Skip to main content
  • Review Article
  • Published:

STAT Signaling in the Pathogenesis and Treatment of Cancer

Abstract

Exceptional advances have been made recently in our understanding of the signaling pathways that control cellular growth, differentiation, and survival. These processes are regulated by extracellular stimuli such as cytokines, cell-cell interactions, and cell-matrix interactions, which trigger a series of intracellular events culminating in the modulation of specific genes. STATs are a highly homologous group of transcription factors that are activated by various pathways and regulate many of the genes controlling cellular function. STATs are activated by tyrosine phosphorylation and modulated by serine phosphorylation, placing them at a convergence point for numerous intracellular signaling pathways. Given the importance of STATs in the control of normal physiologic processes, it is not surprising that inappropriate activation of these proteins has been found in human malignancies. A number of distinct mechanisms have been elucidated by which STATs are activated inappropriately, including autocrine or paracrine stimulation of normal receptors and increased activity of tyrosine kinases through enhanced expression, mutations, or the presence of activating proteins. Furthermore, inappropriate STAT serine phosphorylation has been found in several tumors as well. The increased understanding of signaling pathways in tumors can be translated into therapeutic strategies that have the potential to be more selective and less toxic than current anti-cancer treatments. Approaches which may be effective include the development of antagonists of receptors that can trigger STAT activation, inhibitors of the tyrosine and serine kinases that phosphorylate and activate STATs, agents that decrease STAT levels or inhibit their recruitment to kinases, and molecules that can prevent the binding of STATs to target DNA sequences. Thus, elucidation of cellular and biochemical processes in tumors has enhanced our understanding of the pathogenesis of malignancies and may provide the basis for significant advances in therapy.

Introduction

At the most basic level, cancer represents an abnormal accumulation of cells that arise from a selective advantage in one of three processes: (1) an enhanced ability to grow; (2) a defect in undergoing cell death; or (3) a loss of the capacity to differentiate. Each of these processes is normally initiated by extracellular stimuli such as soluble factors, cell-cell interactions, or cell-matrix interactions, although the actual phenotypic program is determined by the pattern of genes that are activated or repressed. A variety of signaling cascades are present in the cell that link events at the cell membrane with programs of gene expression in the nucleus. In recent years a number of lines of evidence have suggested that pathways that lead to the activation of a group of transcription factors termed STATs may play a particularly important role in the biology of both hematopoietic and nonhematopoietic tumors.

STAT Signaling Pathways

Two approaches can be used to dissect the intracellular events that connect the engagement of a cell surface receptor with the modulation of transcription in the nucleus. The first is to study the receptor, and to discern modifications, such as phosphorylation, that occur following binding by a ligand. Through biochemical and genetic techniques, one can then move step-wise to elucidate proteins that are activated, ultimately arriving at the transcription factors that directly modulate gene expression. The second approach is to start in the nucleus with genes known to be activated by a specific stimulus. One can then define DNA sequences in the promoter of the gene that confer sensitivity to the stimulus. The transcription factors that bind to these DNA sequences can be identified, and the kinases or other enzymes that modulate the transcription factors can then be isolated. In this way, one can work from the inside of the cell outward to discern the pathway controlling new gene transcription. In the last several years, a number of pioneering experiments examining the signal transduction pathways utilized by interferons (IFNs) has employed both of these approaches to define a signaling cascade termed the STAT pathway, which has become important in the understanding of both normal and malignant cell growth.

In the late 1980s, a number of genes were identified whose transcription was induced by IFN-α or IFN-γ. Specific DNA elements, termed the IFN stimulation response element (ISRE) for IFN-α (1) and the gamma (IFN) activated site (GAS) for IFN-γ (2), were identified which mediate responsiveness to each IFN. Using ISRE and GAS DNA sequences as probes, proteins that bound to these sites were isolated and their genes identified (3). The proteins activated in response to IFN-α and IFN-γ are in the cytoplasm under basal conditions, though following stimulation with an IFN, they rapidly translocate to the nucleus and bind to their cognate DNA sequences. It was subsequently shown that the activation of these factors requires tyrosine phosphorylation (4).

These elegant studies defining the activation of IFN-responsive transcription factors were complemented by experiments performed at about the same time on events occurring at the cell membrane. Although it was known that the IFN receptors become tyrosine phosphorylated following stimulation with their ligand, and that inhibitors of tyrosine kinases block the action of IFNs, in contrast to polypeptide growth factor receptors [such as the receptors for epidermal growth factor (EGF), platelet-derived growth factor (PDGF), and insulin], the IFN receptors themselves lack tyrosine kinase activity. To determine how the IFN receptors activate tyrosine kinases, a series of mutants that lacked the ability to respond to IFNs was employed. Genes were transfected into these cell lines, and the restoration of sensitivity to IFNs was examined. These experiments revealed that three members of a recently described family of cytoplasmic tyrosine kinases, termed Jaks, associated with the cytoplasmic domains of these receptors and could restore their signaling ability (59). As a result of these experiments, a unified model for interferon signaling was proposed that is now known to mediate signals generated by a multitude of cytokines (Fig. 1). When a cytokine binds to its receptor, it causes the receptor chains to aggregate, bringing the associated Jak family kinases into juxtaposition. This activates the Jak kinases which mediate the subsequent tyrosine phosphorylation of the Jaks and the cytokine receptor chains. The highly tyrosine phosphorylated receptor-kinase complex then serves as a docking site for proteins, such as STATs, which possess src-homology-2 (SH2) domains that allow binding to specific tyrosine-phosphorylated amino acid sequences (10,11). The STATs recruited in this way become phosphorylated on unique tyrosine residues necessary for activation (12), then dissociate from the receptor-kinase complex and dimerize via reciprocal phosphotyrosine-SH2 interactions (13). The STAT dimers translocate to the nucleus where they bind to the DNA sequences necessary to mediate gene activation in response to the cytokine (14). The actual induction of transcription appears to require interactions between STATs and the p300/CREB binding protein (CBP) family of co-activators (1518), and requires their histone acetyltransferase activity (19).

Fig. 1
figure 1

Cytokine-induced STAT activation. (A) STATs are latent transcription factors found in the cytoplasm of cells. (B) When a cytokine interacts with its cell surface receptor, it induces receptor dimerization, bringing into juxtaposition associated Jak family tyrosine kinases. The Jaks become activated, leading to the tyrosine phosphorylation of themselves and the associated receptor chains. STATs are recruited to the activated kinase-receptor complex, where they become phosphorylated on unique tyrosine residues. (C) Once tyrosine is phosphorylated, the STATs dimerize, translocate to the nucleus, and bind to specific DNA sequences in the regulatory regions of target genes where they can modulate transcription. (D) STATs can also be phosphorylated on specific serine residues. Although this is not sufficient to activate the STATs, it leads to an enhancement of the transcriptional response mediated by tyrosine-phosphorylated STATs.

The activation of STATs is both rapid and transient. Following IFN stimulation, STAT1 becomes maximally tyrosine phosphorylated in 15 to 30 min, and returns to its basal unphosphorylated state in 1 to 2 hr. The de-activation of STATs may occur through dephosphorylation (20,21) and/or proteolysis (2224). In addition, cytokines can induce inhibitors of Jak family kinases or of the STATs themselves, and this may contribute to the transient nature of STAT activation under physiologic conditions (2530).

Although this signaling cascade was initially elucidated in the context of IFNs, it rapidly became clear that many hematopoietic cytokines whose receptors lacked intrinsic tyrosine kinase activity could interact with Jak family members and transduce signals by virtue of STAT activation. Such cytokines include interleukin-2 (IL-2) (3135), IL-3 (36), IL-4 (37,38), IL-6 (39,40), IL-12 (41), leukemia inhibitory factor (4246), erythropoietin (36), thrombopoietin (4749), colony-stimulating factor-1 (CSF-1) (50,51), granulocyte-CSF (G-CSF) (52,53), and granulocyte-macrophage-CSF (GM-CSF) (36). In addition, receptors for other soluble factors were also found to activate Jaks and STATs, such as growth hormone (54,55), oncostatin M (56), prolactin (5759), ciliary neurotrophic factor (60), tumor necrosis factor (TNF) (61), and angiotensin II (62,63). Given the importance of the Jaks in mediating the tyrosine phosphorylation of STATs, this pathway was initially referred to as the Jak-STAT pathway. However, it is clear that other kinases can participate in the phosphorylation of STATs, including src family kinases (64), Tec family kinases (65,66), and the EGF, insulin, and PDGF receptors (50,51,6772). In addition, other stimuli can induce STAT activation such as engagement of the antigen receptor on B cells (73), or of CD2 on T cells (74). STAT activation in response to these stimuli is delayed, suggesting that the mechanism for phosphorylation is distinct from that mediated by cytokines. STAT phosphorylation is also enhanced by reactive oxygen species, which may contribute to the phosphorylation of STATs triggered by physiologic cytokines and growth factors (75,76). Thus, STATs are activated by a large variety of stimuli mediated by a number of tyrosine kinases.

The hypothesis that STATs occupy a convergence point for a variety of cellular pathways was enhanced by the finding that STATs could also be phosphorylated on distinct serine residues (7779). Although serine phosphorylation is not sufficient to lead to STAT activation, the phosphorylation of specific serine residues enhances the transcriptional activation that occurs following tyrosine phosphorylation. The kinases mediating the serine phosphorylation of STATs are still being elucidated, although there is evidence that JNK (80), p38 (80,81), MAP kinase (82), protein kinase C, and protein kinase A (83) might all be able to lead to phosphorylation of these sites.

Six STAT family members have been identified, of which one, STAT5, is encoded by two highly related genes (8488). All STATs share structural similarities including a unique tyrosine residue toward the carboxy terminus that is required for activation, a serine residue distal to the critical tyrosine that also can be phosphorylated, a phosphotyrosine-binding SH2 domain, and a DNA-binding domain. STATs are activated individually or in combination in response to a wide variety of factors. STAT1, STAT3, and STAT5 are each activated by a large number of cytokines; STAT2, STAT4, and STAT6 are activated by relatively few of them. An important area of research centers on determining how the spectrum of STATs activated in response to a given stimulus leads to a unique transcriptional response.

In addition to forming dimers and binding DNA directly, evidence is mounting that STATs can interact with other families of transcription factors. STAT1-STAT2 heterodimers formed in response to treatment with IFN-α associate with a non-STAT DNA-binding protein termed p48, a member of the IFN regulatory factor-1 family (3,89). Subsequently, various STAT family members have been found to associate with c-jun (90), C/EBPα (91,92), SP1 (93), and the glucocorticoid receptor (94). These studies suggest that STATs may have important effects acting both independently and in concert with other transcription factors.

Cellular Functions Modulated by STATs

Proliferation

Given that STATs integrate signals from a variety of pathways, it would be expected that they regulate genes important for critical cellular functions such as growth, differentiation, and survival. In fact, there is evidence that STAT binding sites are located within the promoter regions of genes that affect each of these processes. Protooncogenes such as c-fos, which may play an important role in the progression of the cell cycle, can be induced by STAT proteins through the so-called sis-inducible element (67,82,9599), although maximal induction of c-fos may require concomitant activation of the ras pathway as well (71,95,99). IL-3-induced proliferation is at least partly dependent on STAT5, as dominant inhibitory forms of this STAT decrease the mitogenic effect of IL-3 (99). Similarly, STAT3 is activated during EGF-induced mitogenesis and appears to be necessary for cellular proliferation (100). Furthermore, the increased synthesis of STAT1 induced by IFN-γ can potentiate the mitogenic actions of growth factors such as EGF and PDGF (101). By contrast, STAT1 may also mediate growth arrest via induction of the cyclindependent kinase inhibitor p21 (102). Thus, although STAT activation is necessary for proliferation in many cell types, it can be associated with growth inhibition as well.

Differentiation

In addition to modulating cellular proliferation, STATs can mediate transcriptional events associated with the differentiated functions of cells. For example, STAT5, originally described as “mammary gland factor,” mediates the transcriptional activation of milk proteins in response to prolactin (57,58,103). STAT6, activated in response to IL-4, regulates expression of cell surface proteins such as major histocompatibility (MHC) class II antigens and is involved in immunoglobulin class switching and Th2 differentiation (104106). That STAT-mediated modulation of MHC class II expression is important to host immunity is indicated by the finding that cytomegalovirus (CMV) may escape immune clearance by interfering with STAT-mediated up-regulation of MHC class II molecules (107). Complementing the actions of STAT6, STAT4 plays an important role in the Th1 differentiation pathway for T lymphocytes (108,109). In addition to being associated with cell growth, STAT3 activation has been found to be critical for differentiation of astrocytes (110), keratinocytes (111), and myeloid cells (112), and plays an important role in mediating the formation of epithelial tubules in response to hepatocyte growth factor (113). Thus, STAT activation can be an important approach to overcome blocks in differentiation in malignant cells.

Apoptosis

STATs may play a role both in facilitating apoptosis and preventing it, depending on the system. In human fibroblasts STAT1 is required for the constitutive expression of several caspases, proteases required for executing the cell death pathway (114). Up-regulation of Fas and FasL, molecules that can initiate the apoptotic cascade, is mediated by STAT1 in response to IFN-γ (115). STAT3 appears to mediate apoptosis in Jurkat T cells following ligation of MHC class I molecules (116). On the other hand, STAT1 activation is associated with the activation of the anti-apoptotic protein bcl-x in colorectal carcinoma cells (117) and in cardiac myocytes (118). Thus, STATs may mediate opposing effects on survival in different cell types, perhaps reflecting the fact that the same physiologic stimulus may lead to survival in some cells and apoptosis in others.

STATs in the Pathogenesis of Malignancy

If STATs are involved in the physiologic regulation of processes such as survival, growth, and differentiation, then it would be expected that derangements in STAT signaling could lead to the development of malignancies. Studies in Drosophila, which express a Jak homolog (119) and a STAT homolog (120,121), indicate that gain-of-function mutations affecting the Jak can lead to a form of leukemia (122,123). Over the last several years, evidence has accumulated indicating that, by a variety of mechanisms, inappropriate activation of STATs may play a role in human malignancy as well (Fig. 2). This evidence developed initially from studies on cells taken from patients with hematologic malignancies. Abnormal STAT activation was subsequently found in patients with epithelial and mesenchymal tumors, suggesting that STAT activation might be a common pathway for neoplastic cell growth. Finally, mechanistic studies have begun to shed light on the kinases that mediate STAT phosphorylation in tumors and have confirmed that this activation is necessary for malignant cell growth.

Fig. 2
figure 2

Mechanisms for STAT activation in cancer. A variety of pathways have been described leading to STAT activation in tumor cells, including (1) autocrine or (2) paracrine activation of cytokine receptors, (3) oncoprotein kinases for which STATs are not physiological substrates (such as Bcr-Abl), (4) mutated cellular tyrosine kinases (such as Src), (5) chimeric oncoproteins which activate kinases that normally phosphorylate STATs (such as Tel-Jak), (6) proteins that can activate a receptor independent of ligand (such as in HTLV-I transformation), and (7) activated serine kinases that do not activate the STATs per se, but may contribute to inappropriate responses to physiologic signals.

STAT Activation in Hematologic Malignancies

acute leukemias. The first evidence that inappropriate activation of STATs might be playing a role in human cancer came from studies on leukemic cells taken directly from patients. Using electrophoretic mobility shift assays to detect tyrosine phosphorylated STATs, constitutive activation of STAT5 and STAT1 was found in acute lymphoblastic leukemia (ALL) cells and of STAT1, STAT3, and STAT5 in acute myelogenous leukemia (AML) cells (124126). These studies provided direct evidence that in contrast to normal cells, leukemic cells from untreated patients contain activated STAT transcription factors that could be driving their abnormal growth.

chronic myelogenous leukemia and bcr-abl. Chronic myelogenous leukemia (CML) has long been known to be characterized cytogenetically by the presence of the Philadelphia chromosome, a product of a reciprocal translocation between chromosomes 9 and 22 (127,128). This translocation results in the generation of a fusion protein termed Bcr-Abl, a highly active tyrosine kinase that can transform hematopoietic cells in vitro and in vivo (129132). Whereas many hematopoietic cell lines require exogenous cytokines such as IL-3 or GM-CSF for survival and proliferation, the introduction of Bcr-Abl into these cells relieves the requirement for cytokines and leads to growth factor independence (133). Since IL-3 and GM-CSF exert at least some of their effects by activating STAT transcription factors, the possibility was considered that Bcr-Abl caused factor-independent cell growth and by extension, cellular transformation, through the activation of STATs. In fact, Bcr-Abl transformation of hematopoietic cells leads to the tyrosine phosphorylation of STAT1 and STAT5 in both model systems and in cells of patients with CML (134138). These cells do not display any evidence of Jak activation, suggesting that the tyrosine phosphorylation of the STATs is mediated directly by the Bcr-Abl kinase. Bcr-Abl has two predominant forms, one of 210 kDa associated with CML, and a 190 kDa form that is associated with ALL (139). The reason that these isoforms are associated with distinct hematologic malignancies is unclear, but there are both qualitative (136) and quantitative (135) differences in STAT phosphorylation mediated by these proteins that may underlie their different biological effects. Recent evidence has suggested that the introduction of a dominant interfering form of STAT5, which inhibits STAT5 function, can block Bcr-Abl-mediated cellular transformation (140). This has provided the first direct evidence that cellular transformation by Bcr-Abl requires the activation of specific STATs.

stat serine phosphorylation in chronic lymphocytic leukemia (cll). Given that cytokine-driven cell growth is associated with STAT tyrosine phosphorylation, it is not surprising that constitutive STAT tyrosine phosphorylation is seen in rapidly growing leukemias. However, the form of leukemia most common in the Western world, CLL, is characterized by the gradual accumulation of relatively differentiated B lymphocytes (141). As such, it might be expected that activation of signaling pathways in CLL may be more subtle than the full activation seen in these other malignancies. In fact, no constitutive tyrosine phosphorylation of STATs is found in CLL cells (83). Thus, the possibility was considered that STAT1 and STAT3, transcription factors activated by cytokines important in lymphocyte function such as IL-2 (33,142) and IL-6 (143), might have a distinct modification. In fact, CLL cells from untreated patients uniformly contain STAT1 and STAT3 phosphorylated on the serine residues known to modulate their function, whereas normal peripheral blood or CD5+ B cells lack this modification (83). Although serine phosphorylation does not lead to the nuclear translocation and DNA binding induced by tyrosine phosphorylation, serine phosphorylation does enhance the transcriptional response mediated by tyrosine-phosphorylated STATs. Thus, in CLL, the constitutive serine phosphorylation may amplify the transcriptional induction mediated by STAT1 or STAT3 once these proteins are tyrosine phosphorylated following cytokine stimulation or through other mechanisms. In this way, inappropriate STAT serine phosphorylation would be expected to have a more subtle effect on B cell function, consistent with the biology of CLL.

stat activation in lymphomas. Activation of STAT proteins is not restricted to leukemia cells. A cell line derived from a patient with mycosis fungoides, a form of cutaneous T cell lymphoma, contains STAT3, which is constitutively activated (144). In these cells, the slowly migrating form of STAT3, which is phosphorylated on ser-727 (78,83), is tyrosine phosphorylated as well. This raises the possibility that the dual phosphorylation of STAT3 on both tyrosine and serine residues is an important part of cellular transformation. In the mycosis fungoides cell lines, STAT3 associates with Jak3 and tyk2, and the Jak kinase inhibitor AG490 inhibits STAT3 tyrosine phosphorylation and the growth of these cells (144). These results suggest that inappropriate activation of endogenous Jak family kinases may contribute to STAT activation in mycosis fungoides. However, this Jak activation in the absence of growth factors does not completely recapitulate cytokine-induced signaling. In nonmalignant cell lines and primary cells, IL-2 induces the tyrosine phosphorylation of STAT1 and STAT5 in addition to STAT3. Although STAT5 is not phosphorylated in mycosis fungoides cells under basal conditions, the addition of IL-2 leads to the inducible phosphorylation of STAT5. As such, only part of the IL-2-induced signaling pathway is activated in these cells constitutively. It is possible that of the various biological effects induced by IL-2, STAT3 mediates proliferation, and STAT5 may mediate another effect such as cytokine production. In a transformed cell, only the IL-2-responsive transcription factors necessary for proliferation would need to be activated. In other T cell lymphomas, both STAT3 and STAT5 are activated, and this appears to be driven by autocrine (or perhaps paracrine) activation of the IL-2 receptor (145). Culture of these cells in vitro leads to a gradual loss of STAT phosphorylation, which can be reinduced by exposure to IL-2. Thus, constitutive STAT activation may occur in lymphomas by a variety of mechanisms, many of which center around the IL-2 receptor.

htlv-i-mediated transformation. Another model of hematopoietic cell transformation is represented by lymphocytes infected with human T cell lymphotropic virus I (HTLV-I). Although T cells generally require IL-2 for proliferation, T cells infected with HTLV-I eventually become IL-2-independent. Because IL-2 rapidly activates STAT1, STAT3, and STAT5 in resting or activated lymphocytes (33,142), the activation of STATs in HTLV-I-transformed cells was examined. In HTLV-1-transformed cells prior to the development of IL-2-independent growth, little STAT activation can be detected. By contrast, following development of the IL-2-independent phase of growth, prominent activation of STAT3 and STAT5 is seen (146). These findings are not restricted to in vitro models of HTLV-1 transformation, as primary leukemic cells from 8 of 12 HTLV-I-seropositive patients with adult T cell leukemia/lymphoma (ATLL) displayed constitutive activation of STAT1, STAT3, and/or STAT5 (147). In Bcr-Abl-transformed cells, there is no evidence for Jak activation, and it is proposed that the STATs become tyrosine phosphorylated directly by the Bcr-Abl kinase (134,136). By contrast, in patients with HTLV-I-associated ATLL, Jak kinases are chronically activated (147,148), and in IL-2-independent T cells transformed by HTLV-I in vitro, Jakl and Jak3 are activated (146). Although the mechanism for Jak activation in HTLV-I-transformed cells is unknown, it is likely that a protein encoded by the virus can interact with components of the IL-2 receptor to recapitulate the IL-2 signaling pathway. It is also interesting that IFN-β, which normally exerts an anti-proliferative effect on T cells, does not inhibit the growth of HTLV-I-infected T cells (149). IFN-β induces the phosphorylation of STAT1 in both infected and uninfected cells, but in HTLV-I-infected cells a greater proportion of the phosphorylated STAT1 is the truncated STAT1β form, which may be transcriptionally inactive. It remains to be determined whether this is the mechanism by which HTLV-I infected cells escape IFN-mediated suppression of growth.

stat activation in multiple myeloma. That inappropriate STAT activation might play a role in multiple myeloma was suggested by the finding that both IL-6 and the IL-6 receptor are expressed by myeloma cells. IL-6, which activates STAT1 and STAT3, promotes the growth and survival of myeloma and other B cell tumors (150153), and animals that lack IL-6 cannot support the development of these malignancies (154). The importance of STAT3 in these cancers was demonstrated by the finding that IL-6-independent B cell tumors contain constitutively activated STAT3 (155), and STAT3 is activated in the bone marrow of patients with multiple myeloma but not in normal individuals (156). The majority of STAT3 activation in myeloma cells is due to an IL-6 autocrine loop, as blocking the IL-6 receptor leads to a loss of most, but not all, STAT3 phosphorylation. This may reflect the activation of STAT3 by an independent pathway, activation of the IL-6 receptor intracellularly, or incomplete blockade of the IL-6 receptor. IL-6-mediated STAT3 activation can also be blocked by an inhibitor of Jak family kinases or by the introduction of an inhibitory form of STAT3. Inhibiting STAT3 by any of these approaches sensitizes myeloma cells to both spontaneous and fas-mediated apoptosis, perhaps through down-regulation of the anti-apoptotic protein bcl-xl (156). IL-6 can also act as a growth factor in hairy cell leukemia (157), and elevated serum levels of IL-6 confer a poor prognosis in patients with diffuse large-cell lymphoma (158). Thus, cytokines such as IL-6 which can act by autocrine, paracrine, and endocrine routes, may contribute to the development of a variety of human tumors.

STAT Activation in Nonhematologic Tumors

Extensive evidence suggests that STAT activation is important in the pathogenesis of hematologic malignancies such as leukemias and lymphomas. However, recent studies have shown that STAT activation may play an important role in the function of nonhematopoietic cells as well. As noted, transformation of fibroblasts by a number of mechanisms is associated with STAT3 activation. In addition, many of the cytokines that signal through STATs have receptors on mesenchymal and epithelial cells, and widely expressed polypeptide growth factor receptors, such as those for insulin, EGF, and PDGF, can activate STATs. Analogous to the findings in hematopoietic cells, it would be expected that inappropriate activation of STATs in nonhematopoietic tissue could lead to tumorigenesis.

breast carcinoma. The observation that prolactin, a hormone with major effects on mammary growth and differentiation, activates STAT5, raised the possibility that inappropriate activation of STATs might be a component of the pathogenesis of breast cancer (57,58,103,159). Using nuclear extracts derived from human breast carcinomas and normal mammary tissue, it was found that activated STATs, principally STAT1 and STAT3, were present in the malignant tissue (160). Furthermore, in five of nine breast cancer cell lines, but not in normal mammary epithelium, STAT3 was found to be constitutively activated (161). Since EGF receptor overexpression has been found in breast and other carcinomas, one hypothesis is that the constitutive activation of STAT3 is mediated by the EGF receptor pathway. In fact, in a cell line derived from a minimally invasive breast carcinoma, constitutive STAT3 activation is driven by the autocrine stimulation of an EGF-like molecule (162). However, in other mammary cell lines, an inhibitor of the EGF receptor kinase does not eliminate STAT3 activation, suggesting that a distinct tyrosine kinase is involved in this process (161). In another model of mammary carcinogenesis, overexpression of ornithine decarboxylase (ODC), the rate-limiting enzyme in polyamine biosynthesis, contributes to cell transformation. Overexpression of ODC leads to activation of STAT3, bypassing the need for EGF (163). Thus STAT3 activation may be a common event in mammary transformation induced by diverse stimuli.

other nonhematologic malignancies. Recent evidence has suggested that STAT3 is activated in primary squamous carcinomas of the head and neck (164). Such activation is also found in the normal mucosa of these patients, but not in the mucosa of controls, suggesting that it may be an early change in the transformation of these cells. STAT3 appears to be activated through the EGF receptor and may mediate the increased survival of these cells conferred by EGF family members (165). Similarly, STAT5 has been shown to be activated in squamous carcinoma cells and appears to promote the growth of these cells (166).

Activation of STAT3 has also been found in cell lines derived from human prostate cancer and ovarian cancer (167). Although Jak activation is not found in these cell lines, activation of various src family members is observed, and these kinases may be mediating the phosphorylation of STAT3. Using both microscopic and biochemical techniques, it has been shown that STAT1 and STAT3 are activated in atypical nevi, lesions that are precursors to melanoma (168). Thus abundant evidence suggests that inappropriate STAT activation is present in epithelial and mesenchymal tumors.

autocrine loops in nonhematologic malignancies. Although IL-6 is often viewed as a hematopoietic cytokine, it can affect the biology of epithelial cells as well. Increased expression of both IL-6 and the IL-6 receptor has been seen in colorectal carcinoma, suggesting that an IL-6 autocrine loop may play a role in the pathogenesis of this disease (169). IL-6 can activate STAT1 in colorectal carcinoma cells in vitro (117). While IL-6 does not affect the growth or differentiation of these cells, it does promote their survival, and in this way may contribute to the genesis of colorectal tumors. The presence of an IL-6-mediated autocrine loop may have relevance to the prevention of these neoplasms as well. It has been suggested that people who consume a diet high in fiber have a decreased risk of colorectal cancer (170173). Butyrate, a short-chain fatty acid found in the bowel lumen of people who consume a high-fiber diet, down-regulates expression of the IL-6 receptor in vitro, and in this way breaks the IL-6 autocrine loop (117). Whether this mechanism is important in the development of colon malignancies remains to be determined, but it is a potentially attractive target for developing chemoprotective agents for these cancers. EL-6 can affect the biology of other epithelial malignancies as well. For example, in renal cell carcinoma (174) and prostate cancer (175) IL-6 decreases the sensitivity to chemotherapeutic agents. Thus, cytokine-driven STAT activation may play a major role in the biology of both hematologic and nonhematologic malignancies.

Mechanisms for STAT Kinase Activation in Cancer: Formation of Kinase Fusion Proteins

An important question that arises from these findings is the identification of the kinase(s) catalyzing STAT phosphorylation in tumors. As in the case of Bcr-Abl, recent molecular and genetic evidence has indicated that fusion proteins formed as a result of chromosomal translocations can lead to activation of this pathway. Chromosomal translocations between the short arms of chromosome 9 and chromosome 12 have been described in hematologic malignancies, most commonly, childhood ALL (176178). Subsequently, it has been shown that as a result of this translocation, Jak2 (on chromosome 9) becomes fused with a member of the Ets family of transcription factors (on chromosome 12) (179,180). Ets transcription factors form complexes physiologically through a specific oligomerization domain (181). The fusion proteins contain the kinase domain of Jak2 and the oligomerization domain of the transcription factor. This results in oligomerization of the Jak2 kinase, which recapitulates the activation of Jaks by cytokine-induced dimerization, and leads to constitutive kinase activity (180,182). Since Jak-mediated STAT phosphorylation normally requires the presence of a cytokine receptor, it might not follow that independent Jak activation in the cytoplasm would induce STAT phosphorylation. Nonetheless, in hematopoietic cell lines, the introduction of Tel/Jak2 results in the activation of STAT1, STAT3, and STAT5, and cytokine-independent growth, and in animal models Tel/Jak2 fusions can induce myeloproliferative disorders (183).

Other examples of fusions between kinases and transcription factors have been described, including those of Tel/Abl (184,185), NPM/ALK (186), and ZNF198/FGFR1 (187). In several leukemias, fusions have been found between the PDGF receptor (PDGFR) and proteins that can mediate dimerization (188190). The PDGF receptor is one of the polypeptide receptor tyrosine kinases that can induce STAT activation (191,192). In Tel/PDGFR fusions, analogous to Tel/Jak2, the PDGFβR is activated by dimerization mediated by the oligomerization domain of the transcription factor Tel (193,194). Hematopoietic cells transformed with TEL/PDGFβR become growth factor independent, and display constitutive activation of STAT family members (195), providing further evidence that the forced constitutive activation of STAT family members may be critical to the pathogenesis of these leukemias.

Mechanisms for STAT Kinase Activation in Cancer: Activated Cellular Tyrosine Kinases

Kinases that phosphorylate STATs under physiological conditions, such as Jaks and growth factor receptors, can be activated by mutations to induce STAT phosphorylation continuously. Kinases that may not normally phosphorylate STATs can also become activated through mutations to phosphorylate STATs in models of neoplasia and in human cancer (196). With physiologic stimuli, STAT activation is a transient event. Two key unanswered questions related to STAT signaling in cancer are the following: Why are the mechanisms that normally turn off STAT activation not functioning in these cells? Also, what is the difference in gene induction between continuously activated and transiently activated STATs?

abl. Given that c-abl is a nuclear protein, it is unlikely that it is a physiologic STAT kinase. Nonetheless, the altered activity and subcellular localization of Bcr-Abl clearly allows abl to activate STATs. Pre-B cells transformed with the oncogenic tyrosine kinase v-abl also show constitutive STAT activation (197). In contrast to Bcr-Abl-mediated transformation, v-abl associates with Jakl and Jak3, both of which are activated in these cells. Several lines of evidence suggest that the ability of v-abl to induce the activation of Jakl is critical for its ability to transform hematopoietic cells (198). Thus, transforming tyrosine kinases may lead to STAT activation by direct phosphorylation, or indirectly by activating Jak family members.

src. In contrast to abl, which is unlikely to be a physiologic STAT kinase, src may be involved in mediating some cytokine-induced STAT phosphorylation. Evidence from a model system in which STAT3 phosphorylation is induced by IL-3 indicates that activation of c-src rather than Jak2 is critical for STAT3 activation (64). This suggests that src family kinases may be important in mediating STAT activation in response to cytokines under physiologic conditions. It is clear that mutated forms of src and related kinases can lead to STAT activation and cellular transformation. In models of fibroblast transformation, many tyrosine kinases have been found to lead to STAT3 phosphorylation, including v-src and v-fps, polyoma middle T antigen (which activates several non-receptor tyrosine kinases), and v-sis, which activates the PDGF receptor (161,199201). STAT3 activation is not common to all forms of fibroblast transformation, as SV40 large T antigen, v-raf, and v-ras, all of which operate through other mechanisms, fail to lead to STAT3 activation (161).

A key question that arises in studying neoplastic transformation, especially that induced by a potent tyrosine kinase, is whether the STAT activation seen is integral to the physiologic changes that occur or merely reflects the presence of an activated kinase with a broad substrate range. In the case of src, strong evidence suggests that STAT3 activation is critical for transformation. Wild-type STAT3 enhances the transforming potential of v-src, and dominant interfering forms of STAT3 inhibit src-mediated transformation (202). STAT3β, a naturally occurring splice variant of STAT3, lacks the transactivation domain and the site of serine phosphorylation, and it also functions in a dominant inhibitory capacity (203). Introduction of STAT3β blocks STAT3-mediated gene activation driven by src (204). Furthermore, STAT3β inhibits src-mediated cellular transformation, but not that induced by activated ras, providing additional evidence that STAT activation is central to the mechanism of cellular transformation mediated by some, but not all, dominant transforming oncogenes. These experiments suggest that STAT3 is not necessary for fibroblast transformation, but represents one pathway by which these cells can become malignant, src associates with STAT3 and can phosphorylate this protein in vitro (200,201). However, in contrast to Bcr-Abl, src may also act through Jak kinases, src-transformed cells have been reported to display activation of Jak1 and, to a lesser extent, Jak2 (205), although this finding may be cell type dependent (201). Since the Jaks themselves are activated through tyrosine phosphorylation, it is possible that src-mediated STAT activation can occur either directly or through the activation of the Jaks. Although STAT1 and STAT5 are also phosphorylated in src-transformed cells, these proteins do not coprecipitate with src (201). Thus, distinct mechanisms may govern the phosphorylation and activation of various STAT proteins in src-transformed cells.

lck. Another src family member is the Lck tyrosine kinase, which is important for T cell development and function. In a mouse T cell lymphoma characterized by Lck overexpression, STAT3 and STAT5, as well as Jak1 and Jak2, are constitutively activated (206). Lck-induced STAT3 activation also seems to be important in transformation mediated by the DNA tumor virus herpesvirus saimiri (207,208). The transforming tyrosine kinase-interacting protein Tip-484, which is encoded by the virus, directly interacts with and activates Lck, and leads to the phosphorylation of STAT1 and STAT3. Thus the Lck tyrosine kinase, which is critical for normal T cell activation, can be co-opted in the process of cellular transformation either directly, by over-expression, or indirectly, through a protein that enhances its kinase activity.

eyk. c-Eyk encodes a receptor tyrosine kinase, which can become mutated to form a dominant transforming oncogene. The activated Eyk tyrosine kinase associates with STAT1, STAT3, and Jak1, and leads to the constitutive activation of these STATs (209). Although STAT3 activation is critical for src-mediated transformation, in the case of Eyk, cellular transformation correlates more closely with the activation of STAT1 rather than STAT3 (210). Thus, inappropriate STAT tyrosine phosphorylation can occur in transformed cells as a result of a mutation involving a kinase that may not normally phosphorylate STATs (such as abl) or kinases that do phosphorylate STATs under physiologic conditions (such as Jaks and PDGFβR), or by the inappropriate activation of apparently normal kinases (as with HTLV-I transformation).

Activation of Specific STATs in Tumors

Among the STAT family members involved in experimental and primary human tumors, STAT1, STAT3, and STAT5 appear to be inappropriately phosphorylated most often. This may reflect the fact that these proteins are activated by the widest array of cytokines and may have relatively broad effects in promoting cell growth and/or survival. The three STAT family members that do not play a major role in tumors appear to have more limited and defined actions. STAT2 is activated in response to IFN-α, but generally not in response to other stimuli. Since IFN-α appears to have largely a growth-inhibitory effect, it is not surprising that activation of STAT2 has not been defined as an abnormality in tumor cells. STAT4 and STAT6 are activated by a small number of cytokines, and they appear to play a role largely in regulating the function and differentiation of T lymphocytes and natural killer (NK) cells (35,106,108,109,211,212). Thus, the absence of abnormalities of these transcription factors in malignancy is to be expected.

Anti-Cancer Therapy Targeting STAT Transcription Factors

General Considerations

Given the wide array of tumors in which inappropriate tyrosine or serine phosphorylation of STATs is found, inhibition of STAT signaling is an attractive approach for anti-cancer therapy. In selecting targets for therapeutic intervention, several requirements must be met: (1) the target must be expressed or regulated differently in malignant cells compared to normal cells; (2) it must be amenable to manipulation that can reverse the abnormality; and (3) its modulation in normal cells must not cause undue toxicity. Recent evidence suggests that STATs fulfill each of these criteria. They are clearly activated inappropriately in tumor cells, compared to normal cells, and a variety of approaches can be taken to inhibit their function. However, the most difficult aspect of cancer therapeutics is the issue of selectivity, specifically how to inhibit or kill tumor cells while leaving normal cells unscathed. Although STATs mediate important processes in a variety of cells, it does not follow that inhibition of STAT function will induce toxicity in normal tissue. Both quantitative and qualitative data provide insight into the effects of STAT inhibition. From a quantitative standpoint, the magnitude of STAT activation is considerably greater in models of neoplastic cell growth compared to normal cell growth. Using an IL-2-dependent cell line such as NKL (213), it can be shown that prominent tyrosine phosphorylation of STAT1 and STAT5 occurs after the cells are starved, then stimulated briefly with IL-2 (214). If, however, the cells are allowed to grow continuously in the presence of IL-2, the magnitude of activation of these STATs is greatly reduced (Fig. 3). This may reflect the transient nature of cytokine-induced STAT activation and additional factors such as cell cycle asynchrony. By contrast, the magnitude of STAT activation in tumor systems is usually comparable to that seen in the starvation-stimulation model of STAT activation, which is many-fold higher than that found under physiological growth conditions. It may be that in malignant cells high levels of phosphorylated STATs are required, perhaps to overcome compensatory regulatory pathways. Thus, even partial inhibition of STAT activation might be sufficient to inhibit tumor cell growth without causing toxicity in normal cells.

Fig. 3
figure 3

The magnitude of STAT activation varies with physiologic conditions. IL-2-dependent NKL cells were cultured in IL-2 containing growth medium (lanes 1 and 3) or were starved of IL-2 for 16 hr (lanes 3 and 4). Cells were left untreated (lanes 1 and 3) or were stimulated with supplemental IL-2 (lanes 2 and 4). Tyrosine-phosphorylated STAT1 and STAT5 were determined by performing a Western blot with an antibody that recognizes the activated, tyrosine-phosphorylated form of both proteins. IL-2 treatment of starved cells, which mimics the signaling response in transformed cells, leads to a much greater level of STAT1 and STAT5 activation than that seen in cells growing under physiologic conditions.

A second argument for the feasibility of targeting STATs in anti-cancer therapy makes use of findings from animals in which specific STATs have been eliminated by gene targeting. For example, STAT5 has been found to be activated inappropriately in a number of leukemias and lymphomas. Animals that do not express STAT5a and STAT5b, despite defects with growth hormone and prolactin signaling, are viable and relatively healthy (215). Thus, although STAT5 is involved in mediating the effects of a number of cytokines, there appears to be sufficient redundancy to allow normal processes to occur in its absence. Animals that lack STAT1 also develop normally. However, these animals manifest a defect in cell-mediated immunity, suggesting that targeting of STAT1 may lead to immunosuppression as a side effect. STAT1 null mice may also have an increased susceptibility to malignancies, perhaps through a defect in immune surveillance (216). However, it is unclear whether a similar mechanism is active in humans, as suppression of cell-mediated immunity is generally not associated with epithelial malignancies. The loss of STAT4 (108,109) or STAT6 (106,211,212) affects fairly defined immune functions, but does not lead to widespread abnormalities. One difficulty in interpreting experiments performed in knockout animals is that a defect may result from the absence of a protein during a developmental phase, while the protein itself may be dispensable in adult tissue. In addition to STAT1 and STAT5, the other STAT found to be activated commonly in malignancy is STAT3. Animals that lack this protein die early during embryogenesis (217), making it difficult to predict how inhibition of STAT3 might affect the physiology of an animal in which developmental processes have been completed. Nonetheless, it appears that targeting the activation of STAT proteins in malignancy need not induce intolerable toxicity to normal tissue.

Finally, it is recognized that STATs are not the only transcription factors that may be responsible for malignant cell growth, and neoplastic transformation likely requires abnormalities in more than one pathway. Nonetheless, direct inhibition of STATs may be sufficient to restrict the growth or survival of tumor cells. Furthermore, many of the cellular abnormalities that induce STAT activation, such as the formation of the Bcr-Abl oncoprotein, lead to the activation of other signaling pathways. Thus, targeting molecules upstream of STATs may have multiple beneficial effects in inhibiting tumor cell growth.

Targeting of Receptors

Given our detailed understanding of STAT activation in malignancies, a number of therapeutic strategies can be devised to target this pathway at one or more points (Fig. 4). One mechanism leading to inappropriate STAT activation is the autocrine or paracrine activation of cytokine receptors. As noted earlier, IL-6 plays a role in such pathways in both hematologic and nonhematologic malignancies. If the malignant nature of cells is being driven, even in part, by such stimulation, then the development of antagonists that can block these loops might be effective. Both natural product IL-6 receptor antagonists (218) and genetically modified IL-6 variants, socalled super-antagonists (219,220), have been developed, and evidence in multiple myeloma suggests that IL-6 antagonists inhibit cell growth and make tumor cells more susceptible to cell death (156). Complete absence of IL-6 is compatible with normal existence in mice (221), and antibodies to IL-6 have been used in patients (222), demonstrating that this approach is clinically feasible.

Fig. 4
figure 4

Strategies for STAT inhibition in tumor cells. STAT-mediated cell growth and survival can be inhibited through the use of receptor antagonists, inhibitors of endogenous or mutated tyrosine kinases, STAT SH2 inhibitors (which would block both STAT recruitment to a tyrosine kinase and STAT dimerization), STAT depletion through anti-sense or pharmacologic approaches, protein inhibitors of activated STATs which can be introduced genetically or activated through biologic therapies, decoy oligonucleotides (which can compete for STAT binding to the promoters of target genes), and serine kinase inhibitors.

Since IL-6 autocrine loops may play a role in the genesis of colorectal neoplasms, inhibitors of the IL-6 receptor may be useful in the treatment or prevention of colon tumors (117). For example, an enteric coated form of an IL-6 antagonist could block IL-6 autocrine loops within the lumen of the colon, while causing no systemic effects. Butyrate, which may mediate the protective effect of a high-fiber diet in preventing colon cancers, appears to work by down-regulating the IL-6 receptor and disrupting this autocrine loop. Similarly, aspirin and other nonsteroidal antiinflammatory agents that lower the risk of colon cancer also specifically interfere with IL-6 signaling in these cells (H. Yuan, S. Mahajan, D. Frank, unpublished observations). Thus, inhibition of IL-6-induced STAT1 activation might be an effective assay to screen compounds that could prevent colorectal cancer.

Finally, in T cell lymphomas there is evidence that STAT phosphorylation mediated through autocrine (or paracrine) activation of the IL-2 receptor may be central to the pathophysiology of these tumors (145). These studies indicate that targeting cytokine receptors may be a useful intervention for a variety of hematologic and nonhematologic malignancies.

Targeting of Kinases

Since phosphorylation is required for STAT activation, the inhibition of kinases is an attractive strategy for disrupting STAT function. This can include the tyrosine kinases that are necessary for dimerization, nuclear localization, and DNA binding of STATs and/or the serine kinases that can amplify the transcriptional response mediated by a STAT. Evidence has already accrued suggesting that such a strategy may be beneficial. In ALL, Jak2 activation and STAT activation has been found. A small-molecule inhibitor of Jak2, AG490, specifically decreases Jak2 tyrosine kinase activity and inhibits the growth of ALL cells in a mouse model (223). Normal hematopoiesis is unaffected by this drug. AG490 also inhibits both the spontaneous and IL-2-driven growth of mycosis fungoides cells, a T cell malignancy characterized by constitutively activated STAT3 (144). Similarly, small-molecule inhibitors of growth factor receptor tyrosine kinases and of the abl kinase block the growth of cells transformed by Bcr-Abl, Tel-Abl, and Tel-PDGFR (224230). Such agents may be effective in malignancies by inhibiting the activation of STATs and potentially other pathways as well. In addition, endogenous Jak inhibitors, which likely serve as part of a homeostatic mechanism to limit the signaling response triggered by cytokine stimulation, are inducibly expressed in normal cells (2530). The introduction of such proteins or the genes encoding them into tumor cells could potentially have anti-cancer activity.

Finally, serine phosphorylation of STATs may be important to the biology of CLL (83), cutaneous T cell lymphomas (144), and other malignancies. Although serine phosphorylation does not activate STATs per se, by amplifying physiologic signals received by a cell, it may alter the growth or survival characteristics of a cell sufficiently to cause a tumor. This may be particularly important in relatively low-grade cancers like CLL. Thus, identification of the kinases that mediate STAT serine phosphorylation and the subsequent development of their inhibitors may be an important therapeutic strategy.

Targeting STAT Activation

The mechanisms by which STATs become activated and transduce signals suggest several possible strategies for targeting STATs themselves. STATs become phosphorylated when they are recruited to an activated tyrosine kinase through their SH2 domain (10,11). In addition, the STAT SH2 domain is essential for STAT dimerization, which occurs through reciprocal interactions between the phosphorylated tyrosine of one STAT and the SH2 domain of its dimerization partner (13,14). Thus, small-molecule inhibitors of SH2 domains could interrupt STAT signaling at two points of the pathway: recruitment to an activated kinase and dimerization. With advances in delineating the structural requirements for SH2 interactions, the development of relatively specific inhibitors of this site is feasible.

A second method to inhibit STAT signaling directly is to reduce the concentration of a STAT within a cell. One approach is the use of anti-sense oligonucleotides to directly reduce STAT production. The half-life of a STAT protein is relatively short, less than 24 hr for STAT1 (unpublished observations), making depletion of STATs through anti-sense an appealing and feasible strategy. Such an approach has been used to decrease STAT1 levels in human cells in vitro, with a concomitant reduction in the mitogenic response to growth factors such as EGF and PDGF (101). In addition, recent evidence suggests that drugs currently used as anti-cancer agents might work at least in part through the depletion of STAT1. For example, the purine analog fludarabine is an effective agent in CLL. Although fludarabine was hypothesized to act through incorporation into DNA, relatively few CLL cells are traversing the cell cycle at any given time, making such a mechanism unlikely. Given that inappropriate STAT serine phosphorylation is a hallmark of CLL (83), it was hypothesized that fludarabine might work by interfering with STAT signaling. In fact, fludarabine leads to a pronounced and specific loss of STAT1 (231). This may underlie both the anti-neoplastic actions of fludarabine and the immunosuppression that accompanies the use of this drug, which is similar to that seen in STAT1-deficient animals. These findings strengthen the hypothesis that STAT inhibition is a potentially important target in cancer therapy.

Targeting STAT DNA Binding

Even if STATs become phosphorylated within a cell, they cannot exert their biological effect until they bind to specific DNA sequences in the promoters of target genes. Thus, inhibiting the ability of an activated STAT dimer to bind to its target DNA is an effective strategy to inhibit STAT-mediated transcriptional activation. Several approaches can be envisioned. The first involves inhibiting the translocation of activated STATs from the cytoplasm, where they are phosphorylated, to the nucleus, where they exert their effects. However, the mechanisms that regulate the process of nuclear localization are poorly understood, and thus its inhibition would be a distant goal. More practically, it is conceivable that the ability of an activated STAT to bind to a target sequence in a promoter region can be inhibited. One approach involves the development of small-molecule inhibitors that could interact with the DNA binding sites of activated STATs, thereby preventing the binding to a promoter sequence. A second approach, which could be developed more easily, involves the introduction into the cell of short stretches of double-stranded DNA which mimics the target STAT binding sequence. These “decoy oligonucleotides” would be present in great molar excess over the endogenous sequences within promoter regions. When a STAT becomes activated, it would bind to the decoy oligonucleotide, precluding its interaction with the target gene promoters. Such a strategy could diminish the ability of STATs to activate genes critical for neoplastic cell growth, and might be particularly useful in combination with inhibitors of STAT phosphorylation.

Dominant Inhibitory STATs

STATs can also be inhibited by dominant inhibitory forms. Such STATs, which lack a DNA binding domain or a transactivation domain, can still from dimers with endogenous STATs. However, the dimer is functionally inactive, and thus the mutant STAT can suppress STAT function in a cell in which it is expressed (99,203,232). Although potentially cumbersome to use clinically, this approach has been useful in demonstrating the importance for STAT signaling in a variety of systems and can inhibit malignant cell growth in in vitro models (140,202,204).

STAT Modulation by Biological and Physical Agents

Biological agents in current clinical use may act by modulating STAT function. For example, IFN-α can decrease the rate of recurrence of malignant melanoma after the primary tumor has been removed (233). Although melanoma precursor lesions demonstrate constitutive activation of STAT1 and STAT3, when patients are treated with systemic IFN-α, the DNA binding of these STATs is lost (168). Although IFN-α itself can induce STAT activation acutely in melanoma cells (234), chronic systemic administration may decrease constitutive STAT activation through an independent mechanism. One attractive hypothesis is that the chronic presence of a stimulus for STAT activation induces inhibitors of activated STATs as part of a homeostatic feedback process to limit STAT-mediated transcription (2530). Thus, chronic systemic IFN-α may be able to suppress the function of STATs that have been activated by another pathway. However, an added complexity to the role of STAT signaling in melanoma is the finding that melanoma cell lines that become resistant to the cytostatic effects of IFNs show a loss of IFN-activated proteins, most commonly STAT1 (235). Similarly, in cutaneous T cell lymphoma, resistance to the growth-inhibitory effects of IFN-α is associated with the loss of STAT1 (236). These results suggest that on the one hand, STAT1 activation is present in melanoma precursors and decreases with IFN-α treatment, but that on the other hand, the loss of sensitivity to IFN correlates with a loss of STAT1. The resolution of this paradox may depend on one or more differences between the systems, such as the use of atypical nevi versus melanomas, or the use of primary human tissue versus cell lines. In addition, other modifications that affect STAT activity, such as the concomitant presence of activated STAT3 or the state of phosphorylation of ser-727 of STAT1, may also explain these differences. This is not an inconsequential issue, as activation of STAT1 can lead to growth arrest in response to cytokines other than the IFNs (237). For example, in A431 cells, EGF leads to STAT1 activation and growth inhibition (232). The introduction of a dominant interfering form of STAT1 abrogates the growth-suppressive effects of EGF, indicating that STAT1 mediates this inhibition. Thus, a key unanswered question in the role of STATs in malignant cell physiology concerns this discrepancy between growth-stimulatory and growth-suppressive effects.

STAT activation can be used therapeutically for the induction of cellular differentiation, which may underlie the actions of another class of biological agents, the retinoids. Among this class of compounds, all-trans retinoic acid (ATRA) has potent abilities to induce the differentiation of acute promyelocytic leukemia cells (238). Among the effects of ATRA is the up-regulation of STAT1 and STAT2, the two STATs activated in response to IFN-α. In addition, STAT1 becomes tyrosine phosphorylated for prolonged periods following ATRA treatment, and ATRA potentiates the growth-inhibitory effect of IFN-α (239243). These studies lend further support to the importance of STAT modulation in the mechanism of anti-tumor activity mediated by biological agents.

The inhibition of STAT activation can also occur through nonpharmacologic means. Physical and biological agents that induce immune suppression may act by blocking cytokine-mediated STAT activation. For example, UV light (244) and the adenoviral protein E1A (245) can each prevent the activation of STAT1 induced by IFN-γ. These findings suggest that modulation of STAT activity is an important target for altering cellular behavior and can be achieved through a variety of modalities.

Summary

Extensive work in the last several years has highlighted the importance of the STAT family of transcription factors in mediating the actions of cytokines and growth factors. Because these molecules transduce physiologic signals regulating growth, differentiation, and apoptosis, it is not surprising that inappropriate STAT activation has been found in a wide range of human tumors. The goal now is to translate this enhanced understanding of the molecular pathogenesis of malignancies into effective strategies to treat cancer with more specificity and less toxicity. Reagents developed for this purpose may be useful not only in the treatment of malignancies but also in further dissecting the role that STAT signaling plays in normal and neoplastic cellular physiology.

References

  1. Levy DE, Kessler DS, Pine R, Reich N, Darnell JE Jr. (1988) Interferon-induced nuclear factors that bind a shared promoter correlate with positive and negative transcriptional control. Genes Dev. 2: 383–393.

    Article  CAS  PubMed  Google Scholar 

  2. Khan KD, Shuai K, Lindwall G, Maher SE, Darnell JE Jr, Bothwell AL. (1993) Induction of the Ly-6A/E gene by interferon alpha/beta and gamma requires a DNA element to which a tyrosine-phosphorylated 91-kDa protein binds. Proc. Natl. Acad. Sci. U.S.A. 90: 6806–6810.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Fu X-Y, Schindler C, Improta T, Aebersold R, Darnell JE Jr. (1992) The proteins of ISGF-3, the interferon α-induced transcriptional activator, define a gene family involved in signal transduction. Proc. Natl. Acad. Sci. U.S.A. 89: 7840–7843.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Shuai K, Schindler C, Prezioso VR, Darnell JE Jr. (1992) Activation of transcription by IFN-γ: tyrosine phosphorylation of a 91-kD DNA binding protein. Science 258: 1808–1812.

    Article  CAS  PubMed  Google Scholar 

  5. Velazquez L, Fellous M, Stark G, Pellegrini S. (1992) A protein tyrosine kinase in the interferon α/β signaling pathway. Cell 70: 313–322.

    Article  CAS  PubMed  Google Scholar 

  6. Muller M, Briscoe J, Laxton C, et al. (1993) The protein tyrosine kinase JAK1 complements defects in interferon-alpha/beta and gamma signal transduction. Nature 366: 129–135.

    Article  CAS  PubMed  Google Scholar 

  7. Waiting D, Guschin D, Muller M. (1993) Complementation by the protein tyrosine kinase JAK-2 of a mutant cell line defective in the interferon-gamma signal transduction pathway. Nature 366: 166–170.

    Article  Google Scholar 

  8. Silvennoinen O, Ihle JN, Schlessinger J, Levy DE. (1993) Interferon-induced nuclear signaling by Jak protein tyrosine kinases. Nature 366: 583–585.

    Article  CAS  PubMed  Google Scholar 

  9. Colamonici OR, Uyttendaele H, Domanski P, Yan H, Krolewski JJ. (1994) p135tyk2, an interferon-α-activated tyrosine kinase, is physically associated with an interferon-α receptor. J. Biol. Chem. 269: 3518–3522.

    CAS  PubMed  Google Scholar 

  10. Heim MH, Kerr IM, Stark GR, Darnell JE Jr. (1995) Contribution of STAT SH2 groups to specific interferon signaling by the Jak-STAT pathway. Science 267: 1347–1349.

    Article  CAS  PubMed  Google Scholar 

  11. Stahl N, Farruggella TJ, Boulton TG, Zhong Z, Darnell JE Jr, Yancopoulos GD. (1995) Choice of STATs and other substrates specified by modular tyrosine-based motifs in cytokine receptors. Science 267: 1349–1353.

    Article  CAS  PubMed  Google Scholar 

  12. Shuai K, Stark GR, Kerr IM, Darnell JE Jr. (1993) A single phosphotyrosine residue of Stat91 required for gene activation by interferon-γ. Science 261: 1744–1746.

    Article  CAS  PubMed  Google Scholar 

  13. Shuai K, Horvath CM, Huang LHT, Qureshi SA, Cowburn D, Darnell JE Jr. (1994) Interferon activation of the transcription factor Stat91 involves dimerization through SH2-phosphotyrosyl peptide interactions. Cell 76: 821–828.

    Article  CAS  PubMed  Google Scholar 

  14. Chen X, Vinkemeier U, Zhao Y, Jeruzalmi D, Darnell JE Jr, Kuriyan J. (1998) Crystal structure of a tyrosine phosphorylated STAT-1 dimer bound to DNA. Cell 93: 827–839.

    Article  CAS  PubMed  Google Scholar 

  15. Bhattacharya S, Eckner R, Grossman S, et al. (1996) Cooperation of Stat2 and p300/CBP in signalling induced by interferon-α. Nature 383: 344–347.

    Article  CAS  PubMed  Google Scholar 

  16. Zhang JJ, Vinkemeier U, Gu W, Chakravarti D, Horvath CM, Darnell JE Jr. (1996) Two contact regions between Stat1 and CBP/p300 in interferon γ signaling. Proc. Natl. Acad. Sci. U.S.A. 93: 15092–15096.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Horvai AE, Xu L, Korzus E, et al. (1997) Nuclear integration of JAK/STAT and Ras/AP-1 signaling by CBP and p300. Proc. Natl. Acad. Sci. U.S.A. 94: 1074–1079.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Kurokawa R, Kalafus D, Ogliastro M-H, et al. (1998) Differential use of CREB binding protein-coactivator complexes. Science 279: 700–703.

    Article  CAS  PubMed  Google Scholar 

  19. Korzus E, Torchia J, Rose DW, et al. (1998) Transcription factor-specific requirements for coactivators and their acetyltransferase functions. Science 279: 703–707.

    Article  CAS  PubMed  Google Scholar 

  20. David M, Grimely PM, Finbloom DS. (1993) A nuclear tyrosine phosphatase downregulates interferon-induced gene expression. Mol. Cell. Biol. 13: 5715–5721.

    Article  Google Scholar 

  21. Haque SJ, Flati V, Deb A, Williams BR. (1995) Roles of protein-tyrosine phosphatases in Stat1 alpha-mediated cell signaling. J. Biol. Chem. 270: 25709–25714.

    Article  CAS  PubMed  Google Scholar 

  22. Kim TK, Maniatis T. (1996) Regulation of interferon-gamma-activated STAT1 by the ubiquitin-proteasome pathway. Science 273: 1717–1719.

    Article  CAS  PubMed  Google Scholar 

  23. Haspel RL, Salditt-Georgieff M, Darnell JE Jr. (1996) The rapid inactivation of nuclear tyrosine phosphorylated Stat1 depends upon a protein tyrosine phosphatase. EMBO J. 15: 6262–6268.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Yu CL, Burakoff SJ. (1997) Involvement of proteasomes in regulating Jak-STAT pathways upon interleukin-2 stimulation. J. Biol. Chem. 272: 14017–14020.

    Article  CAS  PubMed  Google Scholar 

  25. Matsumoto A, Masuhara M, Mitsui K, et al. (1997) CIS, a cytokine inducible SH2 protein, is a target of the JAK-STAT5 pathway and modulates STAT5 activation. Blood 89: 3148–3154.

    CAS  PubMed  Google Scholar 

  26. Starr R, Willson TA, Viney EM, et al. (1997) A family of cytokine-inducible inhibitors of signaling. Nature 387: 917–921.

    Article  CAS  PubMed  Google Scholar 

  27. Endo TA, Masuhara M, Yokouchi M, et al. (1997) A new protein containing an SH2 domain that inhibits JAK kinases. Nature 387: 921–924.

    Article  CAS  PubMed  Google Scholar 

  28. Naka T, Narazaki M, Hirata M, et al. (1997) Structure and function of a new STAT-induced STAT inhibitor. Nature 387: 924–928.

    Article  CAS  PubMed  Google Scholar 

  29. Chung CD, Liao J, Liu B, et al. (1997) Specific inhibition of Stat3 signal transduction by PIAS3. Science 278: 1803–1805.

    Article  CAS  PubMed  Google Scholar 

  30. Liu B, Liao J, Rao X, et al. (1998) Inhibition of Statl-mediated gene activation by PIAS1. Proc. Natl. Acad. Sci. U.S.A. 95: 10626–10631.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Beadling C, Guschin D, Witthuhn BA, et al. (1994) Activation of JAK kinases and STAT proteins by interleukin-2 and interferon α, but not the T cell antigen receptor, in human T lymphocytes. EMBO J. 13: 5605–5615.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Nielsen M, Svejgaard A, Skov S, Odum N. (1994) Interleukin-2 induces tyrosine phosphorylation and nuclear translocation of stat3 in human T lymphocytes. Eur. J. Immunol 24: 3082–3086.

    Article  CAS  PubMed  Google Scholar 

  33. Frank DA, Robertson M, Bonni A, Ritz J, Greenberg ME. (1995) IL-2 signaling involves the phosphorylation of novel Stat proteins. Proc. Natl. Acad. Sci. U.S.A. 92: 7779–7783.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Hou J, Schindler U, Henzel WJ, Wong SC, McKnight SL. (1995) Identification and purification of human Stat proteins activated in response to interleukin-2. Immunity 2: 321–329.

    Article  CAS  PubMed  Google Scholar 

  35. Wang KS, Ritz J, Frank DA. (1999) IL-2 induces STAT4 activation in primary NK cells and NK cell lines but not in T cells. J. Immunol. 162: 299–304.

    CAS  PubMed  Google Scholar 

  36. Larner AC, David M, Feldman GM, et al. (1993) Tyrosine phosphorylation of DNA binding proteins by multiple cytokines. Science 261: 1730–1733.

    Article  CAS  PubMed  Google Scholar 

  37. Kotanides H, Reich NC. (1993) Requirement of tyrosine phosphorylation for rapid activation of a DNA binding factor by IL-4. Science 262: 1265–1267.

    Article  CAS  PubMed  Google Scholar 

  38. Schindler C, Kashleva H, Pernis A, Pine R, Rothman P. (1994) STF-IL-4: a novel IL-4-induced signal transducing factor. EMBO J. 13: 1350–1356.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Wegenka UM, Buschmann J, Lutticken C, Heinrich PC, Horn F. (1993) Acute-phase response factor, a nuclear factor binding to acute-phase response elements, is rapidly activated by interleukin-6 at the posttranslational level. Mol. Cell. Biol. 13: 276–288.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Yuan J, Wegenka UM, Lutticken C, et al. (1994) The signalling pathways of interleukin-6 and gamma interferon converge by the activation of different transcription factors which bind to common responsive DNA elements. Mol. Cell. Biol. 14: 1657–1668.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Bacon CM, Petricoin EFI, Ortaldo JR, et al. (1995) Interleukin 12 induces tyrosine phosphorylation and activation of STAT4 in human lymphocytes. Proc. Natl. Acad. Sci. U.S.A. 92: 7307–7311.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Wegenka UM, Lutticken C, Buschmann J, et al. (1994) The interleukin-6-activated acute-phase response factor is antigenically and functionally related to members of the signal transducer and activator of transcription (STAT) family. Mol. Cell. Biol. 14: 186–196.

    Article  Google Scholar 

  43. Lowe C, Gillespie GA, Pike JW. (1995) Leukemia inhibitory factor as a mediator of JAK/STAT activation in murine osteoblasts. J. Bone Min. Res. 10: 1644–1650.

    Article  CAS  Google Scholar 

  44. Symes AJ, Corpus L, Fink JS. (1995) Differences in nuclear signaling by leukemia inhibitory factor and interferon-gamma: the role of STAT proteins in regulating vasoactive intestinal peptide gene expression. J. Neurochem. 65: 1926–1933.

    Article  CAS  PubMed  Google Scholar 

  45. Boulton TG, Zhong Z, Wen Z, Darnell JE Jr, Stahl N, Yancopoulos GD. (1995) STAT3 activation by cytokines utilizing gp130 and related transducers involves a secondary modification requiring an H7-sensitive kinase. Proc. Natl. Acad. Sci. U.S.A. 92: 6915–6919.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Zhao Y, Nichols JE, Bulun SE, Mendelson CR, Simpson ER. (1995) Aromatase P450 gene expression in human adipose tissue. Role of a Jak/STAT pathway in regulation of the adipose-specific promoter. J. Biol. Chem. 270: 16449–16457.

    Article  CAS  PubMed  Google Scholar 

  47. Morella KK, Bruno E, Kumaki S, et al. (1995) Signal transduction by the receptors for thrombopoietin (c-mpl) and interleukin-3 in hematopoietic and nonhematopoietic cells. Blood 86: 557–571.

    CAS  PubMed  Google Scholar 

  48. Pallard C, Gouilleux F, Benit L, et al. (1995) Thrombopoietin activates a STAT5-like factor in hematopoietic cells. EMBO J. 14: 2847–2856.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Sattler M, Durstin MA, Frank DA, et al. (1995) The thrombopoietin receptor c-MPL activates JAK2 and TYK2 tyrosine kinases. Exp. Hematol. 23: 1040–1048.

    CAS  PubMed  Google Scholar 

  50. Silvennoinen O, Schindler C, Schlessinger J, Levy DE. (1993) Ras-independent growth factor signaling by transcription factor tyrosine phosphorylation. Science 261: 1736–1739.

    Article  CAS  PubMed  Google Scholar 

  51. Novak U, Mui A, Miyajima A, Paradiso L. (1996) Formation of STAT5-containing DNA binding complexes in response to colony-stimulating factor-1 and platelet-derived growth factor. J. Biol. Chem. 271: 18350–18354.

    Article  CAS  PubMed  Google Scholar 

  52. Tweardy DJ, Wright TM, Ziegler SF, et al. (1995) Granulocyte colony-stimulating factor rapidly activates a distinct STAT-like protein in normal myeloid cells. Blood 86: 4408–4416.

    Google Scholar 

  53. Chakraborty A, White SM, Schaefer TS, Ball ED, Dyer KF, Tweardy DJ. (1996) Granulocyte colony-stimulating factor activation of Stat3α and Stat3β in immature normal and leukemic human myeloid cells. Blood 88: 2442–2448.

    CAS  PubMed  Google Scholar 

  54. Argetsinger LS, Campbell GS, Yang X, et al. (1993) Identification of JAK2 as a growth hormone receptor-associated tyrosine kinase. Cell 74: 237–244.

    Article  CAS  PubMed  Google Scholar 

  55. Meyer DJ, Campbell GS, Cochran BH, et al. (1994) Growth hormone induces a DNA binding factor related to the interferon-stimulated 91-kDa transcription factor. J. Biol. Chem. 269: 4701–4704.

    CAS  PubMed  Google Scholar 

  56. Symes A, Lewis S, Corpus L, Rajan P, Hyman SE, Fink JS. (1994) STAT proteins participate in the regulation of the vasoactive intestinal peptide gene by the ciliary neurotrophic factor family of cytokines. Mol. Endocrinol 8: 1750–1763.

    CAS  PubMed  Google Scholar 

  57. Gouilleux F, Pallard C, Dusanter-Fourt I, et al. (1995) Prolactin, growth hormone, erythropoietin and granulocyte-macrophage colony stimulating factor induce MGF-Stat5 DNA binding activity. EMBO J. 14: 2005–2013.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Gouilleux F, Wakao H, Mundt M, Groner B. (1994) Prolactin induces phosphorylation of Tyr694 of Stat5 (MGF), a prerequisite for DNA binding and induction of transcription. EMBO J. 13: 4361–4369.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Rui H, Kirken RA, Farrar WL. (1994) Activation of receptor-associated tyrosine kinase JAK2 by prolactin. J. Biol. Chem. 269: 5364–5368.

    CAS  PubMed  Google Scholar 

  60. Bonni A, Frank DA, Schindler C, Greenberg ME. (1993) Characterization of a pathway for ciliary neurotrophic factor signaling to the nucleus. Science 262: 1575–1579.

    Article  CAS  PubMed  Google Scholar 

  61. Guo D, Dunbar JD, Yang CH, Pfeffer LM, Donner DB. (1998) Induction of Jak/STAT signaling by activation of the type 1 TNF receptor. J. Immunol. 160: 2742–2750.

    CAS  PubMed  Google Scholar 

  62. Bhat GJ, Thekkumkara TJ, Thomas WG, Conrad KM, Baker KM. (1994) Angiotensin II stimulates sis-inducing factor-like DNA binding activity. Evidence that the AT1A receptor activates transcription factor-Stat91 and/or a related protein. J. Biol. Chem. 269: 31443–31449.

    CAS  PubMed  Google Scholar 

  63. Marrero MB, Schieffer B, Paxton WG, et al. (1995) Direct stimulation of Jak/STAT pathway by the angiotensin II AT1 receptor. Nature 375: 247–250.

    Article  CAS  PubMed  Google Scholar 

  64. Chaturvedi P, Reddy MV, Reddy EP. (1998) Src kinases and not JAKs activate STATs during IL-3 induced myeloid cell proliferation. Oncogene 16: 1749–1758.

    Article  CAS  PubMed  Google Scholar 

  65. Saharinen P, Ekman N, Sarvas K, Parker P, Alitalo K, Silvennoinen O. (1997) The Bmx tyrosine kinase induces activation of the Stat signaling pathway, which is specifically inhibited by protein kinase Cδ. Blood 90: 4341–4353.

    CAS  PubMed  Google Scholar 

  66. Yamashita Y, Watanabe S, Miyazato A, et al. (1998) Tec and Jak2 cooperate to mediate cytokine-driven activation of c-fos transcription. Blood 91: 1496–1507.

    CAS  PubMed  Google Scholar 

  67. Fu X, Zhang J. (1993) Transcription factor p91 interacts with the epidermal growth factor receptor and mediates activation of the c-fos gene promoter. Cell 74: 1135–1145.

    Article  CAS  PubMed  Google Scholar 

  68. Ruff-Jamison S, Chen K, Cohen S. (1993) Induction by EGF and interferon-γ of tyrosine phosphorylated DNA binding proteins in mouse liver nuclei. Science 261: 1733–1736.

    Article  CAS  PubMed  Google Scholar 

  69. Sadowski HB, Shuai K, Darnell JE Jr, Gilman MZ. (1993) A common nuclear signal transduction pathway activated by growth factor and cytokine receptors. Science 261: 1739–1744.

    Article  CAS  PubMed  Google Scholar 

  70. Park OK, Schaeffer TS, Nathans D. (1996) In vitro activation of Stat3 by epidermal growth factor receptor kinase. Proc. Natl. Acad. Sci. U.S.A. 93: 13704–13708.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Leaman DW, Pisharody S, Flickinger TW, et al. (1996) Roles of JAKs in activation of STATs and stimulation of c-fos gene expression by epidermal growth factor. Mol. Cel. Biol. 16: 369–375.

    Article  CAS  Google Scholar 

  72. Chen J, Sadowski HB, Kohanski RA, Wang L-H. (1997) Stat5 is a physiologic substrate of the insulin receptor. Proc. Natl. Acad. Sci. U.S.A. 94: 2295–2300.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Karras JG, Wang Z, Coniglio SJ, Frank DA, Rothstein TL. (1996) Antigen-receptor engagement in B cells induces nuclear expression of STAT5 and STAT6 proteins that bind and transactivate an IFN-gamma activation site. J. Immunol. 157: 39–47.

    CAS  PubMed  Google Scholar 

  74. Frank DA, Mahajan S, Ritz J. (1998) Activation of T cells through CD2 leads to the delayed and prolonged activation of STAT1 Blood 92: 701a.

    Google Scholar 

  75. Simon AR, Rai U, Fanburg BL, Cochran BH. (1998) Activation of the JAK-STAT pathway by reactive oxygen species. Am. J. Physiol. 275: C1640–1652.

    Article  CAS  PubMed  Google Scholar 

  76. Sattler M, Winkler T, Verma S, et al. (1999) Hematopoietic growth factors signal through the formation of reactive oxygen species. Blood 93: 2928–2935.

    CAS  PubMed  Google Scholar 

  77. Wen Z, Zhong Z, Darnell JE Jr. (1995) Maximal activation of transcription by Statl and Stat3 requires both tyrosine and serine phosphorylation. Cell 82: 241–250.

    Article  CAS  PubMed  Google Scholar 

  78. Zhang X, Blenis J, Li H-C, Schindler C, Chen-Kiang S. (1995) Requirement of serine phosphorylation for formation of Stat-promoter complexes. Science 267: 1990–1994.

    Article  CAS  PubMed  Google Scholar 

  79. Wen Z, Darnell JE Jr. (1997) Mapping of Stat3 serine phosphorylation to a single residue (727) and evidence that serine phosphorylation has no influence on DNA binding of Statl and Stat3. Nucl. Acids Res. 25: 2062–2067.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Turkson J, Bowman T, Adnane J, et al. (1999) Requirement of Rac-1 mediated p38 and JNK signaling for Stat3 transcriptional activity induced by the Src oncoprotein. Mol. Cell. Biol. (in press).

  81. Gollob JA, Schnipper CP, Murphy EA, Ritz J, Frank DA. (1999) The functional synergy between IL-12 and IL-2 involves p38 MAP kinase and is associated with the augmentation of STAT serine phosphorylation. J. Immunol. 162: 4472–4481.

    CAS  PubMed  Google Scholar 

  82. Rajotte D, Sadowski HB, Haman A, et al. (1996) Contribution of both STAT and SRF/TCF to c-fos promoter activation by granulocyte-macrophage colony stimulating factor. Blood 88: 2906–2916.

    CAS  PubMed  Google Scholar 

  83. Frank DA, Mahajan S, Ritz J. (1997) B lymphocytes from patients with chronic lymphocytic leukemia contain STAT1 and STAT3 constitutively phosphorylated on serine residues. J. Clin. Invest. 100: 3140–3148.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Ihle JN. (1996) STATs: signal transducers and activators of transcription. Cell 84: 331–334.

    Article  CAS  PubMed  Google Scholar 

  85. Ihle JN, Kerr IM. (1995) Jaks and stats in signaling by the cytokine receptor superfamily. Trends Genet. 11: 69–74.

    Article  CAS  PubMed  Google Scholar 

  86. Ihle JN, Witthuhn BA, Quelle FW, Yamamoto K, Silvennoinen O. (1995) Signaling through hematopoietic cytokine receptors. Annu. Rev. Immunol 13: 369–398.

    Article  CAS  PubMed  Google Scholar 

  87. Darnell JE Jr. (1997) STATs and gene regulation. Science 277: 1630–1635.

    Article  CAS  PubMed  Google Scholar 

  88. Darnell JE Jr, Kerr IM, Stark GR. (1994) JakSTAT pathway and transcription activation in response to IFNs and other extracellular signaling proteins. Science 264: 1415–1420.

    Article  CAS  PubMed  Google Scholar 

  89. Schindler C, Fu X-Y, Improta T, Aebersold R, Darnell JE Jr. (1992) Proteins of transcription factor ISGF-3: one gene encodes the 91 and 84 kDa ISGF-3 proteins that are activated by interferon-α. Proc. Natl. Acad. Sci. U.S.A. 89: 7836–7839.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Schaefer TS, Sanders LK, Nathans D. (1995) Cooperative transcriptional activity of Jun and Stat3 beta, a short form of Stat3. Proc. Natl Acad. Sci. USA. 92: 9097–9101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Mikita T, Campbell D, Wu P, Williamson K, Schindler U. (1996) Requirements for interleukin-4-induced gene expression and functional characterization of Stat6. Mol. Cell. Biol. 16: 5811–5820.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Delphin S, Stavenezer J. (1995) Characterization of an interleukin 4 (IL-4) responsive region in the immunoglobulin heavy chain germline epsilon promoter: regulation by NF-IL-4, a C/EBP family member and NF-kappa B/p50. J. Exp. Med. 181: 181–192.

    Article  CAS  PubMed  Google Scholar 

  93. Look DC, Pelletier MR, Tidwell RM, Roswit WT, Holtzman MJ. (1995) Statl depends on transcriptional synergy with Spl. J. Biol Chem. 270: 30264–30267.

    Article  CAS  PubMed  Google Scholar 

  94. Stocklin E, Wissler M, Gouilleux F, Groner B. (1996) Functional interactions between Stat5 and the glucocorticoid receptor. Nature 383: 726–728.

    Article  CAS  PubMed  Google Scholar 

  95. Wagner BJ, Hayes TE, Hoban CJ, Cochran BH. (1990) The SIF binding element confers sis/PDGF inducibility onto the c-fos promoter. EMBO J. 9: 4477–4484.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Zhong Z, Wen Z, Darnell Jr JE. (1994) Stat3: a STAT family member activated by tyrosine phosphorylation in response to epidermal growth factor and interleukin-6. Science 264: 95–98.

    Article  CAS  PubMed  Google Scholar 

  97. Ruff-Jamison S, Chen K, Cohen S. (1995) Epidermal growth factor induces the tyrosine phosphorylation and nuclear translocation of Stat 5 in mouse liver. Proc. Natl Acad. Sci. U.S.A. 92: 4215–4218.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Ruff-Jamison S, Zhong Z, Wen Z, Chen K, Darnell JE Jr, Cohen S. (1994) Epidermal growth factor and lipopolysaccharide activate Stat3 transcription factor in mouse liver. J. Biol. Chem. 269: 21933–21935.

    CAS  PubMed  Google Scholar 

  99. Mui AL, Wakao H, Kinoshita T, Kitamura T, Miyajima A. (1996) Suppression of interleukin-3-induced gene expression by a C-terminal truncated Stat5: role of Stat5 in proliferation. EMBO J. 15: 2425–2433.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Grandis JR, Drenning SD, Chakraborty A, et al. (1998) Requirement of Stat3 but not Stat1 activation for epidermal growth factor-mediated cell growth in vitro. J. Clin. Invest. 102: 1385–1392.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Marra F, Choudhury GG, Abboud HE. (1996) Interferon-γ-mediated activation of STAT1α regulates growth factor-induced mitogenesis. J. Clin. Invest. 98: 1218–1230.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Chin YE, Kitagawa M, Su W-CS, You Z-H, Iwamoto Y, Fu X–Y. (1996) Cell growth arrest and induction of cyclin-dependent kinase inhibitor p21WAF1/CIP1 mediated by STAT1. Science 272: 719–722.

    Article  CAS  PubMed  Google Scholar 

  103. Wakao H, Gouilleux F, Groner B. (1994) Mammary gland factor (MGF) is a novel member of the cytokine regulated transcription factor gene family and confers the prolactin response. EMBO J. 13: 2182–2191.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Coffman RL, Lebman DA, Rothman P. (1993) Mechanism and regulation of immunoglobulin isotype switching. Adv. Immunol. 54: 229.

    Article  CAS  PubMed  Google Scholar 

  105. Rothman P, Li SC, Gorham B, Glimcher L, Alt F, Boothby M. (1991) Identification of a conserved lipopolysaccharide-plus-interleukin-4-respon-siveelement located at the promoter of germ line epsilon transcripts. Mol. Cell. Biol. 11: 5551–5561.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Kaplan MH, Schindler U, Smiley ST, Grusby MJ. (1996) Stat6 is required for mediating responses to IL-4 and for the development of Th2 cells. Immunity 4: 313–319.

    Article  CAS  PubMed  Google Scholar 

  107. Miller DM, Rahill BM, Boss JM, et al. (1998) Human cytomegalovirus inhibits major histocompatibility complex class II expression by disruption of the Jak/Stat pathway. J. Exp. Med. 187: 675–683.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Kaplan MH, Sun Y-L, Hoey T, Grusby MJ. (1996) Impaired IL-12 responses and enhanced development of Th2 cells in Stat4-deficient mice. Nature 382: 174–177.

    Article  CAS  PubMed  Google Scholar 

  109. Thierfelder WE, van Deursen JM, Yamamoto K, et al. (1996) Requirement for Stat4 in interleukin-12-mediated response of natural killer and T cells. Nature 382: 171–174.

    Article  CAS  PubMed  Google Scholar 

  110. Bonni A, Sun Y, Nadal-Vicens M, et al. (1997) Regulation of gliogenesis in the central nervous system by the JAK-STAT signaling pathway. Science 278: 477–483.

    Article  CAS  PubMed  Google Scholar 

  111. Hauser PJ, Agrawal D, Hackney J, Pledger WJ. (1998) STAT3 activation accompanies keratinocyte differentiation. Cell Growth Differ. 9: 847–855.

    CAS  PubMed  Google Scholar 

  112. Shimozaki K, Nakajima K, Hirano T, Nagata S. (1997) Involvement of STAT3 in the granulocyte colony-stimulating factor-induced differentiation of myeloid cells. J. Biol. Chem. 272: 25184–25189.

    Article  CAS  PubMed  Google Scholar 

  113. Boccaccio C, Ando M, Tamagnone L, et al. (1998) Induction of epithelial tubules by growth factor HGF depends on the STAT pathway. Nature 391: 285–288.

    Article  CAS  PubMed  Google Scholar 

  114. Kumar A, Commane M, Flickinger TW, Horvath CM, Stark GR. (1997) Defective TNF-α-induced apoptosis in STAT1-null cells due to low constitutive levels of caspases. Science 278: 1630–1632.

    Article  CAS  PubMed  Google Scholar 

  115. Xu X, Fu X–Y, Plate J, Chong AS-F. (1998) IFN-γ induces cell growth inhibition by fas-mediated apoptosis: requirement of STATl protein for upregulation of Fas and FasL expression. Cancer Res. 58: 2832–2837.

    CAS  PubMed  Google Scholar 

  116. Skov S, Nielsen M, Bregenholt S, Odum N, Claesson MH. (1998) Activation of Stat-3 is involved in the induction of apoptosis after ligation of major histocompatibility complex class I molecules on human Jurkat T cells. Blood 91: 3566–3573.

    CAS  PubMed  Google Scholar 

  117. Frank DA, Mahajan S, Yuan H. (1999) The chemoprotectant butyrate downregulates IL-6 induced signaling events in colorectal carcinoma cells. Proc. Amer. Assoc. Cancer Res. 40: 318.

    Google Scholar 

  118. Fujio Y, Kunisada K, Hirota H, Yamauchi-Takihara K, Kishimoto T. (1997) Signals through gp130 upregulate bcl-x gene expression via STAT1-binding cis-element in cardiac myocytes. J. Clin. Invest. 99: 2898–2905.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Binari R, Perrimon N. (1994) Stripe-specific regulation of pair-rule genes by hopscotch, a putative Jak family tyrosine kinase in Drosophila. Genes Dev. 8: 300–312.

    Article  CAS  PubMed  Google Scholar 

  120. Hou XS, Melnick MB, Perrimon N. (1996) marelle acts downstream of the Drosophila HOP/JAK kinase and encodes a protein similar to mammalian STATs. Cell 84: 411–419.

    Article  CAS  PubMed  Google Scholar 

  121. Yan R, Small S, Desplan C, Dearolf CR, Darnell JE Jr. (1996) Identification of a Stat gene that functions in Drosophila development. Cell 84: 421–430.

    Article  CAS  PubMed  Google Scholar 

  122. Luo H, Hanratty WP, Dearolf CR. (1995) An amino acid substitution in the Drosophila hopTum-1 Jak kinase causes leukemia-like hematopoietic defects. EMBO J. 14: 1412–1420.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Harrison DA, Binari R, Nahrenini TS, Gilman M, Perrimon N. (1995) Activation of a Drosophila Janus kinase (JAK) causes hematopoietic neoplasia and developmental defects. EMBO J. 14: 2857–2865.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Gouilleux-Gruart V, Gouilleux F, Desaint C, et al. (1996) STAT-related transcription factors are constitutively activated in peripheral blood cells from acute leukemia patients. Blood 87: 1692–1697.

    CAS  PubMed  Google Scholar 

  125. Weber-Nordt RM, Egen C, Wehinger J, et al. (1996) Constitutive activation of STAT proteins in primary lymphoid and myeloid leukemia cells and in Epstein-Barr virus (EBV)-related lymphoma cell lines. Blood 88: 809–816.

    CAS  PubMed  Google Scholar 

  126. Xia Z, Baer MR, Block AW, Baumann H, Wetzler M. (1998) Expression of signal transducers and activators of transcription proteins in acute myeloid leukemia blasts. Cancer Res. 58: 3173–3180.

    CAS  PubMed  Google Scholar 

  127. Nowell PC, Hungerford DA. (1960) A minute chromosome in human chronic granulocytic leukemia. Science 132: 1497–1499.

    Google Scholar 

  128. Rowley JD. (1973) A new consistent chromosomal abnormality in chronic myelogenous leukemia identified by quinacrine fluorescence and Giemsa banding. Nature 243: 290–291.

    Article  CAS  PubMed  Google Scholar 

  129. Konopka JB, Watanabe SM, Witte ON. (1984) An alteration of the human c-abl protein in K562 unmasks associated tyrosine kinase activity. Cell 37: 1035–1042.

    Article  CAS  PubMed  Google Scholar 

  130. Shtivelman E, Lifshitz B, Gale RP, Canaani E. (1985) Fused transcript of abl and bcr genes in chronic myelogenous leukemia. Nature 315: 550–554.

    Article  CAS  PubMed  Google Scholar 

  131. Gishizky ML, Witte ON. (1992) Initiation of deregulated growth of multipotent progenitor cells by bcr-abl in vitro. Science 256: 836–839.

    Article  CAS  PubMed  Google Scholar 

  132. Daley GQ, Van Etten RA, Baltimore D. (1990) Induction of chronic myelogenous leukemia in mice by the P21OberIabl gene of the Philadelphia chromosome. Science 247: 824–830.

    Article  CAS  PubMed  Google Scholar 

  133. Daley GQ, Baltimore D. (1988) Transformation of an interleukin 3-dependent hematopoietic cell line by the chronic myelogenous leukemia-specific P210bcr/abl protein. Proc. Natl. Acad. Sci. U.S.A. 85: 9312–9316.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Carlesso N, Frank DA, Griffin JD. (1996) Tyrosyl phosphorylation and DNA-binding activity of STAT proteins in hematopoietic cell lines transformed by Bcr/Abl. J. Exp. Med 183: 811–820.

    Article  CAS  PubMed  Google Scholar 

  135. Frank DA, Varticovski L. (1996) BCR/abl leads to the constitutive activation of Stat proteins, and shares an epitope with tyrosine phosphorylated Stats. Leukemia 10: 1724–1730.

    CAS  PubMed  Google Scholar 

  136. Ilaria RL Jr, Van Etten RA. (1996) P210 and P190BCR/ABL induce the tyrosine phosphorylation and DNA binding activity of multiple specific STAT family members. J. Biol Chem. 271: 31704–31710.

    Article  CAS  PubMed  Google Scholar 

  137. Shuai K, Halpern J, ten Hoeve J, Rao X, Sawyers CL. (1996) Constitutive activation of STAT5 by the BCR-ABL oncogene in chronic myelogenous leukemia. Oncogene 13: 247–254.

    CAS  PubMed  Google Scholar 

  138. Chai SK, Nichols GL, Rothman P. (1997) Constitutive activation of JAKs and STATs in BCR-Abl-expressing cell lines and peripheral blood cells derived from leukemic patients. J. Immunol. 159: 4720–4728.

    CAS  PubMed  Google Scholar 

  139. Kurzrock R, Gutterman J, Talpaz M. (1988) The molecular genetics of Philadelphia chromosome-positive leukemias. N. Engl J. Med. 319: 990–998.

    Article  CAS  PubMed  Google Scholar 

  140. Nieborowska-Skorska M, Wasik MA, Salomoni P, Kitamura T, Calabretta B, Skorski T. (1998) STAT5 activation by Bcr/Abl is dependent on its intact SH3 and SH2 domains and is required for leukemogenesis. Blood 92: 91a.

    Google Scholar 

  141. Rozman C, Montserrat E. (1995) Chronic lymphocytic leukemia. New Engl J. Med. 333: 1052–1057.

    Article  CAS  PubMed  Google Scholar 

  142. Lin J-X. (1995) The role of shared receptor motifs and common Stat proteins in the generation of cytokine pleiotropy and redundancy by IL-2, IL-4, IL-7, IL-13, and IL-15. Immunity 2: 331–339.

    Article  CAS  PubMed  Google Scholar 

  143. Tanner JE, Tosato G. (1992) Regulation of B-cell growth and immunoglobulin gene transcription by interleukin-6. Blood 19: 452–459.

    Google Scholar 

  144. Nielsen M, Kaltoft K, Nordahl M, et al. (1997) Constitutive activation of a slowly migrating isoform of Stat3 in mycosis fungoides: tyrphostin AG490 inhibits Stat3 activation and growth of mycosis fungoides tumor cell lines. Proc. Natl. Acad. Sci. U.S.A. 94: 6764–6769.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Zhang Q, Nowak I, Vonderheid EC, et al. (1996) Activation of Jak/STAT proteins involved in signal transduction pathway mediated by receptor for interleukin 2 in malignant T lymphocytes derived from cutaneous anaplastic large T-cell lymphoma and Sezary syndrome. Proc. Natl Acad. Sci. U.S.A. 93: 9148–9153.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Migone T-S, Lin J-X, Cereseto A, et al. (1995) Constitutively activated Jak-STAT pathway in T cells transformed with HTLV-I. Science 269: 79–81.

    Article  CAS  PubMed  Google Scholar 

  147. Takemoto S, Mulloy JC, Cereseto A, et al. (1997) Proliferation of adult T cell leukemia/lymphoma cells is associated with the constitutive activation of JAK/STAT proteins. Proc. Natl Acad. Sci. U.S.A. 94: 13897–13902.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Xu X, Kang SH, Heidenreich O, Okerholm M, O’Shea JJ, Nerenberg MI. (1995) Constitutive activation of different Jak kinases in human T cell leukemia virus type 1 (HTLV-1) tax protein or virus-transformed cells. J. Clin. Invest. 96: 1548–1555.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Smith D, Buckle GJ, Hafler DA, Frank DA, Hollsberg P. (1999) HTLV-I infected T cells evade the anti-proliferative action of IFN-beta. Virology 257: 314–321.

    Article  CAS  PubMed  Google Scholar 

  150. Kawano M, Hirano T, Matsuda T, et al. (1988) Autocrine generation and requirement of BSF-2/IL-6 for human multiple myelomas. Nature 332: 83–85.

    Article  CAS  PubMed  Google Scholar 

  151. Anderson KC, Jones RM, Morimoto C, Leavitt P, Barut BA. (1989) Response patterns of purified myeloma cells to hematopoietic growth factors. Blood 73: 1915–1924.

    CAS  PubMed  Google Scholar 

  152. Levy Y, Tsapis A, Brouet JC. (1991) Interleukin-6 antisense oligonucleotides inhibit the growth of human myeloma cell lines. J. Clin. Invest. 88: 696–699.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Klein B, Zhang XG, Yang LZ, Bataille R. (1995) Interleukin-6 in human multiple myeloma. Blood 85: 863–874.

    CAS  PubMed  Google Scholar 

  154. Hilbert DM, Kopf M, Mock BA, Kohler G, Rudikoff S. (1995) Interleukin 6 is essential for in vivo development of B lineage neoplasms. J. Exp. Med. 182: 243–248.

    Article  CAS  PubMed  Google Scholar 

  155. Hilbert DM, Migone T-S, Kopf M, Leonard WJ, Rudikoff S. (1996) Distinct tumorigenic potential of abl and raf in B cell neoplasia: abl activates the IL-6 signaling pathway. Immunity 5: 81–89.

    Article  CAS  PubMed  Google Scholar 

  156. Catlett-Falcone R, Landowski TH, Oshiro MM, et al. (1999) Constitutive activation of Stat3 signaling confers resistance to apoptosis in human U266 myeloma cells. Immunity 10: 105–115.

    Article  CAS  PubMed  Google Scholar 

  157. Barut B, Chauhan D, Uchiyama H, Anderson KC. (1993) Interleukin-6 functions as an intracellular growth factor in hairy cell leukemia in vitro. J. Clin. Invest. 92: 2346–2352.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Seymour JF, Talpaz M, Cabanillas F, Wetzler M, Kurzrock R. (1995) Serum interleukin-6 levels correlate with prognosis in diffuse large cell lymphoma. J. Clin. Oncol. 13: 575–582.

    Article  CAS  PubMed  Google Scholar 

  159. Pallard C, Gouilleux F, Charon M, Groner B, Gisselbrecht S, Dusanter-Fourt I. (1995) Interleukin-3, erythropoietin, and prolactin activate a STAT5-like factor in lymphoid cells. J. Biol. Chem. 270: 15942–15945.

    Article  CAS  PubMed  Google Scholar 

  160. Watson CJ, Miller WR. (1995) Elevated levels of members of the STAT family of transcription factors in breast carcinoma nuclear extracts. Br. J. Cancer 71: 840–844.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Garcia R, Yu C-L, Hudnall A, et al. (1997) Constitutive activation of Stat3 in fibroblasts transformed by diverse oncoproteins and in breast carcinoma cells. Cell Growth Differ. 8: 1267–1276.

    CAS  PubMed  Google Scholar 

  162. Sartor CI, Dziubinski ML, Yu CL, Jove R, Ethier SP. (1997) Role of epidermal growth factor receptor and STAT-3 activation in autonomous proliferation of SUM-102PT human breast cancer cells. Cancer Res. 57: 978–987.

    CAS  PubMed  Google Scholar 

  163. Manni A, Trout D, Verderame MF, Beaston-Wimmer PR. (1999) Ornithine decarboxylase (ODC) overexpression activates Src and STAT signaling in MCF-10A human breast epithelial cells. Proc. Am. Assoc. Cancer Res. 40: 99.

    Google Scholar 

  164. Endo S, Zeng Q, He Y, et al. (1999) Increased Stat3 activation in head and neck tumors in vivo. Proc. Am. Assoc. Cancer Res. 40: 336.

    Google Scholar 

  165. Rubin Grandis J, Chakraborty A, Meinem MF, Zeng Q, Tweardy DJ. (1997) Inhibition of epidermal growth factor receptor gene expression and function decreases proliferation of head and neck squamous carcinoma but not normal mucosal epithelial cells. Oncogene 15: 409–416.

    Article  CAS  PubMed  Google Scholar 

  166. Zhou MY, Zeng Q, Drenning SD, Rubin Grandis J. (1999) Differential activation of STAT5 isoforms and growth control in head and neck cancer. Proc. Am. Assoc. Cancer Res. 40: 335.

    Google Scholar 

  167. Reddy MVR, Chaturvedi P, Reddy EP. (1999) Src kinase mediated activation of STAT-3 plays an essential role in the proliferation and oncogenicity of human breast, prostate and ovarian carcinomas. Proc. Am. Assoc. Cancer Res. 40: 376.

    Google Scholar 

  168. Kirkwood JM, Farkas DL, Chakraborty A, et al. (1999) Systemic interferon-α (IFN-α) treatment leads to Stat3 inactivation in melanoma precursor lesions. Mol. Med. 5: 11–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Shirota K, LeDuy L, Yuan SY, Jothy S. (1990) Interleukin-6 and its receptor are expressed in human intestinal epithelial cells. Virchows Arch B Cell Pathol 58: 303–308.

    Article  CAS  Google Scholar 

  170. Burkitt DP. (1971) Epidemiology of cancer of the colon and rectum. Cancer 28: 3–13.

    Article  CAS  PubMed  Google Scholar 

  171. Howe GR, Benito E, Castelleto R, et al. (1992) Dietary intake of fiber and decreased risk of cancers of the colon and rectum: evidence from the combined analysis of 13 case-control studies. J. Natl. Cancer Inst. 84: 1887–1896.

    Article  CAS  PubMed  Google Scholar 

  172. Reddy BS, Hedges AR, Laakso K, Wynder EL. (1978) Metabolic epidemiology of large bowel cancer: fecal bulk and constituents of high-risk North American and low-risk Finnish population. Cancer 42: 2832–2838.

    Article  CAS  PubMed  Google Scholar 

  173. Trock B, Lanza E, Greenwald P. (1990) Dietary fiber, vegetables, and colon cancer: critical review and meta-analyses of the epidemiologic evidence. J. Natl Cancer Inst. 82: 650–661.

    Article  CAS  PubMed  Google Scholar 

  174. Mizutani Y, Bonavida B, Koishihara Y, Akamatsu K, Ohsugi Y, Yoshida O. (1995) Sensitization of human renal cell carcinoma cells to cis-diamminedichloroplatinum(II) by anti-interleukin 6 monoclonal antibody or anti-interleu-kin 6 receptor monoclonal antibody. Cancer Res. 55: 590–596.

    CAS  PubMed  Google Scholar 

  175. Borsellino N, Belldegrun A, Bonavida B. (1995) Endogenous interleukin 6 is a resistance factor for cis-diamminedichloroplatinum and etoposide-mediated cytotoxicity of human prostate carcinoma cell lines. Cancer Res. 55: 4633–4639.

    CAS  PubMed  Google Scholar 

  176. Kowalczyk J, Sandberg AA. (1983) A possible subgroup of ALL with 9p-. Cancer Genet. Cytogenet. 9: 383–388.

    Article  CAS  PubMed  Google Scholar 

  177. Chilcote RR, Brown E, Rowley JD. (1985) Lymphoblastic leukemia with features associated with abnormalities of the short arm of chromosome 9. N. Engl J. Med. 313: 286–289.

    Article  CAS  PubMed  Google Scholar 

  178. Pollack C, Hagemeijer A. (1987) Abnormalities of the short arm of chromosome 9 with partial loss of material in hematological disorders. Leukemia 1: 541–548.

    Google Scholar 

  179. Peeters P, Raynaud SD, Cools J, et al. (1997) Fusion of Tel the ETS-variant gene 6 (ETV6), to the receptor-associated kinase Jak2 as a result of t(9;12) in a lymphoid and t(9;15;12) in a myeloid leukemia. Blood 90: 2535–2540.

    CAS  PubMed  Google Scholar 

  180. Lacronique V, Boureux A, Della Valle V, et al. (1997) A TEL-JAK2 fusion protein with constitutive kinase activity in human leukemia. Science 278: 1309–1312.

    Article  CAS  PubMed  Google Scholar 

  181. Wasylyk B, Hahn SL, Giovane A. (1993) The Ets family of transcription factors. Eur. J. Biochem. 211: 7–18.

    Article  CAS  PubMed  Google Scholar 

  182. Ho JM-Y, Beattie BK, Squire JA, Frank DA, Barber DL. (1999) Fusion of the ets transcription factor TEL to Jak2 results in constitutive Jak-Stat signaling. Blood 93: 4354–4364.

    CAS  PubMed  Google Scholar 

  183. Schwaller J, Frantsve J, Aster J, et al. (1998) Transformation of hematopoietic cell lines to growth-factor independence and induction of a fatal myelo- and lymphoproliferative disease in mice by retrovirally transduced TEL/JAK2 fusion genes. EMBO J. 17: 5321–5333.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Janssen JW, Ridge SA, Papadopoulos P, et al. (1995) The fusion of TEL and ABL in human acute lymphoblastic leukemia is a rare event. Br. J. Haematol. 90: 222–224.

    Article  CAS  PubMed  Google Scholar 

  185. Golub TR, Goga A, Barker G, et al. (1996) Oligomerization of the ABL tyrosine kinase by the ETS protein TEL in human leukemia. Mol. Cell. Biol. 16: 4107–4116.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Morris SW, Kirstein MN, Valentine MB, et al. (1994) Fusion of a kinase gene, ALK, to a nucleolar protein gene, NPM, in non-Hodgkin’s lymphoma. Science 263: 1281–1284.

    Article  CAS  PubMed  Google Scholar 

  187. Xiao S, Nalabolu SR, Aster JC, et al. (1998) FGFR1 is fused with a novel zinc-finger gene, ZNF198, in the t(8;13) leukaemia/lymphoma syndrome. Nat. Genet. 18: 84–87.

    Article  CAS  PubMed  Google Scholar 

  188. Golub TR, Barker GF, Lovett M, Gilliland DG. (1994) Fusion of PDGF receptor beta to a novel ets-like gene, tel, in chronic myelomonocytic leukemia with t(5;12) chromosomal translocation. Cell 77: 307–316.

    Article  CAS  PubMed  Google Scholar 

  189. Abe A, Emi N, Tanimoto M, Terasaki H, Marunouchi T, Saito H. (1997) Fusion of the platelet-derived growth factor receptor β to a novel gene CEV14 in acute myelogenous leukemia after clonal evolution. Blood 90: 4271–4277.

    CAS  PubMed  Google Scholar 

  190. Ross T, Bernard O, Berger R, Gilliland DG. (1998) Fusion of the Huntington interacting protein 1 to platelet-derived growth factor β receptor (PDGFβR) in chronic myelomonocytic leukemia with t(5;7)(q33;qll.2). Blood 91: 4419–4426.

    CAS  PubMed  Google Scholar 

  191. Vignais ML, Sadowski HB, Watling D, Rogers NC, Gilman M. (1996) Platelet-derived growth factor induces phosphorylation of multiple JAK family kinases and STAT proteins. Mol. Cell. Biol. 16: 1759–1769.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Valgeirsdottir S, Paukku K, Silvennoinen O, Heldin CH, Claesson-Welsh L. (1998) Activation of Stat5 by platelet-derived growth factor (PDGF) is dependent on phosphorylation sites in PDGF beta-receptor juxtamembrane and kinase insert domains. Oncogene 16: 505–515.

    Article  CAS  PubMed  Google Scholar 

  193. Carroll M, Tomasson MH, Barker GF, Golub TR, Gilliland DG. (1996) The TEL/platelet-derived growth factor beta receptor (PDGF beta R) fusion in chronic myelomonocytic leukemia is a transforming protein that self-associates and activates PDGF beta R kinase-dependent signaling pathways. Proc. Natl. Acad. Sci. U.S.A. 93: 14845–14850.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Joussert C, Carron C, Boureux A, et al. (1997) A domain of TEL conserved in a subset of ETS proteins defines a specific oligomerization interface essential to the mitogenic properties of the TEL-PDGFR beta oncoprotein. EMBO J. 16: 69–82.

    Article  Google Scholar 

  195. Palmer AM, Mahajan S, Frank D, Gilliland DG, Carroll M. (1997) The TEL-PDGFβR transforming protein activates STATL Blood 90: 178a.

    Google Scholar 

  196. Garcia R, Jove R. (1998) Activation of STAT transcription factors in oncogenic tyrosine kinase signaling. J. Biomed. Sci. 5: 79–85.

    Article  CAS  PubMed  Google Scholar 

  197. Danial NN, Pernis A, Rothman PB. (1995) JakSTAT signaling induced by the v-abl oncogene. Science 269: 1875–1877.

    Article  CAS  PubMed  Google Scholar 

  198. Danial NN, Losman JA, Lu T, et al. (1998) Direct interaction of Jakl and v-Abl is required for v-Abl-induced activation of STATs and proliferation. Mol. Cel. Biol. 18: 6795–6804.

    Article  CAS  Google Scholar 

  199. Yu C, Meyer DJ, Campbell GS, et al. (1995) Enhanced DNA-binding activity of a Stat3-related protein in cells transformed by the Src oncoprotein. Science 269: 81–83.

    Article  CAS  PubMed  Google Scholar 

  200. Cao X, Tay A, Guy GR, Tan YH. (1996) Activation and association of Stat3 with Src in v-Srctransformed cell lines. Mol. Cell. Biol. 16: 1595–1603.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Chaturvedi P, Sharma S, Reddy EP. (1997) Abrogation of interleukin-3 dependence of myeloid cells by the v-src oncogene requires SH2 and SH3 domains which specify activation of STATs. Mol. Cell. Biol. 17: 3295–3304.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Bromberg JF, Horvath CM, Besser D, Lathem WW, Darnell JE Jr. (1998) Stat3 activation is required for cellular transformation by v-src. Mol. Cell. Biol. 18: 2553–2558.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Caldenhoven E, van Dijk TB, Solari R, et al. (1996) STAT3beta, a splice variant of transcription factor STAT3, is a dominant negative regulator of transcription. J. Biol. Chem. 271: 13221–13227.

    Article  CAS  PubMed  Google Scholar 

  204. Turkson J, Bowman T, Garcia R, Caldenhoven E, de Groot RP, Jove R. (1998) Stat3 activation by src induces specific gene regulation and is required for cell transformation. Mol. Cell. Biol. 18: 2545–2552.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Campbell GS, Yu C-L, Jove R, Carter-Su C. (1997) Constitutive activation of Jakl in Srctransformed cells. J. Biol. Chem. 272: 2591–2594.

    Article  CAS  PubMed  Google Scholar 

  206. Yu CL, Jove R, Burakoff SJ. (1997) Constitutive activation of the Janus kinase-STAT pathway in T lymphoma overexpressing the Lck protein tyrosine kinase. J. Immunol. 159: 5206–5210.

    CAS  PubMed  Google Scholar 

  207. Lund TC, Garcia R, Medveczky MM, Jove R, Medveczky PG. (1997) Activation of STAT transcription factors by herpesvirus Saimiri Tip-484 requires p561ck. J. Virol. 71: 6677–6682.

    CAS  PubMed  PubMed Central  Google Scholar 

  208. Lund TC, Prator PC, Medveczky MM, Medveczky PG. (1999) The Lck binding domain of herpesvirus saimiri tip-484 constitutively activates Lck and STAT3 in T cells. J. Virol. 73: 1689–1694.

    CAS  PubMed  PubMed Central  Google Scholar 

  209. Zong C, Yan R, August A, Darnell JE Jr, Hanafusa H. (1996) Unique signal transduction of Eyk: constitutive stimulation of the JAK-STAT pathway by an oncogenic receptor-type tyrosine kinase. EMBO J. 15: 4515–4525.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Besser D, Bromberg JF, Darnell JE Jr, Hanafusa H. (1999) A single amino acid substitution in the v-Eyk intracellular domain results in activation of Stat3 and enhances cellular transformation. Mol. Cell. Biol. 19: 1401–1409.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Shimoda K, van Deursen J, Sangster MY, et al. (1996) Lack of IL-4-induced Th2 response and IgE class switching in mice with disrupted Stat6 gene. Nature 380: 630–633.

    Article  CAS  PubMed  Google Scholar 

  212. Takeda K, Tanaka T, Shi W, et al. (1996) Essential role of Stat6 in IL-4 signalling. Nature 380: 627–630.

    Article  CAS  PubMed  Google Scholar 

  213. Robertson MJ, Cochran KJ, Cameron C, Le JM, Tantravahi R, Ritz J. (1996) Characterization of a cell line, NKL, derived from an aggressive human natural killer cell leukemia. Exp. Hematol. 24: 406–415.

    CAS  PubMed  Google Scholar 

  214. Frank DA, Robertson M, Bonni A, Ritz J, Greenberg ME. (1995) IL-2 signaling involves the phosphorylation of novel Stat proteins. Proc. Natl. Acad. Sci. U.S.A. 92: 7779–7783.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Teglund S, McKay C, Schuetz E, et al. (1998) Stat5a and Stat5b proteins have essential and nonessential, or redundant, roles in cytokine responses. Cell 93: 841–850.

    Article  CAS  PubMed  Google Scholar 

  216. Kaplan DH, Shankaran V, Digne AS, et al. (1998) Demonstration of an interferon γ-dependent tumor surveillance system in immunocompetent mice. Proc. Natl. Acad. Sci. U.S.A. 95: 7556–7561.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Takeda K, Noguchi K, Shi W, et al. (1997) Targeted disruption of the mouse STAT3 gene leads to early embryonic lethality. Proc. Natl. Acad. Sci. USA. 94: 3801–3804.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Peppard JV, Loo P, Sills MA, Munster D, Pomponi SA, Wright AE. (1996) Characterization of an interleukin 6 cytokine family antagonist protein from a marine sponge, Callyspongia sp. J. Biol. Chem. 271: 7281–7284.

    Article  CAS  PubMed  Google Scholar 

  219. Savino R, Lahm A, Salvati AL, et al. (1994) Generation of interleukin-6 antagonists by molecular-modeling guided mutagenesis of residues important for gp130 activation. EMBO J. 13: 1357–1367.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  220. Savino R, Ciapponi L, Lahm A, et al. (1994) Rational design of a receptor super-antagonist of human interleukin-6. EMBO J. 13: 5863–5870.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Kopf M, Baumann H, Freer G, et al. (1994) Impaired immune and acute-phase responses in interleukin-6-deficient mice. Nature 368: 339–342.

    Article  CAS  PubMed  Google Scholar 

  222. Klein B, Wijdenes J, Zhang XG, et al. (1991) Murine antiinterleukin-6 monoclonal antibody therapy for a patient with plasma cell leukemia. Blood 78: 1198–1204.

    CAS  PubMed  Google Scholar 

  223. Meydan N, Grunberger T, Dadi H, et al. (1996) Inhibition of acute lymphoblastic leukaemia by a Jak-2 inhibitor. Nature 379: 645–648.

    Article  CAS  PubMed  Google Scholar 

  224. Buchdunger E, Mett H, Trinks U, et al. (1995) 4, 5-bis(4-fluoroanilino)phthalimide: a selective inhibitor of the epidermal growth factor receptor signal transduction pathway with potent in vivo antitumor activity. Clin. Cancer Res. 1: 813–821.

    CAS  PubMed  Google Scholar 

  225. Buchdunger E, Zimmermann J, Mett H, et al. (1996) Inhibition of the ABL protein-tyrosine kinase in vitro and in vivo by a 2-phenylamino-pyrimidine derivative. Cancer Res. 56: 100–104.

    CAS  PubMed  Google Scholar 

  226. Mologni L, Cleris L, Marchesi E, et al. (1999) In vivo eradication of human BCR/ABL-positive leukemia cells with an ABL kinase inhibitor. J. Natl. Cancer Inst. 91: 163–168.

    Article  PubMed  Google Scholar 

  227. Carroll M, Ohno-Jones S, Tamura S, et al. (1997) CGP 57148, a tyrosine kinase inhibitor, inhibits the growth of cells expressing BCR-ABL, TEL-ABL, and TEL-PDGFR fusion proteins. Blood 90: 4947–4952.

    CAS  PubMed  Google Scholar 

  228. Traxler PM, Fyret P, Mett H, Buchdunger E, Meyer T, Lydon N. (1996) 4-(phenylamino)pyrrolopyrimidines: potent and selective, ATP site directed inhibitors of the EGF-receptor tyrosine kinase. J. Med. Chem. 39: 2285–2292.

    Article  CAS  PubMed  Google Scholar 

  229. Druker BJ, Tamura S, Buchdunger E, et al. (1996) Effects of a selective inhibitor of the Abl tyrosine kinase on the growth of Bcr-Abl positive cells. Nat. Med. 2: 561–566.

    Article  CAS  PubMed  Google Scholar 

  230. Tomasson MH, Williams IR, Hasserjian R, et al. (1999) TEL/PDGFbetaR induces hematologic malignancies in mice that respond to a specific tyrosine kinase inhibitor. Blood 93: 1707–1714.

    CAS  PubMed  Google Scholar 

  231. Frank DA, Mahajan S, Ritz J. (1999) Fludarabine-induced immunosuppression is associated with inhibition of STAT1 signaling. Nat. Med. 5: 444–447.

    Article  CAS  PubMed  Google Scholar 

  232. Bromberg JF, Fan Z, Brown C, Mendelsohn J, Darnell JE Jr. (1998) Epidermal growth factor-induced growth inhibition requires Stat1 activation. Cell Growth Differ. 9: 505–512.

    CAS  PubMed  Google Scholar 

  233. Kirkwood JM, Strawdermann MH, Ernstoff MS, Smith TJ, Borden EC, Blum RH. (1996) Interferon alfa-2b adjuvant therapy of high-risk resected cutaneous melanoma: the Eastern Cooperative Oncology Group Trial EST 1684. J. Clin. Oncol. 14: 7–17.

    Article  CAS  PubMed  Google Scholar 

  234. Carson WE. (1998) Interferon-α-induced activation of signal transducer and activator of transcription proteins in malignant melanoma. Clin. Cancer Res. 4: 2219–2228.

    CAS  PubMed  Google Scholar 

  235. Wong LH, Krauer KG, Hatzinisiriou I, et al. (1997) Interferon-resistant human melanoma cells are deficient in ISGF3 components, STAT1, STAT2, and p48-ISGF3gamma. J. Biol Chem. 272: 28779–28785.

    Article  CAS  PubMed  Google Scholar 

  236. Sun WH, Pabon C, Alsayed Y, et al. (1998) Interferon-alpha resistance in a cutaneous T-cell lymphoma cell line is associated with lack of STAT1 expression. Blood 91: 570–576.

    CAS  PubMed  Google Scholar 

  237. Bromberg JF, Horvath CM, Wen Z, Schreiber RD, Darnell JE Jr. (1996) Transcriptionally active Stat1 is required for the antiproliferative effects of both interferon alpha and interferon gamma. Proc. Natl Acad. Sci. U.S.A. 93: 7673–7678.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  238. Warreil RP Jr, Frankel SR, Miller WJJ, et al. (1991) Differentiation therapy of acute promyelocytic leukemia with triretinoin (all-trans retinoic acid). N Engl J. Med. 324: 1385–1391.

    Article  Google Scholar 

  239. Gianni M, Terao M, Fortino I, et al. (1997) Stat1 is induced and activated by all-trans retinoic acid in acute promyelocytic leukemia cells. Blood 89: 1001–1012.

    CAS  PubMed  Google Scholar 

  240. Weihua X, Kolla V, Kalvakolanu DV. (1997) Modulation of interferon action by retinoids. Induction of murine STAT1 gene expression by retinoic acid. J. Biol Chem. 272: 9742–9748.

    Article  CAS  PubMed  Google Scholar 

  241. Kalvakolanu DV, Sen GC. (1993) Differentiation-dependent activation of interferon-stimulated gene factors and transcription factor NF-kappa B in mouse embryonal carcinoma cells. Proc. Natl. Acad. Sä. U.S.A. 90: 3167–3171.

    Article  CAS  Google Scholar 

  242. Kolla V, Lindner DJ, Weihua X, Borden EC, Kalvakolanu DV. (1996) Modulation of interferon (IFN)-inducible gene expression by retinoic acid. Up-regulation of Statl protein in IFN-unresponsive cells. J. Biol Chem. 271: 10508–10514.

    Article  CAS  PubMed  Google Scholar 

  243. Matikainen S, Ronni T, Lehtonen A, et al. (1997) Retinoic acid induces signal transducer and activator of transcription (STAT) 1, STAT2, and p48 expression in myeloid leukemia cells and enhances their responsiveness to interferons. Cell Growth Differ. 8: 687–698.

    CAS  PubMed  Google Scholar 

  244. Aragane Y, Kulms D, Luger TA, Schwarz T. (1997) Down-regulation of interferon γ-activated STAT1 by UV light. Proc. Natl Acad. Sci. USA. 94: 11490–11495.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Look DC, Roswit WT, Frick AG, et al. (1998) Direct suppression of Stat1 function during adenoviral infection. Immunity 9: 871–880.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

This work was supported by NIH grant CA79547, the Cancer Research Foundation of America, and the Brent Leahey Fund.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David A. Frank.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Frank, D.A. STAT Signaling in the Pathogenesis and Treatment of Cancer. Mol Med 5, 432–456 (1999). https://doi.org/10.1007/BF03403538

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/BF03403538