0163-769X/91/1203-0208$03.00/0 Endocrine Reviews Copyright © 1991 by The Endocrine Society

Vol. 12, No. 3 Printed in U.S.A.

LIF: Not Just a Leukemia Inhibitory Factor RAZELLE KURZROCK, ZEEV ESTROV, MEIR WETZLER, JORDAN U. GUTTERMAN, AND MOSHE TALPAZ Department of Clinical Immunology and Biological Therapy, The University of Texas M.D. Anderson Cancer Center, Houston, Texas 77030

I. Introduction

A. Nomenclature

L

EUKEMIA inhibitory factor (LIF) is a pleiotropic cytokine with the ability to influence the function and/or development of seemingly unrelated body elements (Table 1): blood cells, embryonal cells, hepatocytes, neurons, adipose tissues, and bone. These diverse properties account for the bewildering array of names under which this factor has been published (Table 2). Although the most common designation is leukemia inhibitory factor, this molecule can promote both suppression and stimulation of proliferation and/or differentiation depending on the leukemic cell line examined. In addition, LIF may serve as part of the body's energy storage communication network. In this regard, some of the abnormalities that manifest in mammalian hosts with chronic illness include increased serum acute phase proteins and a wasting diathesis accompanied by hypertriglyceridemia; the latter is caused by suppression of lipoprotein lipase, the key enzyme of triglyceride metabolism. It is therefore of interest that LIF induces cachexia in mice, inhibits lipoprotein lipase, and stimulates hepatocyte release of acute phase proteins. In a different arena, the potent effects of LIF on embryonal stem cells suggest a critical role for this molecule in the earliest stages of embryogenesis. Finally, LIF also promotes remodeling of bone and functions as a neuronal differentiation factor. It remains unclear how one molecule can affect distinct organ systems without simultaneous side effects on its other target tissues, although recent data demonstrating that LIF exists both as a diffusible form and a form incorporated into the extracellular matrix (18) suggest that immobilization may serve to restrict its spheres of influence. In this review, we will discuss the multifaceted aspects of LIF functions and implications for understanding the cytokine network.

The diverse nomenclature of LIF reflects the panoply of properties that led researchers in different fields to independent discovery of this protein (Table 2). This explains the distinct and contradictory designations such as differentiation-inducing factor, differentiation-retarding factor, hepatocyte- stimulating factor III, cholinergic neuronal differentiation factor, and melanoma-derived lipoprotein lipase inhibitor I. The most common term for this molecule in current use is, however, LIF. B. Cellular sources Cellular sources for LIF are listed in Table 3. They include certain alloreactive T cell clones, many tumor cell lines, osteoblasts, cultured keratinocytes, and thymic epithelium. In the mouse system, LIF is synthesized in extraembryonic tissue and may therefore regulate embryonic tissues during development (24). Further, embryonic stem cells themselves may express small amounts of LIF, and expression is increased with differentiation induction (25). Recently, we have found that LIF messenger RNA is constitutively expressed by longterm cultures of bone marrow adherent cells, a system felt to be the in vitro analog of the stroma of the bone marrow microenvironment (26). Unstimulated monocytes and lymphocytes do not express LIF, whereas cultured skin fibroblasts and stimulated monocytes do (26, 31); it is therefore the fibroblast-like component of stromal cultures that are probably the LIF producers. Phytohemagglutin-stimulated T lymphocytes also produce LIF (26). C. Molecular biology and biochemistry The reported mol wts of LIF range from 38 to 67K. Much of this heterogeneity can be ascribed to variable glycosylation as removal of the carbohydrate moiety reduces the mol wt to 25K (34, 35). Recombinant forms of LIF that display a different pattern of glycosylation

* Research conducted, in part, by the Clayton Foundation for Research. Dr. Gutterman is a Senior Clayton Foundation Investigator. Address requests for reprints to: Razelle Kurzrock, M.D., Department CIBT, Box 41, University of Texas M.D. Anderson Cancer Center, 1515 Holcombe Boulevard, Houston, Texas 77030.

208

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

August, 1991

LEUKEMIA INHIBITORY FACTOR

(yeast-derived) or no glycosylation (Escherichia coliderived) are still active (1, 35). Both murine and human LIF have been cloned (1, 3537), and the complete nucleotide sequence for their genes is 8.7 and 7.6 kilobase pairs, respectively (38). Both genes comprise three exons, two introns, and an unusually long 3'-untranslated region (3.2 kilobase pairs) (38). The LIF transcript is 4.2 kilobases in length. The murine and human clones that have been isolated predict a sequence with 179 residues for the mature protein (1, 2, 35, 36) and a 79% amino acid sequence identity between the two species. This primary form of LIF is a biologically active, "diffusible" glycoprotein. In addition to the molecule described above, Rathjen and co-workers (18) have recently shown that an alternate "immobilized" form of LIF exists; the latter is incorporated into the extracellular matrix. These different forms are a consequence of the expression of alternative transcripts that diverge throughout the first exon and use distinct promoters. Diffusible and immobilized LIF variants are therefore encoded by mRNAs that are differentially spliced at the exon 1/exon 2 boundary. A transcript termed D (diffusible), encodes the diffusible form; a transcript termed M (matrix-associated) is a result of splicing a novel 5' exon to exons 2 and 3 of the LIF trancription unit, and encodes the matrix-associated form. It should be noted that the two transcripts comigrate on agarose gels and can therefore be distinguished by ribonuclease protection analysis but not by Northern blots. The differential localization of the two forms of LIF can be explained by the modular organization of the

209

TABLE 2. LIF nomenclature

Name

Abbreviation

References

Leukemia inhibitory factor Differentiation factor Differentiation inhibitory activity Differentiation inducing factor Differentiation retarding factor Human IL for DA cells Hepatocyte stimulating factor III Melanoma-derived lipoprotein lipase inhibitor I 9. Cholinergic neuronal differentiation factor

LIF D-factor DIA DIF DRF HILDA HSFIII MLPLI

1,10 11-13 14 12 14, 15 2, 16, 17 8 7

1. 2. 3. 4. 5. 6. 7. 8.

gene. Exons 2 and 3 produce the core hydrophobic secretory sequences and biologically active protein, while extracellular localization is specified by the first exon. Incorporation of mature, functional LIF into the extracellular matrix is therefore directed by changes in the amino terminal of the primary translation product. D. Chromosomal localization LIF has been localized to chromosome 22ql2 in man (39) and 11A1-A2 in the mouse (40). The 22ql2 region in man is of interest because the majority of Ewings sarcomas have a translocation involving this cytogenetic band [(t(ll;22)(q24;ql2)]. However, analyses using somatic cell hybrids and pulsed field gel electrophoresis suggest that the LIF gene is distal to the breakpoint and not close enough to be involved (41). E. Biological properties

TABLE 1. Biological properties of LIF Hematopoiesis In vitro, induces differentiation and inhibits proliferation, or stimulates proliferation without differentiation depending on the cells analyzed (1, 2) In vivo, increases erythroid and megakaryocytic elements while decreasing lymphocytes in mice (3)

LIF, like many other members of the cytokine family, affects diverse body tissues. In the case of LIF, actions on embryonal cells, hematopoietic elements, liver cells, nervous system development, and bone growth have been documented (Table 1). II. Embryogenesis

Embryonal cells Inhibits in vitro differentiation of totipotent mouse embryonal (ES) stem cells, without affecting proliferation (4, 5) Bone Stimulates bone remodeling in vitro (6) Lipid metabolism Inhibits lipoprotein lipase in vitro and may therefore mediate, in part, cachexia (7) Hepatic Stimulates acute phase protein synthesis in hepatocytes (8) Nervous system Directs choice of neurotransmitter phenotype made by sympathetic neurons in vitro (9)

Extraembryonic tissues lie in close proximity to the embryo and secrete factors that influence the development of the latter element. In this regard, Conquet and Brulet (24) have demonstrated that LIF gene transcription is discernible in the extraembryonic tissue of 7.5day, and in the placenta of 9.5-, 10.5-, and 12.5-day (midgestation) postcoitum embryos. LIF transcripts were also detected at the preimplantation mouse blastocyst stage (3.5 days post coitum) (24). The expression of LIF in blastocysts before the appearance of hematopoiesis suggests that this cytokine may regulate embryonal stem cells or the development of trophoblasts (42), and its continued presence in midgestation further implicates this molecule in embryogenesis. A study by Rathjen and

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

KURZROCK ETAL.

210 TABLE 3. Cellular sources of LIF

Mouse Krebs II cells (19, 20) Mitogen-treated splenocytes (12) Ehrlich ascites cells (11) L929 leukemia cells (21) Neutrophils and macrophages of peritoneal infiltrates (22) T cell lines (1) Osteoblast cell line MC3T3E1 (23) Extraembryonic tissue (24) Embryonic stem (ES) cells [differentiation by retinoic acid or 3methoxybenzamide increases LIF levels (25)] Mouse embryonic fibroblasts after induction by IL-la, TGF/3, and bFGF (25) Human Cultures of bone marrow stromal cells (LIF mRNA is expressed constitutively, and expression can be increased by IL-la, IL-1/3, TNFa, and TGF0) (26) Thymic epithelial cells (27) T cell clones (alloreactive) from lymphocytes rejecting kidney allografts (2,10, 17) Cultured keratinocytes (28) Tumor cell lines including 5637 (bladder carcinoma), SVK14 (SV40 transformed keratinocyte), NCI-H23 (lung adenocarcinoma), HLFa (epidermoid carcinoma), SW948 and HRT18 (colon adenocarcinoma), MIA PACA (pancreatic carcinoma), C32 (amelanotic melanoma) HBL 100 (SV-40-transformed mammary epithelium), MDA (breast adenocarcinoma) (29), SEKI (melanoma) (7), COLO-16 (squamous carcinoma) (8), THP-1 (monocytic leukemia) (30) Phytohemagglutinin stimulated T cells (26) Activated monocytes (31) Other Buffalo rat liver cells (14) Rat Yoshida sarcoma cells (32) Rat heart cells (9) Rat osteoblasts (retinoic acid, TNFa, and, in some experiments, TGFjS increases LIF levels) (33)

co-workers (25) on relative expression of LIF D and M transcripts in egg cylinder-stage mouse embryos showed that neither transcript was detectable in the 6.5 or 7.5 day decidua nor in the 7.5 day embryonic regions, and only very low levels of the M transcript were discerned in the 6.5 day embryonic region. However, both D and M transcripts were found at relatively high levels in the extraembryonic region, with the D form approximately 5-fold more abundant than the M form. Expression of both transcripts was greater at 6.5 days than at 7.5 days, with the M transcript barely detectable at the latter timepoint. These data establish both temporal and tissue-specific regulation of LIF D and M transcripts. The activity of LIF has also been studied with the use of in vitro model systems. Embryonal carcinoma (EC) cells, the stem cells of teratocarcinomas, have been exploited for several years to investigate differentiation and development during embryogenesis. Although these cells demonstrate many phenotypic characteristics of

Vol. 12, No. 3

early embryos, they are cytogenetically abnormal, display only limited differentiation potential, and can induce tumors when introduced into embryos. In contrast, embryonal stem (ES) cells, the pluripotent cell lines derived from preimplantation embryos, develop normally and can contribute to all somatic and germ lineages. It has been known for several years that EC and ES cells require the presence of a feeder layer to prevent their differentiation in vitro (43). The feeder factor maintains the embryonal cells (even after multiple passages) in the totipotent state without affecting proliferation (44; reviewed in Ref. 45); it was initially designated differentiation-inhibitory activity or differentiation-retarding factor (14, 15) and is now known to be identical to LIF (4, 5). The above data suggest that LIF functions as a developmental regulatory factor in vivo. How can a pleiotropic molecule that evokes protean responses in different cell types orchestrate complex developmental processes? Some of the answers to this question may lie in the work of Rathjan and co-workers (25); they showed that relative production of the diffusible and matrix-associated, immobilized forms of LIF are developmentally programmed and can be modulated by other bioregulatory molecules. To expand, undifferentiated ES cells express LIF transcript M, albeit at low levels. The possibility of an autocrine process during development of the inner cell mass/primitive ectoderm is thus implied. The level of expression after differentiation is dependent on the particular developmental program. Differentiation induction by exposure to retinoic acid or 3-methoxybenzamide results in a significant increase in steady state levels of both LIF transcripts, whereas only a modest induction occurs in cells that differentiate because of the withdrawal of LIF. These investigators also studied the retinoic acid-induced differentiation of three EC lines. These cells mature into various embryonic cell types whose phenotype varies with the parental cell line; the morphologically distinct progeny exhibit different patterns of LIF expression. Such disparities in relative expression of the two forms of LIF suggest that developmental regulation of the two promoters and/or processing of the two primary transcripts is autonomous. It is therefore conceivable that the diffusible and matrix-associated forms of LIF are produced in different embryonal compartments. Further, the demonstration of enhanced LIF expression in differentiating ES cells suggests a feedback mechanism for stem cell renewal accompanying differentiation (25). The salient characteristic of this model is that early activation of LIF expression by differentiated progeny results in the inhibition of subsequent differentiation and stimulation of self-renewal. One wonders whether such a phenomenon also participates in the control of the differentiation/self renewal program of

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

LEUKEMIA INHIBITORY FACTOR

August, 1991

other systems in which LIF has been implicated, i.e. hematopoiesis. In addition to being developmentally regulated, the expression of LIF is modulated by other bioregulatory molecules including transforming growth factor (TGF)/3, interleukin-1 (IL-1), bovine fibroblast growth factor (bFGF), and steroids. TGF/3 and FGF induce LIF expression in embryonic fibroblasts within 30 and 60 min, respectively (25). The rapid kinetics of this effect and the fact that it proceeds in the presence of protein synthesis inhibition indicate that LIF is a primary response gene and its induction is not mediated via other gene products. Analysis of relative expression of the two transcripts reveals that, in response to some factors (bFGF and IL-1 a), both transcripts increase, whereas, in response to other factors (TGF/3), the increase in LIF D transcript is disproportionately high. In addition, dexamethasone inhibits induction of LIF by IL-la but not by bFGF. The latter result establishes the existence of negative regulators for LIF, a phenomenon that may play a crucial role in restricting physiological expression of LIF. Combinations of a variety of specific inducers and suppressors of LIF coupled with independent developmental programming of the two LIF promoters may therefore contribute to precise regulation of the activity of the two forms of this molecule. Finally, differentiation-based activation of LIF expression ensures stem cell renewal, and hence protects the stem cell pool from exhaustion.

III. Hematopoiesis A. In vitro LIF has a myriad of hematologic effects, and specific effects vary from cell line to cell line (1-3,10, 11,16-22, 34, 46-58) (Table 4). Additionally, some of the findings of different investigators remain to be reconciled, and the data concerning LIF's effects on normal human hematopoietic progenitors are still fraught with uncertainties. LIF was originally characterized by virtue of its ability to induce differentiation of the murine myeloid leukemia cell line Ml (1, 20, 36, 46), a property which it shares with MGI-2 (now known as IL-6) (47). LIF and IL-6 activity have been compared on Ml cells and a second murine myelomonocytic cell line WEHI 3B D+. Despite LIF's potent effects as an inducer of macrophage differentiation of Ml cells (48), it has no effect on WEHI 3B D+ cells, and the latter have no LIF receptors (10, 48). In contrast, IL-6 actively differentiates both Ml and WEHI 3B D+ cells (47, 49). The action of LIF on Ml cells is very rapid, with a decrease in clonogenic cells evident within 24 h of LIF exposure, and significant differentiation induced within 48 h (48). Addition of Mcolony stimulating factor (CSF) to Ml cultures contain-

211

TABLE 4. Hematopoietic effects of LIF In vitro Murine Induces macrophage and granulocyte differentiation and inhibits proliferation of Ml murine myeloid leukemia cells (1) Stimulates proliferation without differentiation of murine IL-3 dependent DA-la leukemia cells (2)

In vivo Blood | platelets Bone marrow [ lymphocytes f erythroid cells | megakaryocytic cells (3)

Enhances IL-3 induced stimulation of megakaryocyte colony formation (58) Human Enhances IL-3 dependent colony formation of very primitive blast colony forming cells (53) and of megakaryocytes (58) In combination with G-CSF, GM-CSF, or IL-6 suppresses clonogenicity of U937 and HL-60 myeloid leukemia cells (51, 52)

ing LIF partially abrogates the reduction in colony number and size that is a consequence of LIF exposure, suggesting that the effects of LIF on Ml cells may be due, in part, to their transformation to CSF-dependent macrophages (48). Examination of LIF's actions on human leukemia cell lines has shown variable results. Wang and colleagues (50) demonstrated that, in two of three acute myelogenous leukemia lines, LIF increased the doubling time of the clonogenic population, possibly by prolonging the stem cell cycle. Additionally, LIF, in combination with G-CSF, GM-CSF, or IL-6, suppresses clonogenicity of U937 and HL-60 myeloid leukemia cells (51, 52). In contrast, LIF also stimulates proliferation of the murine myeloid IL-3-dependent DA-1A leukemia cell line (16, 17). In regard to the effects of LIF on normal progenitors, Metcalf and colleagues (48) found that LIF has no observable colony-stimulating activity in the murine system. Prior exposure of normal progenitor cells to LIF did, however, reduce their survival and/or their subsequent ability to proliferate after stimulation by CSFs (48). Even so, it should be noted that LIF is identical by cloning to a factor originally called HILDA, and activities attributed to HILDA include erythroid burst-promoting activity on human hemopoietic progenitors and the ability to be a chemoattractant and activator of mouse and human eosinophils (16, 17). Because LIF receptors have not been found on eosinophils (10), it has been suggested that these activities could be the result of contamination

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

212

KURZROCK ETAL.

with other molecules. On the other hand, certain investigators have also discerned LIF effects on hematopoietic progenitors. In particular, Leary and colleagues (53) felt that this cytokine was as effective as IL-6 or G-CSF in the enhancement of IL-3-dependent colony formation of very primitive human blast colony-forming cells. LIF alone, however, had no effect on CD34-positive human bone marrow cells. Additionally, Verfaillie and McGlave (54) noted that in serum-containing media (but not in serum-free cultures), LIF stimulated the growth of CFUMIX and CFU-EO in a dose-dependent fashion, and promoted increased CFU-MIX and BFU-E colony size. Depletion of accessory cells had no deleterious effect on these activities. A similar increase was induced by the addition of LIF to CD34+ cells stimulated with IL-3 combined with IL-6. Finally, Metcalf and colleagues (58) have shown that while LIF alone has no effect on murine megakaryocyte colonies, the megakaryocyte colony formation stimulated by IL-3 (multipotential CSF) could be enhanced by the addition of LIF. Recently, we have found that LIF mRNA is constitutively expressed by long-term cultures of bone marrow adherent layers, a population of cells believed to reflect the stroma of the bone marrow microenvironment (26). A role for LIF in modulating hematopoietic processes may therefore be implied, as our stromal cell cultures constitutively express only LIF, TGF/3, and M-CSF, whereas other hemopoietic regulatory cytokine transcripts [IL-la, IL-10, IL-6, GM-CSF, G-CSF, and tumor necrosis factor-a (TNFa)] are only produced after exposure to specific induction stimuli (55). Further, stromal cultures derived from some patients with chronic myelogenous leukemia produce increased amounts of LIF mRNA (26). Increased LIF expression correlates with advanced disease in these patients and is accompanied by the appearance of uninduced expression of IL-1/3 in both the primary malignant cells and in the stromal cells (56). The latter event may serve to drive the disease in an autocrine or paracrine fashion. Since our data also indicate that stromal cell LIF expression can be significantly augmented by IL-lft IL-la, TNF-a, and TGF-0 (26), it may be that enhanced LIF expression in advanced chronic myelogenous leukemia is the corollary of increased IL-1/3 production. B. In vivo Mice engrafted with cells producing high levels of LIF develop a fatal syndrome that includes a neutrophilic leukocytosis and splenic enlargement with excessive hematopoietic cells in the spleen (6, 57). Marrow cellularity is reduced with selective survival of cells derived from the granulocytic lineage. Some of the bone marrow changes are probably caused by myelosclerosis, as in-

Vol. 12, No. 3

creased bone marrow osteoblasts is a cardinal feature of the syndrome (57). When purified recombinant murine LIF is injected into mice (3), many of the characteristics of the engraftment model fail to be reproduced. At high doses (2 ng, three times daily) LIF administration is followed by toxic effects such as behavior changes, loss of body fat, and thymus atrophy. Dose-related rises are observed in blood platelets, erythrocytes, sedimentation rate, and serum calcium-albumin ratios. The latter changes are seen at dose schedules with no toxic effects. LIF also induces a decrease in bone marrow lymphocytes and elevation of erythroid populations. Additionally, a rise in megakaryocytes in the bone marrow and in the spleen are noted. IV. Bone Remodeling Several cytokines have been implicated in bone remodeling. The term "osteoclast activating factor" was originally applied to the bone-resorbing activity accompanying certain hematological malignancies. Osteoclast activating factor activity is now known to be mediated by a family of cytokines that includes IL-1 and TNFa and TNFjfl. Recently, LIF has also been added as a member of this class of substances (12). In vitro, both catabolic and anabolic effects on bone are displayed by LIF (12, 33, 57, 59-62). Moreover, mice engrafted with cells producing high levels of LIF exhibit pronounced bone resorption, calcium deposition in myocardium and skeletal muscle, and the accumulation of osteoblasts in the bone marrow accompanied by excess new bone formation (6, 57). The catabolic effects of LIF are prostaglandin-dependent; the anabolic effects are not. These activities may therefore be mediated through distinct pathways. Alternatively, LIF stimulation of osteoclasts and bone resorption may require the presence of osteoblasts as has been demonstrated for IL-1 and TNF (63, 64). Support for the latter premise is also derived from the observation that osteoblasts, but not multinucleated osteoclasts, display LIF receptors (33). Specific LIF actions on osteoblasts have been documented. In UMR106 cells (rat osteogenic sarcoma cells with osteoblast-like features), LIF increases plasminogen activator inhibitor (33, 59). The subsequent reduction in plasmin generation could be expected to reduce neutral protease activity and contribute to a net positive effect on bone formation. This type of effect is similar to that displayed by TGF/3. In MC3T3E1 cells (murine osteoblast-like line), LIF increases the level of osteopontin (a molecule that plays a role in cell attachment and proliferation), inhibits proliferation and DNA synthesis, and suppresses alkaline phosphatase activity (a differentiation marker) (60). Yet, in a different cell line (RCT-1, a preosteoblastic rat calvaria line), LIF potentiates the

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

LEUKEMIA INHIBITORY FACTOR

August, 1991

increase in alkaline phosphatase produced by retinoic acid (61). The reason for the discrepancy in LIF action on these two cell lines is not immediately apparent, but it should be remembered that the lines are derived from different species and perhaps represent distinct stages of osteoblastic maturation. Reduction of the level of type I procollagen mRNA occurs in both MC3T3E1 and RCT1 cells exposed to LIF. Evidence for the bone resorption activity of LIF has been accrued by assessment of in vitro release of calcium from prelabeled mouse fetal calvaria (12, 62). Bone resorption is associated with an increase in the number of osteoclasts, and the osteolytic effect is inhibited by indomethacin suggesting that it is mediated via local production of prostaglandin. As mentioned earlier, LIF shares many of its prostaglandin-dependent, bone resorption activities with other cytokines such as IL-1 (a and 0), TNF {a and 0), and TGF0. Integration of in vitro observations have indicated that an osteoblast-like cell is probably the primary target of LIF action (59), and bone resorption could therefore be associated with LIF stimulation of osteoblast prostaglandin production. It is also known that LIF transcripts and protein are produced by primary osteoblastic populations and by osteoblast-like cell lines. Further, in these cells, LIF levels can be increased in response to retinoic acid, TNFa, and, in some experiments, TGF0 (33). Since TNFa and TGF0 both affect bone metabolism, these observations identify a potentially interesting cytokine interaction reminiscent of that described above for the embryo and for the bone marrow microenvironment (25, 26). Overall, several lines of evidence converge to implicate LIF as an important paracrine or autocrine modulator of bone remodeling.

V. Hepatic Functions A combination of IL-6, IL-1 (or TNF), and glucocorticoids is minimally required to stimulate the synthesis of most major acute phase plasma proteins in the liver. Cultured keratinocytes and their malignant counterparts, squamous carcinoma cell lines, release additional hepatocyte stimulating factors (HSF) (8, 28). HSFIII is now known to be identical to LIF (8). Interestingly, LIF induces the same set of acute phase plasma proteins in the liver as IL-6. IL-6 is structurally distinct from LIF, and both are bound to a unique receptor. The mechanism for the similarity in their modulatory activity on hepatic cells is not known but could involve common regulatory sequences or receptor subunit similarities. However, since the overall hepatic cell response to LIF and IL-6 also shows some differences, cytokine-specific elements in the regulatory mechanism must exist.

213 VI. Lipid Metabolism

Mammals suffering from chronic infection or malignancy manifest several common clinical and biochemical abnormalities. These abnormalities include a catabolic state with weight loss accompanied by hypertriglyceridemia. The hypertriglyceridemia is associated with elevation of very low-density lipoproteins. The high levels of very low-density lipoproteins result from a clearing defect caused by suppression of the key enzyme of triglyceride metabolism, lipoprotein lipase (65). Inhibition of lipoprotein lipase activity may reduce the intake of fatty acids by adipocytes, resulting in lipid catabolism in adipose tissue and loss of fat. These derangements can be mediated by TNF, IL-1, or interferon-7 (66-68). Recently, investigators have demonstrated that LIF can also inhibit lipoprotein lipase (7). It is, therefore, possible that the cachexia exhibited by mice with high circulating levels of LIF may be explained by the ability of this molecule to suppress adipogenic processes (6).

VII. Nervous System Development In the nervous system, phenotypic decisions can be controlled by factors in the local environment termed neuronal differentiation factors. Rat heart cell conditioned medium produces a protein, cholinergic neuronal differentiation factor, which acts on postmitotic rat sympathetic neurons to specifically induce the expression of acetylcholine synthesis and cholinergic function while suppressing catecholamine synthesis and noradrenergic function (9). It is designated a differentiation factor because it controls phenotypic choices in these neurons without affecting their survival or growth. It has now been established that cholinergic neuronal differentiation factor is identical to LIF (9). Recently, Murphy and colleagues (69) have also shown that LIF stimulates the generation of sensory neurons in cultures of mouse neural crest. They also demonstrate that LIF supports the generation and/or maturation of sensory neurons in cultures of cells from embryonic dorsal root ganglia and, in cultures of postnatal dorsal root ganglia, which contain mature sensory neurons, LIF acts directly to promote neuronal survival. These results suggest that LIF may be a critical part of the process that regulates the development of peripheral neurons from their precursors in the embryonic neural crest. LIF also affects the regulation of the neuropeptide substance P (SP) in sympathetic neurons. In particular, it increases SP in both pure neuronal cell cultures and in cultures containing a mixture of neuronal and nonneuronal cells (70). Interestingly, IL-10 can also increase neuropeptide SP expression; however, unlike LIF, IL-10 has this effect only in dissociated cultures of ganglion neuronal and nonneuronal cells and not in pure neuronal

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

KURZROCK ETAL.

214

cultures. The cell type necessary for the IL-1/3 action is probably the ganglion Schwann cell. Importantly, there is significant additional evidence for the importance of hemopoietic cytokines in neuronal development. IL-1 is found in the brain and receptors for IL-1, -2, and -4 are found on neural cell lines (71-74; reviewed in Ref. 9). IL-1, IL-2, IL-6, and 7-interferon all can influence nervous system cell growth or differentiation (75-78). The opposite is also true: nerve growth factor can promote hemopoietic cell differentiation and growth (79, 80). In fact, there are many parallels in the questions of lineage decisions in the neural crest and hematopoietic systems (81). A. Receptors Cells showing LIF receptors include: rat osteoblasts (33); murine myeloid leukemia Ml cell lines (10, 11); normal murine monocyte macrophages (10); murine embryonal stem cells (ES and EC lines) (4); immature and mature murine megakaryocytes (58); and fetal and adult hepatocytes (82). Receptors are not found on murine neutrophils, mast cells, eosinophils, erythrocytes, or lymphocytes (10). Multinucleated osteoclasts also do not express receptors (33), even though it has been suggested that these cells are derived from the monocyte/macrophage lineage (83). Several distinct LIF-responsive cells have similar numbers of receptors. UMR106-06 rat (osteoblast-like) osteogenic sarcoma cells and RCT-1 preosteoblastic rat calvaria cells each have 300 LIF receptors per cell; their apparent dissociation constant (KD) is 60 and 20 pM, respectively (33, 61). Ml myeloid leukemia cells and ES embryonal cells display 100-500 high-affinity receptors per cell (apparent KD, 100-200 pM) (4, 5,10; reviewed in Ref. 34), and 20% receptor occupancy is needed for biological effects on these cell lines. Thus, the disparate effects of LIF on Ml cells (induction of differentiation) and ES cells (inhibition of differentiation) involve an unclarified molecular basis. A recent study by Hilton et al (82) showed that both fetal and adult parenchymal hepatocytes display a higher number of receptors than monocytes. Specifically, the number of receptors per positive cell ranged from 150 for bone marrow monocytes to 2,000 for adult hepatocytes. In each case, however, binding was of high affinity, with an apparent KD of 34100 pM. VIII. Summary Increasingly it seems that many cytokines are pleiotropic, and individual molecules may have critical roles in several different organ systems. LIF exemplifies this phenomenon: it influences embryogenesis, bone and lipid metabolism, and hematopoietic and nervous system

Vol. 12, No. 3

function. Many of LIF's effects are reminiscent of those of IL-1, TNF, and TGF-,9. Further, even within a single system, LIF can display totally different effects, i.e. induction of differentiation of one leukemic cell line vs. stimulation of proliferation of another. The corollary to these observations is that there appears to be many parallels in developmental systems. For instance, in the case of neuronal "lineage commitment," the events that relate to migration of neural crest cells along various pathways and their ultimate arrest in different locales demonstrate sufficient analogies to hematopoietic lineage commitment phenomena that, in a provocative review, Anderson coined the term "neuropoiesis" (81). This type of analogy becomes even more intriguing when one realizes that some of the same molecules are regulating neuronal and hematopoietic "lineage" proliferation and differentiation. In this respect, several interleukins in addition to LIF are important in neuronal development, and nerve growth factor turns out to also be a hematopoietic regulatory molecule. Similar parallels are enacted in other organ systems as well. The mediation of identical effects by distinct cytokines bound to unique receptors could conceivably be explained by receptor transmodulation or by overlapping signaling sequences. It is nevertheless also unclear how a single cytokine attached to a single receptor can accomplish varied and opposing effects, although divergent intracellular signaling mechanisms could account for some of these phenomena. Yet another enigma relates to how cells from one system can be properly influenced by a pleiotropic molecule such as LIF without significant "cross-effects" on other potentially responsive systems. Cytokine production that is restricted to certain developmental stages, or very localized distribution and spheres of influence within a microenvironment, could be explanatory. The findings of Rathjan and colleagues (18), i.e. that LIF exists as both a diffusible molecule and as a molecule incorporated into the extracellular matrix, is of special interest in relation to the above questions. Indeed, the distinctions between the roles of diffusible and immobilized signaling molecules could be crucial to the multiplicity of LIF's actions (18). Diffusible regulatory factors allow communication between spatially separated cells. Cellular responsiveness to such factors is dictated by the presence of appropriate receptors and postreceptor machinery. In contrast, immobilization of a factor in the extracellular matrix confines its signal to restricted sites. Furthermore, in embryonic cells, differential expression of the two LIF transcripts appears to be developmentally regulated and to be modulated by other cytokines such as IL-la, TGF/?, and bFGF, a phenomenon that may be responsible for fine-tuning control of LIF actions in different physiological situations. It is possible that similar control over localized

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

LEUKEMIA INHIBITORY FACTOR

August, 1991

production of different forms of LIF or analogous stem cell regulators by other specialized microenvironments such as the bone marrow also occurs. Such a concept is strengthened by findings that cultured bone marrow stromal cells produce LIF and that production in this system is also increased by cytokines that include IL-1 and TGF0 as well as TNFa (26). Similarly, TNF« and TGF/? can also increase LIF levels in osteoblasts (33). Overall, the concept which emerges with LIF as a paradigm is that, in many cases, the body does not produce a cytokine for an isolated purpose, but rather exploits individual molecules to perform a multiplicity of distinct functions.

References 1. Gearing DP, Gough NM, King JA, Hilton DJ, Nicola NA, Simpson RJ, Nice EC, Kelso A, Metcalf D 1987 Molecular cloning and expression of cDNA encoding a murine myeloid leukaemia inhibitory factor (LIF). EMBO J 6:3995 2. Moreau J-F, Donaldson DD, Bennett F, Witek-Giannotti J, Clark SC, Wong GG 1988 Leukaemia inhibitory factor is identical to the myeloid growth factor human interleukin for DA cells. Nature 336:690 3. Metcalf D, Nicola NA, Gearing DP 1990 Effects of injected leukemia inhibitory factor on hematopoietic and other tissues in mice. Blood 76:50 4. Williams RL, Hilton DJ, Pease S, Willson TA, Stewart CL, Gearing DP, Wagner EF, Metcalf D, Nicola NA, Gough NM 1988 Myeloid leukaemia inhibitory factor maintains the developmental potential of embryonic stem cells. Nature 336:684 5. Smith AG, Heath JK, Donaldson DD, Wong GG, Moreau J, Stahl M, Rogers D 1988 Inhibition of pluripotential embryonic stem cell differentiation by purified polypeptides. Nature 336:688 6. Metcalf D, Gearing DP 1989 Fatal syndrome in mice engrafted with cells producing high levels of leukemia inhibitory factor. Proc Natl Acad Sci USA 86:5948 7. Mori M, Yamaguchi K, Abe K 1989 Purification of a lipoprotein lipase-inhibiting protein produced by a melanoma cell line associated with cancer cachexia. Biochem Biophys Res Commun 160:1085 8. Baumann H, Wong GG 1989 Hepatocyte-stimulating factor-Ill shares structural and functional identity with leukemia-inhibitory factor. J Immunol 143:1163 9. Yamamori T, Fukada K, Aebersold R, Korsching S, Fann M-J, Patterson PH 1989 The cholinergic neuronal differentiation factor from heart cells is identical to leukemia inhibitory factor. Science 246:1412 10. Hilton DJ, Nicola NA, Metcalf D 1988 Specific binding of murine leukemia inhibitory factor to normal and leukemic monocytic cells. Proc Natl Acad Sci USA 85:5971 11. Yamamoto-Yamaguchi Y, Tomida M, Hozumi M 1986 Specific binding of a factor inducing differentiation to mouse myeloid leukemic Ml cells. Exp Cell Res 164:97 12. Abe E, Tanaka H, Ishimi Y, Miyaura C, Hayashi T, Nagasawa H, Tomida M, Yamaguchi Y, Hozumi M, Suda T 1986 Differentiation-inducing factor purified from conditioned medium of mitogen-treated spleen cell cultures stimulates bone resorption. Proc Natl Acad Sci USA 83:5958 13. Lowe DG, Nunes W, Bombara M, McCabe S, Ranges GE, Henzel W, Tomida M, Yamamoto-Yamaguchi Y, Hozumi M, Goeddel DV 1989 Genomics of cloning and heterologous expression of human differentiation-stimulating factor (leukemic inhibitory factor, human interleukin DA). DNA 8:351 14. Smith AG, Hooper ML 1987 Buffalo rat liver cells produce a

215

diffusible activity which inhibits the differentiation of murine embryonal carcinoma and embryonic stem cells. Dev Biol 121:1 15. Koopman P, Cotton RGH 1984 A factor produced by feeder cells which inhibits embryonal carcinoma cell differentiation: characterization and partial purification. Exp Cell Res 154:111 16. Moreau J-F, Bonneville M, Godard A, Gascan H, Gruart V, Moore MA, Soulillou JP 1987 Characterization of a factor produced by human T cell clones exhibiting eosinophil-activating and burstpromoting activities. J Immunol 138:3844 17. Godard A, Gascan H, Naulet J, Peyrat M-A, Jacques Y, Soulillou J-P, Moreau J-F 1988 Biochemical characterization and purification of HILDA, a human lymphokine active on eosinophils and bone marrow cells. Blood 71:1618 18. Rathjen PD, Toth S, Willis A, Heath JK, Smith AG 1990 Differentiation inhibiting activity is produced in matrix-associated and diffusible forms that are generated by alternate promoter usage. Cell 62:1105 19. Simpson RJ, Hilton DJ, Nice EC, Rubira MR, Metcalf D, Gearing DP, Gough NM, Nicola NA 1988 Structural characterization of a murine myeloid leukaemia inhibitory factor. Eur J Biochem 175:541 20. Hilton DJ, Nicola NA, Metcalf D 1988 Purification of a murine leukemia inhibitory factor from Krebs ascites cells. Anal Biochem 173:359 21. Tomida M, Yamamoto-Yamaguchi Y, Hozumi M, Okabe T, Takaku F 1986 Induction by recombinant human granulocyte colonystimulating factor of differentiation of mouse myeloid leukemic Ml cells. FEBS Lett 207:271 22. Hozumi M, Takenaga K, Tomida M, Okabe J 1977 Role of leucocytes in ascites in the production of factor(s) stimulating differentiation of mouse myeloid leukemia cells. Jpn J Cancer Res 68:643 23. Abe E, Ishimi Y, Takahashi N, Akatsu T, Ozawa H, Yamana H, Yoshiki S, Suda T 1988 A differentiation-inducing factor produced by the osteoblastic cell line MC3T3E1 stimulates bone resorption by promoting osteoclast formation. J Bone Mineral Res 3:635 24. Conquet F, Brulet P 1990 Developmental expression of myeloid leukemia inhibitory factor gene in preimplantation blastocysts and in extraembryonic tissue of mouse embryos. Mol Cell Biol 10:3801 25. Rathjen PD, Nichols J, Toth S, Edward DR, Heath JK, Smith AG 1990 Developmentally programmed induction of differentiation inhibiting activity and the control of stem cell populations. Genes Dev 4:2308 26. Wetzler M, Talpaz M, Lowe DG, Baiocchi G, Gutterman JU, Kurzrock R 1991 Constitutive expression of leukemia inhibitory factor RNA by human bone marrow stromal cells and modulation by IL-1, TNF-a and TGF-/3. Exp Hematol 19:347 27. Le PT, Lazorick S, Whichard LP, Yang Y-C, Clark SC, Haynes BF, Singer KH 1990 Human thymic epithelial cells produce IL6, granulocyte-monocyte-CSF, and leukemia inhibitory factor. J Immunol 145:3310 28. Baumann H, Jahreis GP, Sauder DN, Koj A 1984 Human keratinocytes and monocytes release factors which regulate the synthesis of major acute phase plasma proteins in hepatic cells from man, rat, and mouse. J Biol Chem 259:7331 29. Gascan H, Anegon I, Praloran V, Naulet J, Godard A, Soulillou JP, Jacques Y 1990 Constitutive production of human interleukin for DA cells/leukemia inhibitory factor by human tumor cell lines derived from various tissues. J Immunol 144:2592 30. Abe T, Murakami M, Sato T, Kajiki M, Ohno M, Kodaira R 1989 Macrophage differentiation inducing factor from human monocytic cells is equivalent to murine leukemia inhibitory factor. J Biol Chem 264:8941 31. Anegon I, Moreau JF, Godard A, Jacques Y, Peyrat MA, Hallet MM, Wong G, Soulillou JP 1990 Production on human interleukin for DA cells (HILDA)/leukemia inhibitory factor (LIF) by activated monocytes. Cell Immunol 130:50 32. Hozumi M, Umezawa T, Takenaga K, Ohno T, Shikita M, Yamane

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

216

33.

34. 35.

36.

37. 38. 39. 40. 41.

42. 43. 44. 45. 46.

47. 48. 49. 50. 51. 52. 53.

54.

KURZROCK ETAL.

I 1979 Characterization of factors stimulating differentiation of mouse myeloid leukemia cells from a Yoshida sarcoma cell line cultured in serum-free medium. Cancer Res 39:5127 Allan EH, Hilton DJ, Brown MA, Evely RS, Yumita S, Metcalf D, Gough NM, Ng KW, Nicola NA, Martin TJ 1990 Osteoblasts display receptors for and responses to leukemia-inhibitory factor. J Cell Physiol 145:110 Gough NM, Williams RL 1989 The pleiotropic actions of leukemia inhibitory factor. Cancer Cells 1:77 Tomida M, Yamamoto-Yamaguchi Y, Hozumi M 1984 Purification of a factor inducing differentiation of mouse myeloid leukaemic Ml cells from conditioned medium of mouse fibroblast L929 cells. J Biol CHem 259:10978 Gough NM, Gearing DP, King JA, Willson TA, Hilton DJ, Nicola NA, Metcalf D 1988 Molecular cloning and expression of the human homologue of the murine gene encoding myeloid leukaemia-inhibitory factor. Proc Natl Acad Sci USA 85:2623 Gearing DP, King JA, Gough NM 1988 Complete sequence of murine myeloid leukaemia inhibitory factor (LIF). Nucleic Acids Res 16:9857 Stahl J, Gearing DP, Willson TA, Brown MA, King JA, Gough NM 1990 Structural organization of the genes for murine and human leukemia inhibitory factor. J Biol Chem 265:8833 Sutherland GR, Baker E, Hyland VJ, Callen DF, Stahl J, Gough NM 1989 The gene for human leukemia inhibitory factor (LFI) maps to 22ql2. Leukemia 3:9 Kola I, Davey A, Gough NM 1990 Localization of the murine leukaemia inhibitory factor gene near the centromere on chromosome 11. Growth Factors 2:235 Budarf M, Emanuel BS, Mohandas T, Goeddel DV, Lowe DG 1989 Human differentiation-stimulating factor (leukemia inhibitory factor, human interleukin DA) gene maps distal to the Ewing sarcoma breakpoint on 22q. Cytogenet Cell Genet 52:19 Murray R, Lee F, Chiu C-P 1990 The genes for leukemia inhibitory factor and interleukin-6 are expressed in mouse blastocysts prior to the onset of hemopoiesis. Mol Cell Biol 10:4953 Evans MJ, Kaufman MH 1981 Establishment in culture of pluripotential cells from mouse embryos. Nature 292:154 Pease S, Williams RL 1990 Formation of germ-line chimeras from embryonic stem cells maintained with recombinant leukemia inhibitory factor. Exp Cell Res 190:209 Burgess AW 1989 Life after LIF? BioEssays 10:166 Hilton DN, Nicola NA, Gough NM, Metcalf D 1988 Resolution and purification of three distinct factors produced by Krebs ascites cells which have differentiation-inducing activity on murine myeloid leukaemic cells lines. J Biol Chem 263:9238 Shabo Y, Lotem J, Rubinstein M, Revel M, Clark SC, Wolf SF, Kamen R, Sachs L 1988 The myeloid blood cell differentiationinducing protein MGI-2A is interleukin-6. Blood 72:2070 Metcalf D, Hilton DJ, Nicola NA 1988 Clonal analysis of the actions of the murine leukaemia inhibitory factor on leukaemic and normal murine haemopoietic cells. Leukemia 2:216 Metcalf D 1989 Suppression of myeloid leukemic cells by normal hemopoietic regulators. Proceedings of the Mochida Memorial Symposium. Recent Prog Cytokine Res 10:11 Wang C, Lishner M, Minden MD, McCulloch EA 1990 The effects of leukemia inhibitory factor (LIF) on the blast stem cells of acute myeloblastic leukemia. Leukemia 4:548 Maekawa T, Metcalf D 1989 Clonal suppression of HL60 and U937 cells by recombinant human leukemia inhibitory factor in combination with GM-CSF or G-CSF. Leukemia 3:270 Maekawa T, Metcalf D, Gearing DP 1990 Enhanced suppression of human myeloid leukemic cell lines by combinations of IL-6, LIF, GM-CSF and G-CSF. Int J Cancer 45:353 Leary AG, Wong GG, Clark SC, Smith AG, Ogawa M 1990 Leukemia inhibitory factor/differentiation-inhibiting activity/human interleukin for DA cells augments proliferation of human hematopoietic stem cells. Blood 75:1960 Verfaillie C, McGlave P 1991 Leukemia inhibitory factor/human interleukin for DA cells: a growth factor that stimulates the in

55.

56.

57. 58. 59.

60.

61.

62. 63. 64. 65. 66. 67. 68.

69. 70. 71. 72. 73. 74.

75.

Vol. 12, No. 3

vitro development of multipotential human hematopoietic progenitors. Blood 77:263 Wetzler M, Kurzrock R, Taylor K, Spitzer G, Kantarjian H, Ku S, Gutterman JU, Talpaz M 1990 Constitutive and induced expression of growth factors in normal and chronic phase CML Ph1 bone marrow stroma. Cancer Res 50:5801 Wetzler M, Kurzrock R, Lowe D, Kantarjian H, Gutterman JU, Talpaz M 1990 Alteration in bone marrow stromal growth factor expression—a possible mechanism of disease progression in chronic myelogenous leukemia. Proc Am Assoc Cancer Res 31:179A (Abstract) Metcalf D, Gearing DP 1989 A myelosclerotic syndrome in mice engrafted with cells producing high levels of leukemia inhibitory factor (LIF). Leukemia 3:847 Metcalf K, Hilton D, Nicola NA 1991 Leukemia inhibitory factor can potentiate murine megakaryocyte production in vitro. Blood 77:2150 Allan EH, Gough NM, Gelehrter TD, Zeheb R, Martin TJ 1989 Transforming growth factor ft and leukemia inhibitory factor increase mRNA and protein for plasminogen activator inhibitor1 in osteoblasts. J Bone Mineral Res [Suppl 1] 4:5324 (Abstract) Noda M, Vogel RL, Hasson DM, Rodan GA 1990 Leukemia inhibitory factor suppresses proliferation, alkaline phosphatase activity, and type I collagen messenger ribonucleic acid level and enhances osteopontin mRNA level in murine osteoblast-like (MC3T3E1) cells. Endocrinology 127:185 Rodan SB, Wesolowski G, Hilton DJ, Nicola NA, Rodan GA 1990 Leukemia inhibitory factor binds with high affinity to preosteoblastic RCT-1 cells and potentiates the retinoic acid induction of alkaline phosphatase. Endocrinology 127:1602 Reid IR, Lowe C, Cornish J, Skinner SJM, Hilton DJ, Willson TA, Gearing DP, Martin TJ 1990 Leukemia inhibitory factor: a novel bone-active cytokine. Endocrinology 126:1416 Thomson BM, Saklatavala J, Chambers TJ 1986 Osteoblasts mediate interleukin 1 stimulation of bone resorption by rat osteoclasts. J Exp Med 164:104 Thomson BM, Mundy GR, Chambers TJ 1987 Tumor necrosis factors a and /3 induce osteoblastic cells to stimulate osteoclastic bone resorption. J Immunol 138:775 Cryer A 1981 Tissue lipoprotein lipase activity and its action in lipoprotein metabolism. J Biochem 13:525 Beutler B, Cerami A 1987 Cachectin: more than a tumor necrosis factor. N Engl J Med 316:379 Beutler BA, Cerami A 1985 Recombinant interleukin 1 suppresses lipoprotein lipase activity in 3T3-L1 cells. J Immunol 135:3969 Kurzrock R, Rohde MF, Quesada JR, Gianturco SH, Bradley WA, Sherwin SA, Gutterman JU 1986 Recombinant interferon gamma induces hypertriglyceridemia and inhibits post heparin lipase activity in cancer patients. J Exp Med 164:1093 Murphy M, Reid K, Hilton DJ, Bartlett PF 1991 Generation of sensory neurons is stimulated by leukemia inhibitory factor. Proc Natl Acad Sci USA 88:3498 Freidin M, Kesler JA 1991 Cytokine regulation of substance P expression in sympathetic neurons. Proc Natl Acad Sci USA 88:3200 Breder CD, Dinarello CA, Saper CB 1988 Interleukin-1 immunorective innervation of the human hypothalamus. Science 240:321 Farrar WL, Kilian PL, Ruff MR, Hill JM, Pert CB 1987 Visualization and characterization of interleukin 1 receptors in brain. J Immunol 139:459 Smith LR, Brown SL, Blalock JE 1989 Interleukin-2 induction of ACTH secretion: presence of an interleukin-2 receptor a chain like molecule on pituitary cells. J Neuroimmunol 21:249 Lowenthal JW, Castle BE, Christiansen J, Schreurs J, Rennick D, Arai N, Hoy P, Takebe Y, Howard M 1999 Expression of high affinity receptors for murine interleukin 4 (BSF-1) on hemopoietic and nonhemopoietic cells. J Immunol 140:456 Lindholm D, Heumann R, Meyer M, Thoenen H 1987 Interleukin1 regulates synthesis of nerve growth factor in non-neuronal cells of rat sciatic nerve. Nature 330:658

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

August, 1991

LEUKEMIA INHIBITORY FACTOR

76. Giulaian D, Young DG, Woodward J, Brown DC, Lachman LB 1988 Interleukin-1 is an astroglial growth factor in the developing brain. J Neurosci 8:709 77. Benveniste EN, Merrill JE 1986 Stimulation of oligodendroglial proliferation and maturation by interleukin-2. Nature 321:610 78. Satoh T, Nakamura S, Taga T, Matsuda T, Hirano T, Kishimoto T, Kaziro Y 1988 Induction of neuronal differentiation in PC12 cells by B-cell stimulatory factor 2/interleukin 6. Mol Cell Biol 8:3546 79. Matsuda H, Coughlin MD, Bienenstock J, Denburg JA 1988 Nerve growth factor promotes human hemopoietic colony growth and differentiation. Proc Natl Acad Sci USA 85:6508-6512

217

80. Aloe L, Levi-Montacini R 1977 Mast cells increase in tissues of neonatal rats injected with the nerve growth factor. Brain Res 133:358 81. Anderson DJ 1989 The neural crest cell lineage problem: neuropoiesis? Neuron 3:1 82. Hilton DJ, Nicola NA, Metcalf D 1991 Distribution and comparison of receptors for leukemia inhibitory factor on murine hemopoietic and hepatic cells. J Cell Physiol 146:207 83. Marks SC 1983 The origin of osteoclasts: evidence, clinical implications, and investigative challenges of an extraskeletal source. J Pathol 12:226

American Institute for Cancer Research Conference on "Exercise, Calories, Fat & Cancer" September 4-5, 1991 Ritz Carlton Hotel, Pentagon City, Virginia (Metropolitan Washington, D. C.) Current research will be presented by leading scientists on the role of exercise, calories, and fat in the prevention and treatment of cancer. Platform and poster presentations made by AICR grant recipients, invited participants, and guests will provide a forum to increase understanding of the influences, mechanisms and practical applications of nutritional research on the cancer process, [9 CME hours for RDs and DTs by the ADA and 8.5 hours of AMA CME Category I credit by Georgetown University School of Medicine have been approved.] For information contact Rita Taliferro, Conference Management Division, Associate Consultants, Inc., 1726 M Street, NW, Suite 400, Washington, D. C. 20036. Phone: (202)737-8062.

7th European Workshop on Molecular and Cellular Endocrinology of the Testis Castle Elmau, Bavarian Alps, Germany, May 5-10, 1992 Plenary Sessions: Endocrine and paracrine regulation of testicular function/Inhibin, activin and growth factors/ Gonadotropin receptors/Androgen receptor and testosterone metabolism/Spatial arrangement of germ cells/Genetic control of spermatogenesis/Sperm maturation in the epididymis Invited Speakers: E. Y. Adashi (USA), A. O. Brinkman (NL), B. A. Cooke (UK), T. G. Cooper (D), B. Jegou (F), L. Martini (I), J. Mather (USA), M. van Noort (NL), W. Schulze (D), P. Vogt (D), D. Wolgemuth (USA), P. Wong (Hongkong) Miniposters: Ample time is allocated for the discussion of submitted miniposters and there will be extensive Meet-the-expert-sessions. Deadline for submission of miniposters and registration: January 4, 1992. Instructions for miniposters and registration forms are available from: Prof. Dr. E, Nieschlag, Institute of Reproductive Medicine of the University, Steinfurter Str. 107, D-4400 Minister, Germany, Phone: 49-251-836097, Fax: 49-251-836093.

The Endocrine Society. Downloaded from press.endocrine.org by [${individualUser.displayName}] on 14 November 2015. at 13:45 For personal use only. No other uses without permission. . All rights reserved.

LIF: not just a leukemia inhibitory factor.

Increasingly it seems that many cytokines are pleiotropic, and individual molecules may have critical roles in several different organ systems. LIF ex...
1MB Sizes 0 Downloads 0 Views