory neurons in the adult rat. Brain Res. 1999;815:60–69. doi: 10.1016/s0006-8993(98)01094-4. [DOI] [PubMed] [Google Scholar]
  • 119.Schindler C, Darnell JE., Jr Transcriptional responses to polypeptide ligands: the JAK-STAT pathway. Annu Rev Biochem. 1995;64:621–651. doi: 10.1146/annurev.bi.64.070195.003201. [DOI] [PubMed] [Google Scholar]
  • 120.Schindler U, Wu P, Rothe M, Brasseur M, McKnight SL. Components of a Stat recognition code: evidence for two layers of molecular selectivity. Immunity. 1995;2:689–697. doi: 10.1016/1074-7613(95)90013-6. [DOI] [PubMed] [Google Scholar]
  • 121.Schneider A, Kruger C, Steigleder T, Weber D, Pitzer C, Laage R, Aronowski J, Maurer MH, Gassler N, Mier W, Hasselblatt M, Kollmar R, Schwab S, Sommer C, Bach A, Kuhn HG, Schabitz WR. The hematopoietic factor G-CSF is a neuronal ligand that counteracts programmed cell death and drives neurogenesis. J Clin Invest. 2005;115:2083–2098. doi: 10.1172/JCI23559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Schwaiger FW, Hager G, Schmitt AB, Horvat A, Hager G, Streif R, Spitzer C, Gamal S, Breuer S, Brook GA, Nacimiento W, Kreutzberg GW. Peripheral but not central axotomy induces changes in Janus kinases (JAK) and signal transducers and activators of transcription (STAT) Eur J Neurosci. 2000;12:1165–1176. doi: 10.1046/j.1460-9568.2000.00005.x. [DOI] [PubMed] [Google Scholar]
  • 123.Schweizer U, Gunnersen J, Karch C, Wiese S, Holtmann B, Takeda K, Akira S, Sendtner M. Conditional gene ablation of Stat3 reveals differential signaling requirements for survival of motoneurons during development and after nerve injury in the adult. J Cell Biol. 2002;156:287–297. doi: 10.1083/jcb.200107009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Sehgal PB, Kumar V, Guo G, Murray WC. Different patterns of regulation of Tyr-phosphorylated STAT1 and STAT3 in human hepatoma Hep3B cells by the phosphatase inhibitor orthovanadate. Arch Biochem Biophys. 2003;412:242–250. doi: 10.1016/s0003-9861(03)00050-x. [DOI] [PubMed] [Google Scholar]
  • 125.Seidel HM, Milocco LH, Lamb P, Darnell JE, Jr, Stein RB, Rosen J. Spacing of palindromic half sites as a determinant of selective STAT (signal transducers and activators of transcription) DNA binding and transcriptional activity. Proc Natl Acad Sci USA. 1995;92:3041–3045. doi: 10.1073/pnas.92.7.3041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Sheu JY, Kulhanek DJ, Eckenstein FP. Differential patterns of ERK and STAT3 phosphorylation after sciatic nerve transection in the rat. Exp Neurol. 2000;166:392–402. doi: 10.1006/exnr.2000.7508. [DOI] [PubMed] [Google Scholar]
  • 127.Shirogane T, Fukada T, Muller JM, Shima DT, Hibi M, Hirano T. Synergistic roles for Pim-1 and c-Myc in STAT3-mediated cell cycle progression and antiapoptosis. Immunity. 1999;11:709–719. doi: 10.1016/s1074-7613(00)80145-4. [DOI] [PubMed] [Google Scholar]
  • 128.Silva M, Grillot D, Benito A, Richard C, Nunez G, Fernandez-Luna JL. Erythropoietin can promote erythroid progenitor survival by repressing apoptosis through Bcl-XL and Bcl-2. Blood. 1996;88:1576–1582. [PubMed] [Google Scholar]
  • 129.Singer CA, Figueroa-Masot XA, Batchelor RH, Dorsa DM. The mitogen-activated protein kinase pathway mediates estrogen neuroprotection after glutamate toxicity in primary cortical neurons. J Neurosci. 1999;19:2455–2463. doi: 10.1523/JNEUROSCI.19-07-02455.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Siren AL, Knerlich F, Poser W, Gleiter CH, Bruck W, Ehrenreich H. Erythropoietin and erythropoietin receptor in human ischemic/hypoxic brain. Acta Neuropathol. 2001;101:271–276. doi: 10.1007/s004010000297. [DOI] [PubMed] [Google Scholar]
  • 131.Sola A, Rogido M, Lee BH, Genetta T, Wen TC. Erythropoietin after focal cerebral ischemia activates the Janus kinase-signal transducer and activator of transcription signaling pathway and improves brain injury in postnatal day 7 rats. Pediatr Res. 2005;57:481–487. doi: 10.1203/01.PDR.0000155760.88664.06. [DOI] [PubMed] [Google Scholar]
  • 132.Solaroglu I, Cahill J, Jadhav V, Zhang JH. A novel neuroprotectant granulocyte-colony stimulating factor. Stroke. 2006;37:1123–1128. doi: 10.1161/01.STR.0000208205.26253.96. [DOI] [PubMed] [Google Scholar]
  • 133.Solaroglu I, Solaroglu A, Kaptanoglu E, Dede S, Haberal A, Beskonakli E, Kilinc K. Erythropoietin prevents ischemia-reperfusion from inducing oxidative damage in fetal rat brain. Childs Nerv Syst. 2003;19:19–22. doi: 10.1007/s00381-002-0680-2. [DOI] [PubMed] [Google Scholar]
  • 134.Solaroglu I, Tsubokawa T, Cahill J, Zhang JH. Anti-apoptotic effect of granulocyte-colony stimulating factor after focal cerebral ischemia in the rat. Neuroscience. 2006;143:965–974. doi: 10.1016/j.neuroscience.2006.09.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Stahl, Davis S, Wong V, Taga T, Kishimoto T, Ip NY, Yancopoulos GD. Cross-linking identifies leukemia inhibitory factor-binding protein as a ciliary neurotrophic factor receptor component. J Biol Chem. 1993;268:7628–7631. [PubMed] [Google Scholar]
  • 136.Stahl N, Farruggella TJ, Boulton TG, Zhong Z, Darnell JE, Jr, Yancopoulos GD. Choice of STATs and other substrates specified by modular tyrosine-based motifs in cytokine receptors. Science. 1995;267:1349–1353. doi: 10.1126/science.7871433. [DOI] [PubMed] [Google Scholar]
  • 137.Stephanou A, Brar BK, Knight RA, Latchman DS. Opposing actions of STAT-1 and STAT-3 on the Bcl-2 and Bcl-x promoters. Cell Death Differ. 2000;7:329–330. doi: 10.1038/sj.cdd.4400656. [DOI] [PubMed] [Google Scholar]
  • 138.Stromberg H, Svensson SP, Hermanson O. Distribution of the transcription factor signal transducer and activator of transcription 3 in the rat central nervous system and dorsal root ganglia. Brain Res. 2000;853:105–114. doi: 10.1016/s0006-8993(99)02260-x. [DOI] [PubMed] [Google Scholar]
  • 139.Suzuki S, Tanaka K, Nagata E, Ito D, Dembo T, Fukuuchi Y. Cerebral neurons express interleukin-6 after transient forebrain ischemia in gerbils. Neurosci Lett. 1999;262:117–120. doi: 10.1016/s0304-3940(99)00051-8. [DOI] [PubMed] [Google Scholar]
  • 140.Suzuki S, Tanaka K, Nogawa S, Dembo T, Kosakai A, Fukuuchi Y. Phosphorylation of signal transducer and activator of transcription-3 (Stat3) after focal cerebral ischemia in rats. Exp Neurol. 2001;170:63–71. doi: 10.1006/exnr.2001.7701. [DOI] [PubMed] [Google Scholar]
  • 141.Suzuki S, Tanaka K, Nogawa S, Ito D, Dembo T, Kosakai A, Fukuuchi Y. Immunohistochemical detection of leukemia inhibitory factor after focal cerebral ischemia in rats. J Cereb Blood Flow Metab. 2000;20:661–668. doi: 10.1097/00004647-200004000-00003. [DOI] [PubMed] [Google Scholar]
  • 142.Suzuki S, Tanaka K, Nogawa S, Nagata E, Ito D, Dembo T, Fukuuchi Y. Temporal profile and cellular localization of interleukin-6 protein after focal cerebral ischemia in rats. J Cereb Blood Flow Metab. 1999;19:1256–1262. doi: 10.1097/00004647-199911000-00010. [DOI] [PubMed] [Google Scholar]
  • 143.Suzuki S, Yamashita T, Tanaka K, Hattori H, Sawamoto K, Okano H, Suzuki N. Activation of cytokine signaling through leukemia inhibitory factor receptor (LIFR)/gpl30 attenuates ischemic brain injury in rats. J Cereb Blood Flow Metab. 2005;25:685–693. doi: 10.1038/sj.jcbfm.9600061. [DOI] [PubMed] [Google Scholar]
  • 144.Taga T, Kishimoto T. Signaling mechanisms through cytokine receptors that share signal transducing receptor components. Curr Opin Immunol. 1995;7:17–23. doi: 10.1016/0952-7915(95)80024-7. [DOI] [PubMed] [Google Scholar]
  • 145.Takagi Y, Harada J, Chiarugi A, Moskowitz MA. STAT1 is activated in neurons after ischemia and contributes to ischemic brain injury. J Cereb Blood Flow Metab. 2002;22:1311–1318. doi: 10.1097/01.WCB.0000034148.72481.F4. [DOI] [PubMed] [Google Scholar]
  • 146.Takanaga H, Yoshitake T, Hara S, Yamasaki C, Kunimoto M. cAMP-induced astrocytic differentiation of C6 glioma cells is mediated by autocrine interleukin-6. J Biol Chem. 2004;279:15441–15447. doi: 10.1074/jbc.M311844200. [DOI] [PubMed] [Google Scholar]
  • 147.Takeda K, Noguchi K, Shi W, Tanaka T, Matsumoto M, Yoshida N, Kishimoto T, Akira S. Targeted disruption of the mouse Stat3 gene leads to early embryonic lethality. Proc Natl Acad Sci USA. 1997;94:3801–3804. doi: 10.1073/pnas.94.8.3801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Takizawa T, Yanagisawa M, Ochiai W, Yasukawa K, Ishiguro T, Nakashima K, Taga T. Directly linked soluble IL-6 receptor-IL-6 fusion protein induces astrocyte differentiation from neuroepithelial cells via activation of STAT3. Cytokine. 2001;13:272–279. doi: 10.1006/cyto.2000.0831. [DOI] [PubMed] [Google Scholar]
  • 149.ten HJ, de J, I, Fu Y, Zhu W, Tremblay M, David M, Shuai K. Identification of a nuclear Stat1 protein tyrosine phosphatase. Mol Cell Biol. 2002;22:5662–5668. doi: 10.1128/MCB.22.16.5662-5668.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Teng YD, Lavik EB, Qu X, Park KI, Ourednik J, Zurakowski D, Langer R, Snyder EY. Functional recovery following traumatic spinal cord injury mediated by a unique polymer scaffold seeded with neural stem cells. Proc Natl Acad Sci USA. 2002;99:3024–3029. doi: 10.1073/pnas.052678899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Tolwani RJ, Buckmaster PS, Varma S, Cosgaya JM, Wu Y, Suri C, Shooter EM. BDNF overexpression increases dendrite complexity in hippocampal dentate gyrus. Neuroscience. 2002;114:795–805. doi: 10.1016/s0306-4522(02)00301-9. [DOI] [PubMed] [Google Scholar]
  • 152.Tsai SY, Yang LY, Wu CH, Chang SF, Hsu CY, Wei CP, Leu SJ, Liaw J, Lee YH, Tsai MD. Injury-induced Janus kinase/protein kinase C-dependent phosphorylation of growth-associated protein 43 and signal transducer and activator of transcription 3 for neurite growth in dorsal root ganglion. J Neurosci Res. 2007;85:321–331. doi: 10.1002/jnr.21119. [DOI] [PubMed] [Google Scholar]
  • 153.Uemura A, Takizawa T, Ochiai W, Yanagisawa M, Nakashima K, Taga T. Cardiotrophin-like cytokine induces astrocyte differentiation of fetal neuroepithelial cells via activation of STAT3. Cytokine. 2002;18:1–7. doi: 10.1006/cyto.2002.1006. [DOI] [PubMed] [Google Scholar]
  • 154.Vignais ML, Sadowski HB, Watling D, Rogers NC, Gilman M. Platelet-derived growth factor induces phosphorylation of multiple JAK family kinases and STAT proteins. Mol Cell Biol. 1996;16:1759–1769. doi: 10.1128/mcb.16.4.1759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Vinkemeier U, Moarefi I, Darnell JE, Jr, Kuriyan J. Structure of the amino-terminal protein interaction domain of STAT-4. Science. 1998;279:1048–1052. doi: 10.1126/science.279.5353.1048. [DOI] [PubMed] [Google Scholar]
  • 156.Wang J, Pham-Mitchell N, Schindler C, Campbell IL. Dysregulated sonic hedgehog signaling and medulloblastoma consequent to IFN-alpha-stimulated STAT2-independent production of IFN-gamma in the brain. J Clin Invest. 2003;112:535–543. doi: 10.1172/JCI18637. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Wang J, Schreiber RD, Campbell IL. STAT1 deficiency unexpectedly and markedly exacerbates the pathophysiological actions of IFN-alpha in the central nervous system. Proc Natl Acad Sci USA. 2002;99:16209–16214. doi: 10.1073/pnas.252454799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Wang M, Zhang W, Crisostomo P, Markel T, Meldrum KK, Fu XY, Meldrum DR. Endothelial STAT3 plays a critical role in generalized myocardial proinflammatory and proapoptotic signaling. Am J Physiol Heart Circ Physiol. 2007;293:H2101–H2108. doi: 10.1152/ajpheart.00125.2007. [DOI] [PubMed] [Google Scholar]
  • 159.Wen TC, Peng H, Hata R, Desaki J, Sakanaka M. Induction of phosphorylated-Stat3 following focal cerebral ischemia in mice. Neurosci Lett. 2001;303:153–156. doi: 10.1016/s0304-3940(01)01711-6. [DOI] [PubMed] [Google Scholar]
  • 160.West DA, Valentim LM, Lythgoe MF, Stephanou A, Proctor E, van der Weerd L, Ordidge RJ, Latchman DS, Gadian DG. MR image-guided investigation of regional signal transducers and activators of transcription-1 activation in a rat model of focal cerebral ischemia. Neuroscience. 2004;127:333–339. doi: 10.1016/j.neuroscience.2004.05.022. [DOI] [PubMed] [Google Scholar]
  • 161.Wu TR, Hong YK, Wang XD, Ling MY, Dragoi AM, Chung AS, Campbell AG, Han ZY, Feng GS, Chin YE. SHP-2 is a dual-specificity phosphatase involved in Stat1 dephosphorylation at both tyrosine and serine residues in nuclei. J Biol Chem. 2002;277:47572–47580. doi: 10.1074/jbc.M207536200. [DOI] [PubMed] [Google Scholar]
  • 162.Wu YY, Bradshaw RA. Induction of neurite outgrowth by interleukin-6 is accompanied by activation of Stat3 signaling pathway in a variant PC 12 cell (E2) line. J Biol Chem. 1996;271:13023–13032. doi: 10.1074/jbc.271.22.13023. [DOI] [PubMed] [Google Scholar]
  • 163.Xia XG, Hofmann HD, Deller T, Kirsch M. Induction of STAT3 signaling in activated astrocytes and sprouting septal neurons following entorhinal cortex lesion in adult rats. Mol Cell Neurosci. 2002;21:379–392. doi: 10.1006/mcne.2002.1180. [DOI] [PubMed] [Google Scholar]
  • 164.Yadav A, Kalita A, Dhillon S, Banerjee K. JAK/STAT3 pathway is involved in survival of neurons in response to insulin-like growth factor and negatively regulated by suppressor of cytokine signaling-3. J Biol Chem. 2005;280:31830–31840. doi: 10.1074/jbc.M501316200. [DOI] [PubMed] [Google Scholar]
  • 165.Yamashita K, Wiessner C, Lindholm D, Thoenen H, Hossmann KA. Post-occlusion treatment with BDNF reduces infarct size in a model of permanent occlusion of the middle cerebral artery in rat. Metab Brain Dis. 1997;12:271–280. doi: 10.1007/BF02674671. [DOI] [PubMed] [Google Scholar]
  • 166.Yamashita T, Sawamoto K, Suzuki S, Suzuki N, Adachi K, Kawase T, Mihara M, Ohsugi Y, Abe K, Okano H. Blockade of interleukin-6 signaling aggravates ischemic cerebral damage in mice: possible involvement of Stat3 activation in the protection of neurons. J Neurochem. 2005;94:459–468. doi: 10.1111/j.1471-4159.2005.03227.x. [DOI] [PubMed] [Google Scholar]
  • 167.Yamauchi K, Osuka K, Takayasu M, Usuda N, Nakazawa A, Nakahara N, Yoshida M, Aoshima C, Hara M, Yoshida J. Activation of JAK/STAT signalling in neurons following spinal cord injury in mice. J Neurochem. 2006;96:1060–1070. doi: 10.1111/j.1471-4159.2005.03559.x. [DOI] [PubMed] [Google Scholar]
  • 168.Yanagisawa M, Nakashima K, Taga T. STAT3-mediated astrocyte differentiation from mouse fetal neuroepithelial cells by mouse oncostatin M. Neurosci Lett. 1999;269:169–172. doi: 10.1016/s0304-3940(99)00447-4. [DOI] [PubMed] [Google Scholar]
  • 169.Yi JH, Park SW, Kapadia R, Vemuganti R. Role of transcription factors in mediating post-ischemic cerebral inflammation and brain damage. Neurochem Int. 2007;50:1014–1027. doi: 10.1016/j.neuint.2007.04.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Zhang F, Wang S, Cao G, Gao Y, Chen J. Signal transducers and activators of transcription 5 contributes to erythropoietin-mediated neuroprotection against hippocampal neuronal death after transient global cerebral ischemia. Neurobiol Dis. 2007;25:45–53. doi: 10.1016/j.nbd.2006.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Zhong Z, Wen Z, Darnell JE., Jr Stat3: a STAT family member activated by tyrosine phosphorylation in response to epidermal growth factor and interleukin-6. Science. 1994;264:95–98. doi: 10.1126/science.8140422. [DOI] [PubMed] [Google Scholar]
  • 172.Zigmond RE, Hyatt-Sachs H, Mohney RP, Schreiber RC, Shadiack AM, Sun Y, Vaccariello SA. Changes in neuropeptide phenotype after axotomy of adult peripheral neurons and the role of leukemia inhibitory factor. Perspect Dev Neurobiol. 1996;4:75–90. [PubMed] [Google Scholar]

RESOURCES