:0 . I ‘ , . . .: _ . ‘ , m+§luh «hm. . . . . V w. 4... ... swam? .W .Q a: $3.. fin»... ,. .V 2 s L . a 2:, . .fiw “.3va W #1.. , L.. .émfiw flag, . 3.... .. flu in”, «a.» Efi This is to certify that the dissertation entitled The Negative Effect of Corticosterone on Murine Bone Marrow B Lymphocytes: Modulation by IL—7 and Stromal Cells presented by Tonya S. Laakko has been accepted towards fulfillment of the requirements for Ph.D. Biochemistry and Molecular de ee in gr Biology ajor professor afloat Ml 07* Date 0& [137016 MS U is an Affirmative Action/Equal Opportunity Institution 0- 12771 LIBRARY Michigan State University PlACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE ”7270' JUL 2 8' 4 6/01 cJClHCIDaiODWpSS-DJS THE NEGATIVE EFFECT OF CORTICOSTERONE ON MURINE BONE MARROW B LYMPHOCYTES: MODULATION BY IL-7 AND STROMAL CELLS By Tonya S. Laakko A Dissertation Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Doctorate of Philosophy The Department of Biochemistry and Molecular Biology 2000 ABSTRACT THE NEGATIVE EFFECT OF CORTICOSTERONE ON MURINE BONE MARROW B LYMPHOCYTES: MODULATION BY IL-7 AND STROMAL CELLS By Tonya S. Laakko B cells are critical to many facets of host defense making their development and production in the marrow of vital importance. Glucocorticoids (Ge), especially those produced endogenously in trauma patients, burn victims, the malnourished, etc., where the neuroendocrine stress axis is activated, can greatly impair lymphopoiesis. This reduces the numbers of circulating B cells and compromises immune defense. Considering the importance of B cell development, surprisingly little was known about the mechanisms involved in the downregulation of lymphopoiesis by Go. The experiments performed in this thesis were designed to address the following questions: 1) How do increased concentrations of the natural glucocorticoid, corticosterone (Cs), effect the various stages of development of cells of the B lineage in murine bone marrow? 2) Can stromal cells that support hematopoiesis modulate the in vitro response of precursor B cells to corticosterone? 3) Can the cytokines interleukin-7 and/or stem cell factor, which are essential to B cell development, modulate the adverse effects of corticosterone on precursor B cells in vitro? Using a Cs implantation system, circulating Cs was elevated to concentrations analogous to that seen during normal physiological stress in the mouse model system. Flow cytometric identification of the subpopulations of developing B lymphocytes indicate that Cs had adverse effects within 12 hours especially on pre B cells. By 36 hours nearly all (70-90%) of early-pro. late-pro, pre and immature cells were lost from the bone marrow. Only mature B cells and the earliest defined progenitor, the pre-pro B cell, showed resistance to Cs. Phenotypic and DNA analysis via flow cytometry showed that Cs induced apoptosis in pro, pre and IgM+ B cells fiom murine bone marrow and decreased cell cycling in pro and pre B cells in vitro. Surprisingly, interleukin-7 which promotes lymphopoiesis, but not stem cell factor, completely inhibited Cs—induced apoptosis and cell cycle arrest in the earliest B lymphocytes, the pro B cells. IL-7 also had a modest protective effect on pre B cells, reducing apoptosis by 30%, but it did not protect them from the Cs-induced decrease in cell cycling. It provided only limited protection to immature IgM+ bearing cells. Stromal cells also significantly reduced Cs—induced apoptosis (30-50%) for all stages of B lymphocyte development in vitro although they did not restore cell cycling to normal levels. Clearly these experiments have shown that Cs can have a rapid and adverse effect on the development of cells of the B lineage in mouse BM in vivo and in vitro. Nevertheless, substantial protection afforded the precursor B cells by interleukin-7 and stromal cells in vitro suggest that with further investigations this cytokine and other factors produced by stromal cells could potentially promote immune recovery from stress and Cs induced damage. TABLE OF CONTENTS LIST OF TABLES ............................................................................ vi LIST OF FIGURES ........................................................................... vii ABBREVIATIONS ........................................................................... ix INTRODUCTION ........................................................................... 1 CHAPTER I: LITERATURE REVIEW ................................................... 4 B Cell Lymphopoiesis ............................................................... 5 Background .................................................................. 5 Model Systems ............................................................... 6 Hematopoiesis ................................................................ 8 Flow cytometry: Phenotypic Markers for Monitoring B Cell Lymphopoiesis ............................................. 8 Irnmunoglobulin Gene Rearrangement .................................. 11 B Lymphocyte Homeostasis and Selection .............................. 12 Stage Specific Gene Expression .......................................... 15 Microenvironmental Effects on B Lymphopoiesis ..................... 16 Cell-Cell and Cell-Extracellular Matrix Contact ....................... 20 Stromal Cell Derived Cytokine Support ................................. 22 Summary ..................................................................... 28 Glucocorticoids and the Immune System ........................................ 30 Background .................................................................. 3O Glucocorticoid Biochemistry .............................................. 33 The Glucocorticoid Receptor .............................................. 34 Glucocorticoid Receptor Signaling ....................................... 36 Summary ..................................................................... 4O Apoptosis .............................................................................. 41 History ........................................................................ 41 Apoptotic Morphology and Detection .................................... 42 Apoptosis Activation (Caspases and Mitochondria) ................... 44 Apoptosis Inhibition (Bel-2 Family of Proteins) ........................ 47 Bel-2 Family Members in B Cell Lymphopoiesis ...................... 50 Cytokine-Withdrawal Induced Apoptosis ................................ 51 Glucocorticoid Induced Apoptosis ....................................... 53 Summary ..................................................................... 55 CHAPTER II: THE IN V1 V0 EFFECT OF CORTICOSTERONE ON DEVELOPING B LYMPHOCYTES: FROM COMMITTED PROGENITOR TO MATURITY .................................................. 64 Abstract ................................................................................ 65 Introduction ........................................................................... 67 iv Materials and Methods ............................................................... 72 Results ................................................................................. 77 Discussion ............................................................................. 83 CHAPTER III: STROMAL CELL PROTECTION AGAINST CS-INDUCED APOPTOSIS IN BONE MARROW B LYMPHOCYTES ..................... 100 Abstract ................................................................................ 101 Introduction ........................................................................... 103 Materials and Methods ............................................................... 107 Results ................................................................................. 111 Discussion ............................................................................. 1 17 CHAPTER IV: INTERLEUKIN-7 PROTECTS EARLY STAGES IN E LYMPHOYCTEDEVELOPMENT FROM CORTICOSTERONE-INDUCED APOPTOSIS AND CELL CYCLE ARREST .................................... 132 Abstract ................................................................................ 133 Introduction ........................................................................... 135 Materials and Methods ............................................................... 138 Results ................................................................................. 141 Discussion ............................................................................. 148 REFERENCES ................................................................................ 164 Table 2.1 Table 3.1 Table 4.1 Table 4.2 LIST OF TABLES Bone marrow cellularity of Cs or sham implanted mice ............... 91 Bone marrow B cells in S/Gz/M phases of the cell cycle .............. 125 Recovery and viability of bone marrow cultured for 15 to 16 Hours ............................................................... 152 Effects of IL-7 and SCF on apoptosis in pro B cells ................... 163 vi Figure l .1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 1.5 Figure 1.6 Figure 1.7 Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5 Figure 2.6 Figure 3.1 Figure 3.2 Figure 3.3 Figure 3.4 LIST OF FIGURES Differentiation of various blood cell lineages from a pluripotent stem cell ......................................................... 57 Bone marrow B lymphocyte development scheme ..................... 58 Developing B lymphocyte association with stromal cells in the bone marrow ............................................................ 59 Corticosteroid biosynthesis ................................................ 60 The glucocorticoid receptor ............................................... 61 Apoptotic pathways ........................................................ 62 Cytokine survival via Bad inactivation .................................. 63 Plasma corticosterone concentrations .................................... 88 The effect of Cs or cholesterol implants on thymic weights .......... 90 Flow cytometric profile of developing bone marrow B lymphocytes .................................................................. 93 Cs-induced changes in bone marrow B lymphocyte populations over time ....................................................... 95 Cs-induced cell cycle changes in pro and pre B cells .................. 97 Cs effect on major hematopoietic lineages in the bone marrow ........................................................................ 99 Spontaneous and Cs-induced apoptosis in pro, pre and IgW B cells in vitro ........................................................................ 122 Stromal cell modulation of spontaneous and Cs-induced apoptosis ..................................................................... 124 Flow cytometry of S10 and/or IL—7 effects on apoptosis and cycling of Pro B cells .................................................. 127 Relative expression of IL-7 and SCF message .......................... 129 vii Figure 3.5 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Cs effect on stromal cell expression of SCF and IL-7 RNA ........... 131 IL-7 modulation of spontaneous and Cs—induced apoptosis ........... 154 IL-7 and/or Cs effect on cell cycle distribution among pro and pre B cells .................................................................... 156 Flow cytometry of IL-7 and/or Cs effect on pro B cell apoptosis and cell cycle .................................................... 158 IL-7 effect on apoptosis induced by varying concentrations of Cs among B cells ........................................................... 160 SCF effect on spontaneous and Cs-induced apoptosis ................. 162 viii BM Cs Gc Dex IL SCF BCR TCR HSA SCID RAG tdt , BSAP-S NFKB IKB LTBMC VCAM ECM SDF-l TSLP CSF VLA PI3K yc ACTH CRH HPA LDL HDL CBG GcR SHR GRE SCR GRIP HAT AIDS PS TNFR ICE LIST OF ABBREVIATIONS bone marrow corticosterone glucocorticoid dexamethasone interleukin stem cell factor immunoglobulin B cell receptor T cell receptor heat stable antigen severe combined immunodeficiency recombination activation gene terminal deoxynucleotidal transferase B-cell-specific activator protein—5 nuclear factor kappa B inhibitor of nuclear factor kappa B long term bone marrow culture vascular cellular adhesion molecule extracellular matrix stromal derived factor-1 thymic stromal lymphopoietin colony stimulated factor very late antigen phosphatidylinositol 3-kinase common gamma chain adrenal corticotrophin hormone corticotrophin releasing hormone hypothalamic-pituitary—adrenal low density lipoprotein high density lipoprotein corticosterone binding globulin glucocorticoid receptor steroid hormone receptor heat shock protein glucocorticoid response element steroid receptor coactivator glucocorticoid receptor interacting protein histone acetyltransferase aquired immunodeficiency syndrome phosphatidylserine tumor necrosis factor receptor interleukin l-B converting enzyme ix APAF DD STAT Jak Smase PI-PLC apoptotic protease activating factor death domain signal transducers and activators of transcription janus kinase sphingomyelinase phosphatidylinositol phospholipase C INTRODUCTION The successful generation of mature B cells (B cell lymphopoiesis) is critical to maintenance of the immune system. Aberrations in lymphopoiesis can result in a variety of diseases and/or failure to adequately respond to pathogens. Research in this area of immunology has been limited due to the complexity of the development process, the difficulty in identification of developing B lymphocytes and the very heterogeneous nature of the bone marrow where B cells develop in adult mammals. Therefore much of the research on B lymphocytes has been performed using immortalized cell lines, rather than natural cells. However, over the past decade methodology has been developed to successfully identify B lymphocytes from the other cell types that exist in the bone marrow (BM) (Coffinan and Weissman, 1981). In addition, distinct stages in B lymphocyte development have been defined based on the differential expression of surface proteins (Hardy et al., 1991). These stages have also been defined functionally based on Ig gene rearrangement. This thesis utilizes this sophisticated methodology to study, in depth, the effect of glucocorticoid (Go) elevation on different substages of cells of the B lineage. Chronic elevation of Ge, due to either physiological stress or pharmacological administration, can result in immunosuppression. In viva research from our lab has shown that Go elevation reduces the development of B cells in mice (Garvy et al., 1993). It is likely that induction of apoptosis in these developing cells is at least one mechanism whereby Gc elevation adversely affects lymphopoiesis. However in viva studies of apoptosis are severely limited due to the rapid phagocytosis of cells early in the apoptotic process. Evidence that Go can induce significant amounts of apoptosis in these populations have come from in vitra studies; the synthetic Gc, dexamethasone (dex), has been used for decades as a classic inducer of apoptosis in thymocytes (developing T lymphocytes). Although fewer studies have been performed using normal developing B lymphocytes, some research has shown that glucocorticoids can also induce apoptosis in these cells, mostly IgM' B cells, in vitra (Garvy et al., 1993; Griffiths et al., 1994; Merino et al., 1994). Very few in vitra studies of the effect of Gc on B or T lymphocytes have been performed using the natural Ge (corticosterone [Cs] in rodents). Natural and synthetic Gc have been reported to have different affinities for the Ge receptor and different biological efficacies. To focus on physiological rather than pharmacological effects of Go, the effect of the natural steroid hormone, Cs, rather than synthetic Gc has been investigated here. In viva studies were performed herein to determine the effect of increased concentrations of circulating Cs on each stage in B lymphocyte development to determine if losses induced by Cs correspond with populations thought to be more susceptible to apoptosis. In vitra studies were also performed to thoroughly determine the apoptotic effect of Cs on developing B lymphocytes in culture. Additionally, this thesis investigates potential modulation of Cs-induced apoptosis in developing B lymphocytes in vitra by cells and factors that would normally be found in the BM microenvironment. B cell lymphopoiesis depends on contact with, and factors secreted from, large fibroblast-like cells found in the bone marrow called stromal cells. Stromal cell lines have been utilized here to determine if, in addition to promoting B cell lymphopoiesis, these cells could modulate the B lymphocyte response to Cs. A cytokine derived fi'om stromal cells, interleukin-7 (IL-7), has been shown to support B cell lymphopoiesis, when added to B lymphocyte cultures not containing stromal cells (Namen et al., 1988). This cytokine has been shown to be specific for early stages of B and T cell lymphopoiesis and clinical interest in this cytokine for immunotherapy has recently developed (Maeurer et al., 2000; Westermann et al., 1998). Therefore we have investigated the direct effect of IL-7 addition on B lymphocyte response to Cs in vitra. The cytokine stem cell factor (SCF) promotes the commitment of early blood cell progenitors to the B lineage in combination with IL-7 (McNiece et al., 1991). SCF also amplifies the lymphopoietic effect of IL-7 on early stages during B cell development, so the affect of this cytokine on B lymphocyte induced apoptosis, alone and in combination with lL-7 has also been investigated. In Chapter 1 of this document the literature on B cell lymphopoiesis, glucocorticoids and apoptosis will be reviewed. This will be followed by three data chapters covering 1) the in viva effect of Cs elevation on developing B lymphocytes 2) stromal cell mediated modulation of the affect of Go on B lymphocytes in vitra and 3) the in vitra effect of IL-7 and/or SCF on Gc-induced apoptosis and cell cycle arrest in bone marrow B lymphocytes. CHAPTER 1: LITERATURE REVIEW B CELL LYMPHOPOIESIS Backgron B cell lymphopoiesis occurs in the bone marrow of adult marmnals and is a highly complex, multi-stage process that results in the generation of mature B cells that are involved in host defense. Membrane bound immunoglobulins (Ig) are the molecules responsible for B cell recognition of antigen and are composed of two heavy chains and two light chains. The generation of the immense diversity required for antigen recognition is via the recombination of the heavy and light chain genes (Rast and Litman, 1998). Briefly, recombination is achieved by rearrangement of several variable gene segments to the constant domain. This occurs for both the heavy and light chains. In addition to diversity derived from recombination, there are five heavy chain isotypes that result in various effector functions. The genes are mu, delta, gamma, alpha and epsilon which result in five isoforrns, IgM, IgD, IgG, IgA and IgE, respectively. On the cell surface Ig molecules are associated with Igor and IgB coreceptors that are part of the B cell receptor (BCR) complex (Hombach et al., 1990; Kashiwamura et al., 1990; Wienands, 2000). Ligation of Ig causes an intracellular phosphorylation cascade through the cytoplasmic domains of [go and IgB that results in increased proliferation and differentiation to plasma cells that produce large quantities of soluble lg (antibodies) (Justernent, 2000; Matsuo et al., 1993; Nomura et al., 1991). Antibodies mediate neutralization of toxins, initiate the complement cascade and elicit T cell responses in defense of foreign pathogens. The dysregulation of B cell lymphopoiesis can result in several diseases. Multiple disorders such as systemic lupus erythematosis, asthma and rheumatoid arthritis can result from B cells with faulty or “anti-self” Igs produced via dysregulated lymphopoiesis. This thesis is concerned with the down-regulation of lymphopoiesis, which can occur during physiological stress and can result in compromised immune responses. In fact, this lab has extensively studied on the affects of zinc deficiency (a common example of physiological stress) on lymphopoiesis (DePasquale—Jardieu and Fraker, 1984; Fraker et al., 1977; King etal., 1995; Osati-Ashtiani et al., 1998). Zinc deficiency results in massive losses of precursor (3220+IgM') B lymphocytes in the bone marrow, suggesting that downregulation of B lymphopoiesis may be the major mechanism whereby peripheral B lymphocytes are decreased, leading to a compromised immune system. During zinc deficiency and other conditions of chronic stress, a steroid hormone, glucocorticoid (Go), is believed to induce many of the physiological responses. In fact this lab has also shown that chronic elevation of Go decreased BM B cell in mice (Garvy et al., 1993). The mechanism by which lymphopoiesis is downregulated by glucocorticoids is not well defined and this is the phenomenon that this thesis will focus 011. Model Systems There are currently few treatments available for the devastating, sometimes fatal, health scenarios presented above and there are often no reliable cures. Lack of progress in this field can, at least in part, be attributed to a lack of understanding of the mechanisms whereby B lymphocytes develop into functional mature cells and how homeostasis is maintained. A multi-stage, complex maturation process in the highly heterogeneous bone marrow (BM) compartment can be concluded to be one main reason for our lack of advancement. The majority of our knowledge about B cell lymphopoiesis has been attained from cell lines, short-term whole bone marrow cultures or long-term bone marrow cultures. The murine (mouse) model has proven to be very useful in studying lymphopoiesis, and while not identical to the human system has shown fundamental similarities (Akashi et al., 2000; LeBien, 1998). The use of immortalized cell lines to study B lymphocytes has proven to be useful for determining important intracellular processes such as transcription and signaling in specific stages of B cell development. However, a limitation of cell lines is that they are not always a reliable model for determining cellular functions such as differentiation, cell cycle and cell death responses, since these are the systems often modified from normal by irnmortalization. Short-term bone marrow cultures have been valuable for studying some responses of primary developing B cells, although within a few days the majority of B cells will die due to the absence of microenvironmental support factors. The advent of a long term culture system developed by Whitlock and Witte (Whitlock et al., 1984), which will be revisited later, has resulted in a reliable in vitra system to specifically grow and regenerate B lymphocytes for long periods of time. This review addresses the murine system and the experiments presented in this thesis were performed on primary mouse cells or stromal cell lines unless otherwise specified. Hematopoiesis Figure 1.1 is a diagram of the generation of the various lineages of blood cells from a hematopoietic stem cell. In the BM of adult mammals all lineages of blood cells are generated and, for the most part, go through multiple stages of development there. T cells are a notable exception; they are generated in the BM and then migrate to the thymus for maturation. Overall, there are at least eight different blood cell lineages present in various stages of maturation in the BM. Developing and mature B lymphocytes compose approximately 30% of the total cells of the marrow. The B cells are a branch of the lymphoid lineage along with T and natural killer (NK) cells. In fact many parallels are seen between the development of B and T cells, including differentiation, proliferation and selection based on the generation of a firnctional BCR or T cell receptor (TCR), respectively. The study of T cell lymphopoiesis has been aided by the nearly homogeneous tissue that the majority of their development takes place in, the thymus. On the other hand, to study primary B lymphocytes there was a vital need for methodology to identify specific subsets of cells within the very heterogeneous population of cells found in the marrow. Flow Cytometry: Phenotypic Markers for Monitoring B Cell Lymphopoiesis F low cytometry has proven to be of great use in analysis of heterogeneous tissues, such as the BM. F luorochrome conjugated antibodies to surface proteins can be used to identify various cell types. Multi-parameter flow cytometry can allow for the analysis of several fluorochromes simultaneously, thus the expression of several cellular proteins can be analyzed simultaneously. This “multi-color” labeling has been useful in not only identifying the various developmental stages of B lymphocytes in BM, but also in identifying distinct developmental subsets within the population. For many years mature B lymphocytes were identified and purified based on surface Ig expression. Precursor cells, committed to the B lineage but lacking surface lg, could not be identified from other cell types found in the marrow. By immunizing rats with murine B cells and neoplastic pre B cell lines, a monoclonal antibody was produced which could identify lg expressing B cells and precursor B cells (Coffinan and Weissman, 1981). The antibody recognized a 220,000 MW glycoprotein (B220, CD45RA); it was determined that this antibody was specific for the B lineage. 3220 is a phosphotyrosine phosphatase that dephosphorylates and inactivates src tyrosine kinases such as 1yn (Katagiri et al., 1999; Satterthwaite and Witte, 1996). The precursor cells have been further differentiated since then based on stage-specific expression of various proteins. Several labs have developed murine B lymphocyte identification schemes independently and the incongruent nomenclature and differing phenotypic markers used in identification has added confusion to the field, although each supports the same order of development based on the progression of Ig recombination (Lu et al., 1998; Osmond et al., 1998). This lab has adopted the established phenotypic nomenclature developed by Hardy et a1 (Hardy et al., 1991). Figure 1.2 is a detailed representation of the ordered development of B lymphocytes showing surface proteins, relative size and expression of significant proteins in each defined stage in development. It incorporates information that will be presented throughout this review on B lymphopoiesis. Pro B cells can be identified by the surface expression of the molecule recognized by the S7 monoclonal antibody (leukosialin, CD43) (Hardy et al., 1989). S7 is the clone designation and it is also used as the nomenclature for the antigen it recognizes. Although it is expressed on many cell types in the marrow, it is specific for the pro B cell stage of development in 8220+ B lymphocytes of the bone marrow. S7 is a highly sialylated integral membrane protein which has been shown to play a role in adhesion (Laferte and Dennis, 1988, Walker, 1999 #195). The pro B cells can be subdivided further based on expression of heat stable antigen (HSA) (CD24) and a membrane bound aminopeptidase, BP—l (Ly-51, 6C3) (Wu et al., 1991). HSA (CD24) is a highly glycosylated protein which is involved in adhesion and in the regulation of lymphopoiesis. In transgenic mice that overexpress HSA and in HSA knockout mice, the generation of early B lymphocyte precursors is downregulated (Hough et al., 1996; Nielsen et al., 1997). Although BP-l expression is tightly regulated in B cell development a functional role has not been assigned, and lymphopoiesis is normal in BP-l deficient mice (Lin et al., 1998; Wang et al., 1998). The earliest pro B cells, pre-pro, do not express either HSA or BP-l (BZ20+S7+HSA'BP- 1'). The expression of HSA on the surface identifies the next stage of development, the early-pro B cells (3220+S7IHSA+BP-l'). Late-pro B cells are identified by the expression of BP-l (B220+S7+HSA+BP-l+) as development proceeds toward the pre B cell stage the expression of S7 is lost (8220+S7‘HSA+BP-l+). Irmnature B cells have lgM on their surface, and maturity of this lineage is marked by the coexpression of surface lgM and IgD. The orderly progression through these stages to maturity is based on the rearrangement of Ig heavy and light chain genes and the expression of various proteins that play a role in the rearrangements, differentiation, proliferation and apoptosis. This will be discussed more in-depth below and is addressed throughout the experimental sections. Immunoglobulin Gene Rearrangement In general pro B cells are defined as cells committed to the B lineage prior to expression of cytoplasmic 1; heavy chain. The identification of a variety of other distinctive features of these cells has led to a better functional understanding of their development and maturation. Using reverse transcriptase polymerase chain reaction (RT- PCR) amplification techniques, the ordered progression of heavy chain gene rearrangements was elucidated (Hardy et al., 1991). The earliest pro B cells (pre-pro as defined by Hardy) have Ig genes in a gerrnline configuration and the cells are small and noncycling. Although evidence exists which suggests that this population may not be entirely composed of cells committed to the B lineage, a substantial fi'action of these cells retain the ability to differentiate into mature B cells. The next defined stage, the early- pro B cells, are a large, proliferating cell type undergoing recombination of the p. heavy chain diversity regions (DH) to joining (JH) regions. In the late pro B cells, DJH recombination is completed and joined to one of several variable gene segments (V H). These cells are also large and proliferating. Upon completion of [.1 heavy chain rearrangement, and subsequent cytoplasmic expression, the cells are considered pre B cells. There exists a stage where the pre B cells are large and actively cycling followed by a stage of non-proliferation and smaller size (Coffrnan, 1982, Osmond, 1986 #132) (see Figure 2). At this stage the pre B cells undergo recombination of the joining region to either a kappa or lambda light chain variable region (VK or VA, respectively). Completed lgM rearrangement results in export to the cell surface, defining the cells as 11 immature B lymphocytes. Maturity is accomplished following production and surface expression of IgD. B Lymphocyte Homeostasis and Selection The complexity of this recombination process suggests a high potential for faulty rearrangements, yet its importance suggests a need for quality control. In fact studies on the population dynamics of B lymphopoiesis indicated that the production of BM B lymphocytes far exceeds the number of mature cells generated (Osmond, 1990; Osmond, 1986). Using tritiated thymidine incorporation to study in viva proliferation it was discovered that certain precursor B cells actively cycle. These were identified as being large cellsgin the late pro B cell stage and in the pre B cell stages of development. Further experiments using vincristine to block cells in metaphase were performed to determine the rate of proliferation of the various B cell subsets (Opstelten and Osmond, 1983). Together, these experiments led to the estimation that approximately fifty million B cells are generated per day in the bone marrow, but only around ten percent of those are exported fiom the marrow. The major losses appear to occur in transition from the pro B to pre B and from pre B to immature B cells (Lu and Osmond, 1997). This suggested that, indeed, many B lymphocytes were somehow being eliminated and this was consistent with the probability of the generation of faulty and anti-self Ig gene rearrangements. Therefore, it is widely accepted that B lymphocytes with nonsense and anti-self lg gene rearrangements are eliminated before reaching maturity. The mechanism by which cells identify faulty or functional heavy chains began to be elucidated by the identification of the expression of a surrogate light chain in early stages of B cell development. The surrogate light chain is composed of two non-variable molecules, termed A5 and VpreB (Kudo and Melchers, 1987; Sakaguchi and Melchers, 1986). Surrogate light chain associates with rearranged )1 heavy chains via a disulfide bond and allows for its transient surface expression (Karasuyama et al., 1990). These associated molecules along with Igor and IgB are termed the pre-B cell receptor (pre- BCR). The pre-BCR appears to be used as a checkpoint molecule for either firrther differentiation or elimination of the cell (Karasuyama et al., 1994; Rolink et al., 2000). Transgenic mice with a defective pre-BCR, due to a disrupted A5 gene, show a block in the progression of pro B cells into pre B cells (Kitarnura et al., 1992; Rolink et al., 1993). In addition, the SCID defect in mice results in the inability to rearrange Ig genes and therefore B lymphopoiesis is halted early in the pro B cell stage (Bosma et al., 1999; Bosma et al., 1983; Schuler et al., 1986). Insertion of a completed it heavy chain and pre-BCR expression in SCID mice allows the progression to the pre B cell stage (Chang et al., 1995). Another suggested checkpoint in B lymphopoiesis occurs immediately following surface expression of the completed lgM molecule. This can again be demonstrated using the SCID model. After insertion of a completed 11 and progression to the pre B cell stage, maturity only occurs with insertion of a recombined light chain (Reichman—Fried et al., 1990). The transgenic expression of a kappa light chain allowed pre B cells to differentiate into mature B cells, although at a two to three fold lower number of cells. These experiments have given strong evidence for the stages in B cell development where cells with functional lg gene rearrangements are positively selected. It is now widely accepted that elimination of deve10ping B lymphocytes during lymphopoiesis occurs via apoptosis. Several lines of evidence support this hypothesis. In viva detection of apoptotic cells is minimal due to rapid phagocytosis of cells at early stages of death (F adok and Henson, 1998; F adok et al., 1992), but in vitra culturing techniques have given some insight into death in developing B lymphocytes. Isolation of BM and short term culture have shown that B lymphocytes undergo apoptosis. Pre, late pro and immature B cells undergo a higher rate of apoptosis, relative to the early pro and mature B cells, when removed from the BM microenvironment (Lu and Osmond, 1997). In addition, it has been shown that early pro B cells and mature B cells have elevated levels of Bel—2, a protoonco gene with anti-apoptotic potential (Li et al., 1993; Merino et al., 1994). ‘Cursory studies have also shown that precursor B cells, not expressing surface IgM, are more susceptible to apoptosis induced by the synthetic glucocorticoid, dexamethasone (Garvy et al., 1993; Lu and Osmond, 1997; Merino et al., 1994). In transgenic mice with elevated expression of Bel-2 in B lymphocytes, cell survival is heightened (Strasser et al., 1991). In Bel-2 deficient mice, B lymphopoiesis is normal through birth, but soon thereafter B lymphocytes are absent in the bone marrow, spleen and periphery (Nakayama et al., 1993; Veis et al., 1993). These observations are fundamental to this thesis, since we are thoroughly studying Gc induced apoptosis in B cell lymphopoiesis. An in depth review of the mechanisms of apoptosis and glucocorticoid affects on immune cells will be covered later. Stage Specific Gene Expression The expression of some important genes, whose products are involved in B lymphopoiesis, have also been determined. RT-PCR was used to detect these genes from sorted primary cell populations (Li et al., 1993). The recombinase activating genes, 14 recombination activating genes 1 and 2 (RAG-l and RAG-2), code for proteins involved in the recombination machinery (Schatz et al., 1989); their gene expression pattern correlates specifically with the B cell populations undergoing active rearrangement. They are upregulated in pro B cells, transiently down regulated in large, proliferating pre B cells and again are expressed in small pre B cells, undergoing light chain rearrangements. Mice lacking functional RAG-1 or 2 are unable to undergo I g gene rearrangements and B lymphopoiesis is blocked in the pro B cell stage (Mombaerts et al., 1992; Shinkai et al., 1992). Terminal deoxynucleotidal transferase (tdt) is an enzyme involved in adding random nucleotides to the heavy chain joining regions to increase diversity (Desiderio et al., 1984; Landau et al., 1987). It is expressed in the pro B cell populations, but it is not detectable in the pre B cells that have completed heavy chain rearrangements (Opstelten et al., 1986). These expression patterns correlate with the stages in which their functions are critical, therefore in addition to surface phenotype and status of Ig rearrangement, the presence of certain intracellular proteins can define the various stages of development. Recently several transcription factors have been identified which appear to be critical in the differentiation of developing B lymphocytes. PU.1 and Ikaros are two such factors. Transgenic PU.1 -/- or Ikaros -/- mice, have arrested B cell development at a stage prior to the generation of pro B cells (McKercher et al., 1996; Scott et al., 1994; Wang et al., 1996). Another transcription factor indicated in development of B lymphocytes are the E2A gene products E12 and E47. Mice with a targeted null E2A mutation fail to develop B cells early in development, before DJ rearrangements (Bain et al., 1994; Bain and Murre, 1998). B-cell-specific activator protein (BSAP-S) is a transcription factor that has been shown to be involved in heavy chain recombination (Hagman et al., 2000). It is expressed in B cells through the pro B cell stage of development and its elimination blocks the development of cells after DJ H rearrangements (Nutt et al., 1997; Urbanek et al., 1994). Nuclear factor kappa B (N FKB) is a molecule that was originally identified as a transcription factor that binds to an enhancer element for the K light chain (Sen and Baltimore, 1986). Several family members have been identified and appear to have redundant function in B lymphopoiesis, since eliminating one of these factors does not halt lymphopoiesis (Burkly et al., 1995; Sha et al., 1995). Although, the constitutive expression of a dominant negative form of the NFKB inhibitor (IKB) does result in inhibition of VKJ K rearrangement and transcription of the K light chain (Scherer et al., 1996). These findings demonstrate that specific transcription factors are involved in stage-specific differentiation during B lymphopoiesis. This underscores the complex and tightly regulated nature of the development of B lymphocytes through various differentiation stages. Microenvironmental Effects on B Lymphopoiesis The BM microenvironment is critical for B cell lymphopoiesis. In the BM, precursor B lymphocytes develop in close association with the cytoplasmic projections of reticular-like cells. Figure 1.3 is a diagram depicting the development of B cells within the context of the matrix of the microenvironment. In viva injection of radiolabeled antibodies to B lymphocytes demonstrated that the earliest progenitors are found near the bone cortex and as the cells mature they progress towards the central core of the marrow for exportation to the periphery (Jacobsen and Osmond, 1990; Jacobsen et al., 1990; 16 Osmond, 1990; Osmond et al., 1992). The microenvironment plays an integral role in lymphopoiesis as is demonstrated by high losses of B lymphocytes following their removal from the BM. It was believed that the reticular cells that the lymphocytes were in contact with were supporting lymphopoiesis by providing direct contact and soluble factors for the developing cells. Cells that provide support during the development of other cell types are called stromal or nurse cells. Here we use the term stroma as the nomenclature for the cells of the bone marrow that support B cell lymphopoiesis Long term bone marrow cultures (LTBMC) developed by Whitlock and Witte (Whitlock et al., 1984) provided a system by which B lymphopoiesis could be generated and sustained for months or even years. Whole bone marrow is placed in culture and after approximately 2 weeks an adherent layer of cells, that contains stromal cells, is established and following 3-4 weeks after initiation lymphoid cells are constitutively generated. In young cultures, through 8 weeks, the predominant lymphocytes in culture are IgM' B cells with few lgM cells. These precursor cells retain the ability to fully recombine Ig genes and differentiate into mature B cells. In older cultures the predominant cell type is lgMI. This culture system has provided a means to reliably elucidate many key players in B lymphopoiesis. Prior to the development of the B lymphopoietic culture system a culture system that promoted myelopoiesis was developed (Dexter et al., 197 7). Interestingly, a major difference between these culture systems was the addition of glucocorticoids to the medium, which promoted the development of myeloid cells. This suggested that perhaps glucocorticoids might suppress lymphopoiesis and/or promote myelopoiesis. The work presented in this thesis addresses the negative effect of glucocorticoids on B lymphopoiesis, thereby providing 17 insight into the mechanism whereby these two different culture systems can lead to the development of two different types of blood cells. Additionally, the affect of increased concentrations of Cs on myelopoiesis will be addressed in the experimental portion of this dissertation. Morphological analysis of the adherent layer of LTBMC showed two distinct cell types, a large, cytoplasmically spread, fibroblast-like cell and a macrophage-like cell (Witte et al., 1987). The fibroblast-like cells were termed stromal cells, since the B precursors developed in close association specifically with this cell type (W itte et al., 1987). Although stromal cells have not been shown to be identical in nature, some general characteristics of the cell type were established. For example, stromal cells fiom LTBMC express surface vascular cellular adhesion molecule (VCAM l) and CD44, and various components of the extracellular matrix (ECM) such as collagen IV, fibronectin and laminin are also expressed (Witte et al., 1993; Witte et al., 1987). More importantly, stromal cells are responsible for the production of various cytokines that support lymphopoiesis (Funk et al., 1995). In fact, two cytokines that will be discussed thoroughly later in this review, interleukin 7 (IL-7) and stem cell factor (SCF, kit ligand, steel factor), are produced specifically by stromal cells in LTBMC. In addition, stromal cells produce a pre B cell growth stimulating factor (stromal derived factor-1, SDF-l) and thymic stromal lymphopoietin (TSLP) (Levin et al., 1999; Nagasawa et al., 1994). The production of macrophage colony stimulating factor (M-CSF) is also specific to stromal cells, whereas the macrophages produce the interleukin 1 beta (IL-113) cytokine. Stromal cell lines have also been established from LTBMC, many of which can support both myelopoiesis and lymphopoiesis depending on the culture conditions used (Collins and Dorshkind, 1987; Dorshkind et al., 1986). Some variation in expression of proteins can be observed with different clones and the degree of support provided to B lymphocytes can also vary (Henderson et al., 1990). In fact, some BM stromal cell lines have been shown to have a negative affect on lymphopoiesis (Borghesi et al., 1997). This suggests that stromal cells have the ability to both support or inhibit B cell development, yet little is known about the mechanism by which these cells can affect the lymphocyte so differently. It has also been noted that stromal cell function can be modulated by exogenous factors that in turn can affect lymphopoiesis. For example, the addition of the aryl hydrocarbon 7,12-dimethylbenzanthracene to cultures containing B lymphocytes and bone marrow stromal cells resulted in apoptosis of the B cells (Yarnaguchi et al., 1997). This induction of death was due to either the stromal cells or stromal cell derived factors since the B lymphocytes were shown to contain very few aryl hydrocarbon receptors and, in cultures without stromal cells, the B lymphocytes did not undergo apoptosis. Estrogen also appears to down regulate B cell lymphopoiesis in long term cultures via modulating stromal cell function rather than by directly affecting the B cells (Smithson et al., 1995). Treatment of B lymphocytes alone, did not induce cell death or inhibit proliferation, whereas a reduction in B cell lymphopoiesis occurred when B cells were cultured on stromal cells that were pretreated with estrogen. Specific cytokines have also been shown to modulate the expression of surface proteins on stromal cells. IL-l B addition to stromal cell clones resulted in the upregulation of both CD54 and VCAM-1 on the cell surface, whereas IL-4 upregulated only VCAM-l and TGF-B downregulated expression of VCAM-1 on stromal cells (Dittel et al., 1993). Both CD54 and VCAM-1 have been 19 indicated in providing direct cell-cell contacts with B lymphocytes that promote B cell lymphopoiesis. Clearly various molecules that would normally be found in the bone marrow microenvironment can cause modification in stromal cell firnction, thereby potentially modifying their affect on the development of blood cells. Although the normal effects of stromal cells on B cell development have been studied in depth, little is known about the effect that Cs might have on these cells. Considering what is known of Cs negative long-term effects on B lymphocyte development this thesis addressed the potential of stromal cells to modify B cell responses to Cs and the potential for Cs to directly modify the production of cytokines known to be involved in lymphopoiesis. Cell-Cell and Cell-Extracellular Matrix Contact Considering the close association of the precursor B lymphocytes with stromal cells both in LTBMC and in situ, the role of direct contact appeared to be important in lymphopoiesis. One hypothesis regarding the close association between B lymphocytes and stroma was simply the physical retention of developing cells in the marrow. However, it was observed that suspending lymphocytes over stromal cells by using a porous membrane resulted in decreased cell viability and loss of cell differentiation, thus arguing for a more intricate role for cell-cell and cell-extracellular matrix (ECM) adhesion in lymphopoiesis (Borghesi et al., 1997; Kiemey and Dorshkind, 1987; Manabe et al., 1994). Contact between pro B cells and stromal cells enhanced survival of the pro B cells and resulted in an upregulation in the anti-apoptotic Bel-2 protein and a concomitant decrease in pro-apoptotic Bax protein levels (Gibson et al., 1996). In addition to adhesion playing a role in supporting B cell lymphopoiesis it has also been 20 suggested that lymphocytes can initiate signaling cascades in stromal cells (Jarvis and LeBien, 1995; Jarvis et al., 1997). The incubation of B lymphocytes directly on a stromal cell line, resulted in tyrosine phosphorylation at focal adhesion points in stromal cells, indicating the initiation of adhesion mediated signaling. Therefore the interactions between stromal cells and lymphocytes appear to be more complex than originally thought with each cell type potentially able to affect the other in a positive or negative manner. There are multiple potential adhesive interactions that could play a role in stromal cell supported B lymphopoiesis. Miyake et a1 (Miyake et al., 1991) attempted to identify if adhesion molecules on bone marrow stromal cells could directly affect B cell development. This was done by creating monoclonal antibodies against cell surface proteins. Two antibodies, M/K-l and M/K-2, had specific affinity to a surface molecule on stromal cells present in LTBMC and in the bone marrow. Addition of the antibodies to newly initiated LTBMC resulted in the ablation of lymphopoiesis in preestablished cultures and also resulted in the rapid detachment of lymphoid cells. This antibody recognized a molecule with sequence homology to human VCAM-1. One ligand of VCAM-1 is the 01481 integrin (VLA-4), therefore an antibody to the or subunit of VLA-4 was made to determine if it would have similar effects on lymphopoiesis (Miyake et al., 1991). In fact, it did cause rapid detachment of lymphocytes in LTBMC and its addition to newly established cultures completely blocked lymphopoiesis. This indicated that VCAM-l on stroma and VLA-4 on lymphocytes interact and facilitate lymphopoiesis. VLA-4 is also known to use fibronectin as a ligand, but blocking peptides that disrupted this interaction did not disrupt adhesion or lymphopoiesis. Additionally, irnmunostaining 21 showed that VCAM-l and VLA-4 are located together in areas of contact between stromal cells and precursor B lymphocytes (Murti et al., 1996, Jacobsen, 1996 #78). It has also been demonstrated that adhesion of human precursor B cells to stromal cells is mediated to a great extent by the interactions of VLA-4 and VCAM-l (Dittel et al., 1993). Thus, direct cell-cell interactions between stromal cells and B lymphocytes are important in the normal development of B lymphocytes in the bone marrow. Stromal Cell Derived Cytokine Support In addition to the important role of contact in stromal cell support of lymphopoiesis another critical role of stromal cells is the production of cytokines, although cytokine effects and close association between stromal cells and lymphocytes can be interrelated. Many cytokines are produced in membrane bound forms as well as soluble forms. Although close association with the stroma may augment these effects it has also been shown that soluble support via the stroma can also increase B lymphocyte survival and proliferation, although not to the extent of direct association (Borghesi et al., 1997). In 1988 a stromal cell derived factor, IL-7, was identified and purified, that supports proliferation of pre B lymphocytes (N amen et al., 1988; Namen et al., 1988). Biochemical analysis identified it as a 25-kilodalton glycoprotein and its structure of four ’ alpha helical bundles placed it in the Type-I class of cytokines. In culture, without stromal cell support, recombinant IL-7 can support prolonged proliferation of IgM' cells (Lee et al., 1988). Initially the responding cells are consistent with late pro B cells and pre B cells: cytoplasmic if, BP-l+ and IgM' (Lee et al., 1989). Extended culture periods with IL-7 ultimately resulted in an accumulation of cytoplasmic u” B lymphocytes, 22 corresponding with the early pro B cell stage. Subsequently it was discovered that IL-7 also played a similar role in thyrnocyte development (Okazaki et al., 1989; Watson et al., 1989). It induced proliferation in early CD8'CD4' thymocytes. Since then, T cells and thymocytes have been used extensively in the study of IL-7 and the IL—7 receptor (IL-7R) and relatively little work has been done using B lymphocytes. It can be surmised that, at least in part, the complexity of the B cell system has hindered research efforts in this area. In addition to inducing proliferation, IL—7 could promote survival of early B and T cells. One mechanism whereby IL-7 might promote survival is by the activation of the phosphatidyl inositol 3-kinase (PI3K) pathway (Dadi et al., 1993; Pallard et al., 1999). This survival pathway will be discussed in greater detail later in this thesis, but briefly PI3K phosphorylates AKT (protein kinase B) thereby inactivating the pro-apoptotic Bad protein. It has also been shown that IL-7 can promote survival by maintaining or upregulating anti-apoptotic Bcl-2 levels and decreasing pro-apoptotic Bax levels (Lu et al., 1999). IL-7 upregulated Bel-2 protein levels and increased cell survival in early precursor thymocytes and in a T lymphoma cell line (Kim et al., 1998; Lee et al., 1996; von Freeden-Jeffry et al., 1997). Although Bel-2 may play a role in IL-7 induced survival it may not be the exclusive survival mechanism. In mice deficient for the IL-7 receptor, overexpression of Bel-2 did not rescue B lymphopoiesis, but did promote the survival of circulating (mature) B cells (Maraskovsky et al., 1998). It should be noted that IL-7 receptor deficiency also resulted in a lack of proliferation in the developing B lymphocytes, therefore Bel-2 could potentially promote the survival of B lymphocytes but this effect would not be seen due to the lack of expansion of the cells. Therefore, the role of Bel-2 in IL-7 mediated survival has not been completely elucidated. This thesis 23 will provide further insight into the effect of IL-7 on apoptosis and cell cycle arrest in B lymphocytes induced by Cs. In 1990 the receptor for human and murine IL-7 was cloned (Goodwin et al., 1990). Binding studies revealed that the receptor exists in both a high and low affinity form. The receptor is observed on many murine cell types, both myeloid and lymphoid. The high affinity form of the receptor is found on IgM' B lymphocytes and in CD4'CD8' or single positive thymocytes. Injection of a blocking antibody to the high affinity form of the IL-7R resulted in a dramatic decrease in B cell precursors and thymocytes, although peripheral lymphocytes remained for two weeks (Sudo et al., 1993). These data supported an important role for IL-7 in lymphopoiesis. Later studies, utilizing transgenic mice, showed that elimination of IL-7 or IL-7R resulted in ablation of B lymphopoiesis, except for an odd population of immature B cells found in the spleen (Corcoran et al., 1998; von Freeden-Jeffry et al., 1995). The exact role of IL-7 in B lymphopoiesis is still being investigated. It was originally thought that the cytokine was involved strictly in proliferation and survival of precursor B cells. More recently support is growing for a role of IL—7 in differentiation as well. It was shown that IL—7Ror knockout mice have precursor B lymphocytes with normal D-J recombination of the heavy chain but there is a marked reduction in V gene recombination, correlating with the physical distance of the V gene from the recombination site (Corcoran et al., 1998). In addition, the expression of the pax-5 gene, whose product, BSAP-S, binds the heavy chain and stimulates recombination is significantly reduced. To determine if signaling through the IL7R could result in lg gene rearrangement, a mutation was made on the cytoplasmic domain of the receptor (Corcoran et al., 1996). This mutation caused elimination of the proliferative effects of IL-7 but did not affect Ig heavy chain rearrangement. The IL-7R, therefore, has been shown to promote several distinct intracellular responses through its cytoplasmic domain. Crosslinking studies showed that the IL-7Ror subunit was associated with the common y chain (yo) of the IL-2 receptor (Noguchi et al., 1993). Association with yo enhanced the binding affinity of IL-7Ror and increased receptor internalization upon ligand binding. In addition, blocking antibodies to yo inhibited IL-7 induced proliferation of a B cell line (Kondo et al., 1994). IL-4, IL-9 and IL-15 also contain the yc as part of their receptors. The cytoplasmic tail yc has been shown to associate with janus kinase family member Jak3 and stimulate STAT signaling resulting in proliferation (Malabarba et al., 1995; Miyazaki et al., 1994). Src kinasesgand phosphatidyl inositol 3- kinases (PI3K) have also been shown to be activated in response to IL-7. Another cytokine that appears to be involved in B cell lymphopoiesis is stem cell factor (SCF). SCF is a glycosylated homodimer that plays a broad role in aiding the commitment of hematopoietic progenitors to the various lineages when present in combination with specific grth factors for those specific cell types (W itte, 1990). To date, the only cytokine that has been shown to be critical in B lymphopoiesis is IL-7, but SCF synergizes with IL-7 to increase proliferation of pro and pre B cells (McNiece et al., 1991). In addition, this combination of cytokines results in differentiation of 3220' murine bone marrow cells into B220+ B lymphocytes, whereas neither cytokine, alone, can promote the commitment to the B lineage. SCF can be produced in both soluble and membrane bound forms. Experiments whereby only the soluble isoforrn of SCF was 25 produced indicated that the membrane bound form likely is the active form of the cytokine (Miyazawa et al., 1995). The receptor for SCF is c-kit which is a receptor tyrosine kinase expressed on the surface of many hematopoietic cells; on B cells its expression is restricted to the pro B cell stage (Rico-Vargas et al., 1994). Like IL-7, SCF can initiate multiple signaling cascades in the cell and it has shown some ability to promote survival. It may, in part, promote survival through the PI3K induced activation of AKT (Blume-Jensen et al., 1998). Other activating cell signaling pathways that have been shown to be activated by SCF binding of c-kit are the JAK/ STAT, MAPK and Src family transduction pathways (Linnekin, 1999). Another tyrosine kinase receptor, flt-3, appears to parallel c-kit in its function on B lymphocytes. Like c-kit, it is expressed on pro B lymphocytes and its ligand, FL, can induce proliferation in pre-pro B cells (Hirayama et al., 1995; Matthews et al., 1991). Considering the similarities and the differences between the biological effects and signaling pathways of IL-7 and SCF this thesis investigates the effect of both of these cytokines on background and Cs-induced apoptosis to determine whether they have similar, different or additive effects on B lymphocytes. Since the discovery of the above-mentioned factors, other stromal cell derived molecules that may play a role in lymphopoiesis have been discovered. Two such factors are thymic stromal lymphopoietin (TSLP) and stromal derived factor-l (SDF-l). TSLP has been shown to act as a costimulatory factor with IL-7 and promotes proliferation through the immature stage of development (Friend et al., 1994). Interestingly, it was shown that the receptor for TSLP is composed of the IL-7Ror chain and the novel TSLP receptor (Levin et al., 1999). Whereas TSLP appears to affect later stages in B cell 26 development, SDF-l appears to elicit its effect on earlier B cells. SDF-l can induce proliferation in pre B cells and has also been shown to be a chemoattractant factor for various progenitor populations (Aiuti et al., 1997; Nagasawa et al., 1994). Mice deficient in SDF-1 or its receptor, CXCR4, have a block in fetal lymphopoiesis before the pro B cell stage. It is becoming clear that many factors are involved in the successful generation of mature B lymphocytes and that they may elicit their effects at different developmental stages. 27 Summary The generation and development of B lymphocytes in adult mammals is a tightly regulated, multi-stage process that occurs in the highly heterogeneous bone marrow. It is the complex and critical rearrangement of Ig genes that appears to drive the development of B cells. Two advances have aided in progress in this field, flow cytometry and a long term culturing technique that promotes B cell lymphopoiesis. Flow cytometry has provided a powerful tool for analysis of the B lymphocytes in their various stages of development and LTBMC has allowed for the discovery of several microenvironmental support factor critical in lymphopoiesis. lL-7 is a stromal cell derived factor that, in particular, has been shown to be critical in early stages of B lymphocyte development in mice. Without IL-7 pro and pre B lymphocytes cannot survive or proliferate long-term in culture. In addition to LTBMC, transgenic mice have provided a system for defining the in viva effects of many of these lymphopoietic molecules. Taken together, it is becoming clear that during the development of B lymphocytes a myriad of factors work together to promote cell survival, proliferation and differentiation and that response to these factors depends on the stage of development. Each developmental stage can be defined by selective expression of relevant proteins and surface markers and proper recombination of lg is critical for advancement to maturity. Even with the experimental advances that have been made in the last two decades, much work remains in eliciting the mechanism of B lymphopoiesis, especially in determining how normal physiological changes may modulate this process. This lab has a unique advantage in studying normal bone marrow B lymphocytes in various stages of development in part due to access to a multi- pararneter flow cytometer. Considering this advantage we have chosen to investigate the 28 effect of Go on normal developing B lymphocytes from progenitor through maturity. As will be further discussed in the next portion of this literature review, Gc can have dramatic negative effects on B cells, yet little is known of this phenomenon. Here we will determine in viva and in vitra responses of normal B lymphocytes to Go, as well as the potential modification of these responses by stromal cells, IL-7 or SCF. GLUCOCORTICOIDS AND THE IMMUNE SYSTEM Background Glucocorticoids (Gc) are steroid hormones that are produced in the cortex of the adrenal gland. Their basal production is critical to mammals, since the lack of Ge can result in death. Natural variations in serum levels of Ge occur in a diurnal fashion with levels in humans being highest in the morning hours and in nocturnal animals, such as rodents, levels are highest at night (Nichols and Tyler, 1967). Levels of Go are regulated by adrenal corticotrophin hormone (ACTH; corticotrophin) produced in the anterior pituitary gland. ACTH is upregulated by vasopressin and corticotrophin releasing hormone (CRH), whereas glucocorticoids cause its down-regulation (Engler et al., 1999). Together this cascade of hormones link together neurobiology and endocrinology and is referred to as the hypothalamic-pituitary—adrenal (HPA) axis. The major external effector of the HPA axis is stress. Stress can be defined as anything that disturbs homeostasis. Some stressors, such as exercise, are thought of as positive and elicit a transient increase in Go. Other stressors induce a chronic increase in the production and circulation of Ge. Classic examples of negative stress are seen in burn patients, trauma victims, malnourishment and depression. It is chronic exposure to Gc that can have a variety of negative effects including compromised immune function. Therefore it is this phenomenon that is of interest here. As early as 1947 the relationship between stress and Ge elevation was suggested (Selye, 1947). Just 2 years later, it was proposed by Hench (Hench, 1949) that Go mediate anti-inflammatory action. In fact, for decades synthetic Gc such as prednisone 30 and dexamethasone have been used as anti-inflammatory drugs in diseases such as arthritis, autoimmunity and asthma. This provides both pharmacological as well as physiological reasons for studying the effects of glucocorticoids. Along with the noted anti-inflammatory effects of glucocorticoids, immunosuppression is evident. In patients with the aforementioned stressors, there is often a marked decrease in peripheral lymphocytes (lymphopenia) and ultimately compromised immune defense against infection. This lab became interested in glucocorticoid induced irrnnunosuppression while studying zinc deficiency in mice. It was observed that zinc deficiency resulted in chronic elevation of the naturally occurring Gc, Corticosterone (Cs) (in humans the major circulating glucocorticoid is cortisol) (DePasquale-Jardieu and Fraker, 1980). Adrenalectomy of these mice resulted in protection of the immune system during zinc deficiency (DePasquale-Jardieu and Fraker, 1980). There is evidence that suppression of lymphopoiesis is at least partially the mechanism whereby Gc induces lyrnphopenia. Thyrnic atrophy has been a hallmark of elevated Gc, indicating that it may decrease the numbers of developing T lymphocytes. It was later determined that indeed developing T lymphocytes were being lost, most notably the double positive (CD4+CD8+) precursor thymocytes (Cohen, 1992). The effect of elevated Go on B cell lymphopoiesis was not investigated until relatively recently, when our lab and others analyzed the BM compartment in mice with elevated levels of the steroid. Garvy et a1 (Garvy et al., 1993) demonstrated that chronic elevation of Cs in mice resulted in selective depletion of the developing B lymphocytes in the BM. It was also noted that IgM‘ precursor B cells were more sensitive to these Cs induced losses and suggested that the mechanism may be due 31 to Go mediated cell cycle arrest and apoptosis in these cells. The role of glucocorticoid induced apoptosis and cell cycle arrest will be thoroughly discussed in the apoptosis section of this review. In addition to the negative regulatory effect of elevated Go on developing B and T lymphocytes, it appears that Go can also play a role in normal lymphocyte selection and even survival under some conditions. In the thymus it was recently shown that corticosterone could be produced by epithelial cells (Vacchio et al., 1994). In experiments using radiolabeled steroid precursors, thymic tissue was capable of complete glucocorticoid metabolism, suggesting that the thymus contains all of the necessary enzymes for steroid biosynthesis (Lechner et al., 2000). In thymic organ cultures, the elimination of Cs production resulted in the loss of cells that would normally be positively selected by crosslinkage of the TCR (V acchio and Ashwell, 1997; Vacchio et al., 1998). In addition to promoting survival of “selected” T cells, it was observed that glucocorticoids could inhibit apoptosis induced by serum depletion in a Bel-2 expressing T lymphoma cell line (Huang and Cidlowski, 1999). To date, there does not appear to be any evidence for the role of Go in the positive selection of B lymphocytes, but given the parallels between the two processes it seems likely that Go may play a similar role in B lymphopoiesis. These observations underscore the complex role that Go plays in the immune system and suggest that the steroid may act as an effector for cell survival or cell death depending on the context of the cell. This thesis addresses Gc induced downregulation of B lymphopoiesis, induction of apoptosis and cell cycle arrest on normal B lymphocytes, thus defining conditions whereby Gc has a negative effect on B cell development. 32 Glucocorticoid Biochemistry As previously stated the synthesis of glucocorticoids occurs in the cortex of the adrenal gland. The precursor for all steroid hormones is cholesterol and the majority of cholesterol used in steroid hormone biosynthesis is trafficked into the adrenal gland by either low density lipoprotein (LDL) in humans or high density lipoprotein (HDL) in nrice. In the adrenal the majority of cholesterol is stored in lipid vesicles until stimulation of synthesis (Vinson et al., 1992). Figure 1.4 is a diagram of the pathways of steroid biosynthesis and the enzymes involved. The first step is the generation of pregrrenalone by the cytochrome P450 side chain cleavage enzyme. This is followed by 3 B- hydroxysteroid dehydrogenase A 4,5-isomerase reaction resulting in progesterone. In humans a Nor-hydroxylase followed by a 21- and 113- hydroxylase reaction generates cortisol. It has been long believed that the Nor-hydroxylase is absent in rodents, therefore the 21-hydroxylase and 1 1 B-hydroxylase reactions result in the production of corticosterone, which differs from cortisol only in lacking a hydroxyl group at carbon 17. More recently though, evidence for l l B-hydroxylase activity in murine in vitra adrenal cultures has been demonstrated (Touitou et al., 1990). In addition to synthesis of glucocorticoids, mineral corticoids and sex hormones are also derived from cholesterol in the adrenal gland. Following synthesis and release from the adrenal gland, Gc are transported through the system bound to the corticosteroid binding globulin (CBG, transcortin) (W estphal and Devenuto, 1966). However, it is believed that the biologically active form of Go is not bound to CBG, but exists in a fi'ee state (Bright, 1995; Mendel et al., 1989). A dynamic equilibrium between free and bound Go is apparent, therefore suggesting that 33 regulation of Go binding to CBG might result in partial control of Go activity. In fact, CBG levels have been shown to decrease in response to stress such as surgery in humans (V ogeser et al., 1999). These decreases have also been observed in rodents exposed to classical stressors (Tannenbaum et al., 1997). It is believed that the downregulation of CBG results in an upregulation of free Go. In addition, at sites of inflammation it has been shown that elastase can cleave CBG, drastically reducing its affinity for Go (Pemberton et al., 1988). These results suggest that not only are Go actions controlled by upregulated production of the steroid but also by modulation of its availability. The Glucocorticoid Receptor Following transport throughout the system, Gc elicit their effects by binding to the glucocorticoid receptor (GcR). The GcR belongs to the steroid hormone receptor (SHR) superfamily of proteins and is present in nearly all mammalian tissues at thousands of molecules per cell (Ballard et al., 1974). This family includes receptors for the sex hormones, thyroid hormones and arylhydrocarbons (such as dioxin). In 1985 the human GcR was cloned, demonstrating the existence of both or and B isoforrns (Hollenberg et al., 1985). The or isoform was shown to be the ligand binding form composed of 777 amino acids whereas the non-ligand binding B form was missing 15 amino acids at the carboxyl terminus. It was recently shown that the GcRB is expressed at modest yet varying levels and may act as a ligand independent negative regulator of GcR (Bamberger et al., 1995). The or form of GcR has several functional domains, which are consistent throughout the SHR family. The protein contains a ligand binding domain, a DNA binding domain and two transactivating domains (AF-l and AF -2). A diagram of 34 the GcR is shown in Figure 1.5. Sequence analysis and crystallography of the DNA binding domain revealed the presence of two zinc finger motifs (Freedman and Luisi, 1993). The loop structure formed by zinc binding between four cysteine residues is a common feature of DNA binding proteins. The ligand binding domain is located at the carboxyl terminal end of the protein. Radiolabeled ligand affinity studies have shown that dexamethasone binds the receptor with high affinity. The natural steroids, cortisol or corticosterone, bind with modest affinity and sex hormones bind with only low affinity. The transcription activating domain, AF- 1 , is located at the amino-terminal end of the receptor and its activity is independent of ligand binding (Almlof et al., 1997). In contrast the AF -2 transactivating domain is located in the ligand binding region and is dependent on steroid binding (Danielian et al., 1992). Steroid binding is thought to cause a transformation of the AF-2 region, allowing coactivating proteins, which mediate transcription, to bind the SHR (Feng et al., 1998; Wagner et al., 1995). Early studies showed that under hormone free conditions a large form of GcR was located in the cytosolic fraction of cell extracts (Baxter and Tomkins, 1971). Interestingly, the addition of steroid resulted in a lower mass receptor that could be isolated from nuclear fractions (Higgins et al., 1973; Schmidt et al., 1982). This led to the hypothesis that Go binds to a large inactive receptor complex in the cytosolic compartment resulting in release of associated factors and translocation to the nucleus. It was later determined by irnmunohistochernical studies that indeed Gc binding to the GcR resulted in translocation of the receptor from the cytosol to the nucleus. In pursuit of identification of the putative associated proteins of the inactive GcR complex, cytosolic receptor was isolated and purified (Gustafsson et al., 1989; Rexin et al., 1992). A 90- 35 kilodalton protein was isolated and firrther immunolabeling and sequence analysis showed the protein to be heat shock protein 90 (hsp90). Point mutations in the ligand binding domain of the GcR resulted in loss of hsp90 binding. This suggested that hsp90 bound the ligand binding domain of the OCR (Chakraborti et al., 1992). Further coprecipitation studies of the GcR suggested that there were a variety of associated proteins. To date the cytosolic form of the OCR has been shown to have two hsp90 molecules, hsp70, hsp56 and ‘a small immunophillin p23 molecule associated with it. Upon binding of steroid to the receptor, this protein complex dissociates allowing bound receptor to translocate to the nucleus. GcR are phosphorylated proteins that undergo hyperphosphorylation upon agonist ligand binding (Bodwell et al., 1998). The hyperphosphorylation appears to be critical in GcR regulation of transcription, since receptors mutated at certain phosphorylation sites have severely impaired transcriptional transactivation. Additionally, phosphorylation occurs in regions of association with cell cycle dependent kinases. Interestingly, phosphorylation of the OCR appears to be dependent on the cell cycle status. In the experiments performed thus far, ligand binding induced hyperphosphorylation in the S phase of the cycle, but not during szM (Hu et al., 1997). Collectively these data suggest that the regulation of GcR activity is modulated by more than just ligand binding. Glucocorticoid Receptor Signaling In 1974 it was observed that the insect glucocorticoid, ecdysteroid, caused puffing on polytene chromosomes of Drasaphila melanagaster, suggesting that Go may elicit its effect by modulating transcription (Ashbumer et al., 1974). It was later determined that, 36 in fact, GcR was affecting transcription of certain genes via binding to a DNA consensus sequence, 5’-GGTACAnnnTGTTCT-3’ (Beato et al., 1989; Chandler et al., 1983; Karin et al., 1984). The GcR DNA binding site was termed the glucocorticoid response element (GRE). Exactly how receptor binding to DNA regulates the transcription machinery has not been determined. There is evidence for direct interactions between GcR and the transcription initiation complex and for association of the GcRs with coactivator proteins (Tsai and O'Malley, 1994). Members of the p160 family of proteins (steroid receptor coactivator 1 [SCRl ], GcR-interacting protein 1 [GRIP1]) were found to associate with the AF -2 region of nuclear receptors and act as coactivators of transcription following ligand binding (Hong et al., 1996; Onate et al., 1995). Recently these coactivators have been found to promote transcription activation by binding the histone acetyltransferases (HAT) CBP or p300 (Torchia et al., 1997). The HAT activity promotes transcription by acetylating histone H3 of chromatin, thus modifying chromatin structure to allow transcription machinery access to the DNA (Kuo and Allis, 1998). In addition to HAT activity CBP and p300 have been shown to directly bind the basal transcription machinery. Therefore, GcR, may play a direct role in stimulating transcription by controlling chromatin structure and perhaps by directly interacting with the transcription initiation complex. Active ligand-bound GcR has also been implicated in cross-talk with other signaling cascades. Interestingly, it was long believed that steroid hormones mediated transcription activation, but recently it was shown that SHR can also negatively regulate transcription of certain genes. Gc appears to negatively regulate transcription of AP-l and NF -kB regulated genes. AP-l promotes transcription of immediate early genes 37 stimulated by growth factors and NF -kB stimulates transcription of genes involved in immune cell activation and inflammatory cytokines. In vitra experiments suggested that GcR negatively regulates AP-l transcription by directly interacting with the fos/jun heterodimer (Heck et al., 1994; Lucibello et al., 1990). This interaction is ligand dependent, but does not require GcR binding to the DNA. In contrast, the repressive effects of GcR on NF-kB activities do not appear to be mediated by direct interactions between the two proteins. NF -kB is bound to 1ch in its inactive cytoplasmic form [reviewed in Hatada, 2000 #323]. Activation of NF-kB requires the degradation of 1ch followed by nuclear translocation and transcription initiation. It appears that GcR inhibits NF-kB activity by upregulating 1ch gene transcription, thus rendering NF-kB to its inactive form (Auphan et al., 1995). Cross-talk between nuclear receptors and transcription factors is an active area of study and is proving to be important in the understanding of how Gc affects cells. In addition to the direct effect of GcR on cellular processes, it is also becoming apparent that hsp90 may also mediate signaling following its dissociation fi'om the GcR. The creation of a fusion protein containing the GcR hormone binding domain (also the hsp90 binding domain) and the adenovirus ElA protein, resulted in steroid responsiveness by the otherwise innocuous ElA (Picard et al., 1988). The results suggested that release of the inhibitory protein complex may result in at least a partial steroid effect. In addition, hsp90 also interacts with pp60"'src kinase and has been shown to diminish its kinase activity (Xu and Lindquist, 1993). It is becoming apparent that signal transduction via the GcR is a multi-faceted process that can influence or be influenced by various other cellular signals. 38 Another mechanism for regulating GcR mediated signaling may be via the direct transcriptional upregulation or downregulation of the receptor. It has been shown that Go itself can downregulate the transcription of GcR, ultimately resulting in decreased levels of the receptor protein (Bumstein et al., 1994; Cidlowski and Cidlowski, 1981). In addition to the negative effect on transcription, Gc binding to the OCR results in destabilization of the receptor. Recently, the phosphorylation state of the GcR had been implicated in the stability of the receptor, with hyperphosphorylation induced by ligand binding correlating with decreased receptor stability (Bodwell et al., 1998). Abolishment of the eight phosphorylation sites, either individually or in combination, has shown that decreased levels of phosphorylation correlates with decreased transactivation of responsive genes and increased receptor half-life (Webster et al., 1997). The exact mechanism of GcR stability and modification is not well understood, although it is becoming clear that GcR mediated pathways are tightly regulated. 39 Summary Glucocorticoids are hormones that have a diverse range of effects on many physiological functions in mammals. Here we have focussed on the effect of Go on the immune system, specifically the development of B lymphocytes. At basal concentrations Gc appear to play a positive role in normal homeostasis during thymocyte development, yet the effect of basal levels of Go on B lymphocyte development has not been confirmed but can be surmised to likely play a similar role. A chronic increase in Ge production and circulation, due to the chronic activation of the HPA stress axis, clearly has a negative affect on the immune system. This lab has had an integral role in determining how the induction of the stress axis via zinc deficiency has had a negative affect on the immune system and B lymphocyte development and the specific downregulation of BM B lymphocytes by chronic Cs elevation. Since then identification techniques and methodologies have advanced dramatically in this field, allowing for further investigation into the effects that Go have on developing B lymphocytes in viva and in vitra. The downregulation of developing B lymphocytes by Go likely occurs by both a decrease in cell survival (apoptosis) and a downregulation in proliferation. These phenomenon will be discussed in the next section of this review. Additionally, many in vitra experiments studying B lymphocyte responses to Go have utilized the synthetic Gc, dex, yet little is known of the effects of the naturally occurring Cs on B lymphocytes. This thesis will thoroughly investigate the effects of Cs on normal BM B lymphocytes from mice to provide new insights into the mechanism of Ge negative actions on these developing immune cells. 40 APOPTOSIS History In 1972 Kerr and Wyllie (Kerr et al., 1972) observed that in normal tissue development, certain cells underwent cell death in an ordered manner with distinct morphological characteristics. This form of death was subsequently termed apoptosis and varied fiom necrotic cell death in that cytoplasmic contents were not released to the extracellular environment, thus avoiding an uncontrolled immune response. Nearly a decade passed before a modest interest in apoptosis was scientifically revived. It wasn’t until the 1990’s that the fundamental importance of apoptosis became widely recognized and sparked intensive research in this area. It is now clear that cell death is germane to the homeostasis of multicellular organisms and dysregulation of it can have serious health consequences. Excessive apoptosis has been indicated as playing a central role in several diseases. A classic example is apparent in the etiology of acquired immunodeficiency syndrome (AIDS). Central to the manifestation of AIDS is the excessive elimination of helper T lymphocytes by apoptosis (Meyaard et al., 1992; Wang et al., 1999). The faulty induction of apoptosis is also apparent in neurodegenerative diseases, such as Parkinson’s and Alzheimer’s (Waggie et al., 1999). Along with excessive apoptosis, the failure to undergo appropriate cell death also can contribute to disease. In B cell follicular lymphoma, a genetic translocation results in the upregulation of the anti-apoptotic protein Bel-2 (this protein will be discussed in detail later in this chapter) (Bakhshi et al., 1985; Cleary et al., 1986). The malignant cells in this type of cancer did not exhibit increased 41 proliferation as was thought to be a hallmark of malignancies, rather they failed to die normally. Apoptosis is also thought to play a significant role in the proper development of B and T lymphocytes with death being induced in cells with faulty Ig or T cell receptor rearrangements, respectively. The failure of B or T cells with anti-self antigen receptors to undergo proper elimination can lead to autoimmune diseases such as Systemic Lupus Erythematosis, asthma and rheumatoid arthritis (O'Reilly and Strasser, 1999). Relative to this thesis, our lab has implicated apoptosis in the mechanism of downregulation of developing B lymphocytes in response to Go elevation (Garvy et al., 1993). A downregulation in the development of B and/or T lymphocyte can result in decreased circulating cells (lymphopenia) that ultimately results in a compromised immune response. Clearly the past ten years have brought apoptosis to the forefiont in the study of the etiology of many diseases and thus firrther study of this process could significantly contribute in the understanding, the control, and potentially the cure of sometimes fatal diseases. Apoptotic Morphology and Detection Morphologically most apoptotic cells undergo a decrease in cell size and volume, nuclear condensation, blebbing of the plasma membrane and eventual formation of membrane bound apoptotic bodies. During early research on apoptosis a seminal observation was made in thymocytes induced to undergo apoptosis with dexamethasone (Wyllie, 1980). It was shown that in the apoptotic nucleus an endonuclease is activated that caused DNA fragmentation, consistent with cleavage at the intemucleosomal linker regions of chromatin. As previously stated, apoptotic cell death varies from necrotic cell 42 death in that the cytosolic contents remain membrane bound whereas during necrosis the dying cell swell and ultimately releases its contents to the surrounding milieu. During apoptosis the plasma membrane undergoes changes in fluidity, resulting in a transition of phosphatidylserine (PS) from the inner to the outer leaflet of the phospholipid bilayer (Fadok et al., 1992). The exposure of PS then triggers rapid engulfinent by phagocytic cells. Recently, a protein kinase C delta mediated scramblase has been proposed as the mechanism whereby PS is exposed on the outer leaflet (Frasch et al., 2000). Another firndamental difference between apoptosis and necrosis is the lack of intemucleosomal DNA cleavage in necrotic cells. Late in necrosis DNA cleavage can occur, but results in indistinct fragmentation, in contrast to the nucleosomal cleavage pattern seen in apoptosis (Bicknell and Cohen, 1995). Early studies in apoptosis depended on the morphologic analysis of the dying cells, therefore more sophisticated research such as the analysis of cell death in heterogeneous tissue was difficult if not impossible. Later an electrophoretic method was developed that could determine the degree of apoptosis in cell populations by the presence of a DNA “ladder”, indicative of intemucleosomal DNA cleavage. This methodology aided somewhat in studying apoptosis, but it only provided quantitative data and not qualitative identification of dying cells. This lab made a significant contribution in methodology for the identification of apoptosis in whole cells (Telford et al., 1991; Telford et al., 1994). Cells could be fixed in ethanol and the DNA stained with propidium iodide whereby cells containing hypodiploid DNA were determined to be apoptotic. This method allowed for the analysis of heterogeneous tissues, since mixed populations could be phenotyped for specific cell types and analyzed on an individual 43 cell basis. Prior to this, apoptosis in bone marrow B lymphocytes, the major topic of this thesis, could not be adequately studied unless complex purification protocols or cell lines were utilized. Apoptosis Activation (Caspases and Mitochondria) In the 1990’s intensive research efforts were applied to the identification of a central biochemical pathways that lead to the commitment to and execution of apoptosis. Many intracellular processes were implicated as central death effectors, but further investigation showed apoptosis initiation could occur independent of the proposed effectors. Calcium influx, generation of reactive oxygen species and endonuclease activation were just a few processes implicated as central to the activation of apoptosis (reviewed in Schwartzman and Cidlowski, 1993). Each of these were later shown to be common features of apoptosis under many circumstances but not necessary for death. In recent years with the effort of many laboratories a model of the central pathway of apoptosis has been generated (Figure 1.6). It integrates both mitochondrial changes and the activation of a class of proteases called caspases. The model also defines two distinct types of apoptosis, receptor-mediated death that is initiated in healthy cells (caspase dependent) and death that occurs in damaged or neglected cells (mitochondria dependent). A brief overview of receptor-mediated apoptosis verses apoptosis in damaged cells will be presented below, followed by a detailed overview of the central death pathway. Briefly, a classic example of cells that are healthy but eliminated is evident in the immune response. This type of autoregulatory death occurs via the crosslinking of death receptors of the tumor necrosis factor receptor (TNFR) superfamily 44 (Nagata, 1997). This results in the activation of the caspase cascade and ultimately apoptosis. Mitochondrial depolarization and the release of cytochrome c is a common feature seen in this type of death, but it is not necessary for it to occur (Los et al., 1995). On the other hand, cells that undergo apoptosis due to neglect or damage depend on mitochondrial changes but caspase activation is not required for death. The activation of caspases is common and responsible for the morphology associated with apoptosis, but blocking caspase activity does not block cell death. Neglect by serum deprivation or DNA damage are two situations whereby blockage of caspase activity does not inhibit cell death, and the death that occurs is more similar to necrosis in morphology (McCarthy et al., 1997; Mills et al., 1998). Research on cell death in the nematode, Caenarhabditis elegans, has led to fundamental discoveries of the central molecular mechanisms of apoptosis. Genetic mutants with cell death defects were created and two proteins called CED-3 and CED-4 were found to be activators of apoptosis. CED-3 was later cloned, and shown to have homology to the mammalian cysteine protease, interleukin-1B converting enzyme (ICE) (Yuan et al., 1993). At the time, ICE was known to process IL-l B into an active form, but it had not been implicated in apoptosis. Further studies showed, indeed, that ICE overexpression could itself induce apoptosis (Miura et al., 1993). Since then several ICE- like proteases have been identified and were renamed caspases. To date 13 caspases have been identified. A unique characteristic of the caspase family is their specificity for cleavage after aspartate residues in proteins (Howard et al., 1991). There are many intracellular substrates for caspase cleavage that once cleaved can elicit a biological effect. The cleavage and activation of these proteins leads to the distinct morphology 45 seen in apoptosis. For example, caspase 3 has been shown to cleave and degrade proteins involved in the cytoskeleton, such as actin (Kayalar et al., 1996). In addition to generating morphological changes, the caspases can also act to amplify the death pathway by cleaving anti-apoptotic proteins such as Bel-2, rendering them pro-apoptotic (Cheng et al., 1997). Therefore many of the cellular changes associated with apoptosis are due to caspase activity. Along with action on cellular substrates caspases can cleave and activate themselves, thus rendering a cascade of activated caspases. The caspases exist in an inactive procaspase form. Both cytochrome c from the mitochondria and the newly identified CED-4 homologue, apoptotic protease activating factor (APAF-1), are necessary for caspase activation in apoptosis due to neglect or cell damage (Zou et al., 1997). Cell neglect, such as lack of growth factors or cell damage can result in the depolarization of the mitochondrial membrane and the release of cytochrome c (Vander Heiden et al., 1997). Cytochrome c is thought to bind APAF-l and activate binding to procaspase 9 resulting in autocleavage of the procaspase and thus formation of an active caspase 9 (Srinivasula et al., 1998). It is then believed that caspase 9 activates caspase 3, ultimately resulting in a cascade of caspase activity (Hakem et al., 1998). Caspase activation due to TNF R family mediated apoptosis (death in otherwise healthy cells) initiation proceeds by a different mechanism. TNFR family members that promote death have a homologous cytoplasmic region termed the death domain (DD) (Nagata, 1997; Tartaglia et al., 1993). The receptors include FAS (CD95), TNFR-I and the death receptors DR3, 4 and 5. The DD allows recruitment of adaptor molecules that can directly bind caspase 8 (Boldin et al., 1996; Muzio et al., 1996). Receptor 46 crosslinking via ligand binding is then thought to promote the cleavage of procaspase 8 into its active form, followed by activation of the caspase cascade. Mitochondrial membrane depolarization and the release of cytochrome c typically occurs in this death pathway, but it is not essential. Therefore it appears that there are at least two distinct mechanisms for initiating the caspase cascade, one that is dependent on the mitochondrial release of cytochrome c for caspase activation and another that can directly activate caspases via receptor ligation. Apoptosis Inhibition (Bel-2 Family of Proteins) Before the discovery of caspases the homologue for the C. elegans cell survival protein CED-9 was discovered. As previously stated, a gene translocation in B cell follicular lymphoma resulted in the discovery of Bel-2. It was later discovered that CED- 9 and Bel-2 shared sequence and functional homology (Hengartner and Horvitz, 1994). Prior to this observation it was discovered that Bcl-2 could functionally replace CED-9 as a cell survival factor in C. elegans (V aux et al., 1992). Bel-2 is part of a large family of proteins. The proteins are defined as having at least one of four Bel-2 homology domains (BHl, 2, 3 or 4). Interestingly, several members of the family have anti-apoptotic potential, whereas others can induce apoptosis. Family members that promote survival (e. g. Bel-2 and Bcl-X long isoform; Bel-XL) contain each of the BH domains. Death promoting members (e. g. Bax, Bad and Bel-X short isoform; Bcl-Xs) vary in the number of BH domains. Rather than functioning independently, it is commonly accepted that the ratio of pro.death versus anti-death factors determines the activity. It is thought that heterodimerization, between pro-apoptotic and anti-apoptotic proteins, disrupts activity 47 by disturbing homodimerization. Therefore, the protein in excess will form functional homodirners. It is also believed that the majority of Bel-2 family members are targeted to membranes via a transmembrane tail (Nguyen et al., 1993). In fact early observations suggested that Bel-2 elicited its effects by maintaining the mitochondrial membrane potential. The exact function of the Bel-2 family members has not been fully elucidated, but in recent years progress has been made. In 1996 the structure of Bcl-XL was determined using x-ray crystallography and nuclear magnetic resonance (Muchmore et al., 1996). This led to speculation that Bel-XL could potentially control mitochondrial membrane homeostasis via pore formation due to structure similarities to bacterial toxins that function by forming membrane pores. Since then Bel-XL, as well as Bel-2 and pro- apoptotic Bax, have been shown to form ion channels in synthetic lipid bilayers (Minn et al., 1997). Functionally, Bel-2 has been shown to maintain mitochondrial polarization and block the release of cytochrome c into the cytosol (Marchetti et al., 1996). Bax, on the other hand, promotes mitochondrial membrane depolarization and cytochrome c release from the inner mitochondrial membrane space (Jurgensmeier et al., 1998). Observations that Bax-regulated release of cytochrome c could occur prior to membrane depolarization had also been made. This suggested that perhaps Bax might function as a channel for cytochrome c release, but this has not been proven and remains a subject of debate. Although the mechanism of action of these proteins have not been completely defined, it has become widely accepted that the Bel-2 family members do play an important role in regulating mitochondrial activity in apoptosis. 48 Recent studies on the regulation of Bel-2 family members, have provided evidence that their function can be modulated by intracellular signaling cascades. It was commonly believed that the major regulatory mechanism was via expression levels of the anti and pro-apoptotic molecules. Studies on the pro-apoptotic member, Bad, have indicated phosphorylation as another regulatory mechanism (Zha et al., 1996). Bad was discovered in 1995 (Yang et al., 1995). It varies from other pro-apoptotic family member in that it contains only the BH3 domain and does not have a membrane binding motif. Bad has been shown to be phosphorylated on the BH3 domain. This results in association with the cytosolic protein 14-3-3. Further investigation showed that in an IL- 3 dependent cell line, IL-3 addition resulted in activation of the protein kinase AKT (or protein kinase B) (Datta et al., 1997). AKT activation led to the phosphorylation and cytosolic sequestration of Bad by 14-3-3. Removal of IL—3 resulted in the dephosphorylation of Bad and subsequent heterodirnerization with Bel-XL. This resulted in inactivation of Bcl-XL function and allowed for Bax homodimerization and pro- apoptotic function. Interestingly, the cytokine lL-4 has also been shown to activate AKT and promote thymocyte survival (Cerezo et al., 1998). A proposed model of Bad function is presented in Figure 1.7. These fairly recent findings stimulate the hypothesis that perhaps cytokine induced survival may at least in part proceed by the inactivation of Bad activity. Cytokine induced survival will be discussed further later in this chapter. In addition to phosphoryl-regulation through Bad, direct phosphorylation of Bel-2 may be another regulatory mechanism. The Raf-l serine/threonine kinase has been shown to associate with Bel-2 at the mitochondrial membrane (Wang et al., 1994). It can phosphorylate Bel-2 leading to its inactivation. This presents an interesting paradox, 49 since the Raf-l activating MEK pathway promotes cell survival, rather than cell death. The SAPK pathway has also been indicated in Bel-2 phosphorylation (Maundrell et al., 1997). Considering the death promoting activity of the SAPK pathway, inactivation of Bel-2 is a logical progression. It is clear that the regulation of Bel-2 family members via cell signaling cascades has not been fully elucidated and will remain an active area of research. Bel-2 Family Members in B Cell Lymphopoiesis The role of Bel-2 family members in the development of mammals has been addressed by creating gene knockout mice. Both Bcl-2 -/- and Bel-XL -/- mice result in lethality (Kamada et al., 1995; Motoyama et al., 1995; Veis et al., 1993). Bel-2 -/- mice die around three weeks after birth as a result of renal failure. In contrast, Bel-XL knockout mice die prior to birth at approximately embryonic day 13. Lethality appears to be due to faulty development of the brain. The immune system in both models also exhibits dysregulation. In Bel-XL -/- mice, lymphocyte development in the embryonic liver is blocked. To determine the effects of Bel-XL on lymphopoiesis, Bel-XL deficient embryonic stem cells were used to generate chimeric mice (Ma et al., 1995). In this case, survival of developing thymocytes and B lymphocytes was dramatically reduced. The Bel-2 -/- mice initially generated mature B and T lymphocytes, but by three weeks of age both B and T cells were significantly reduced. These results suggested that both Bel-XL and Bel-2 were independently involved in generating functional lymphocytes. Bel-XL appears to play a significant role in early generation of lymphocytes, whereas Bel-2 functions in survival after the initial generation of the lymphoid compartment. 50 The expression of Bel-2 in murine B lymphocytes was shown to correlate with developmental populations that were shown to be resistant to apoptosis in vitra. High levels of Bel-2 were shown to be present in pre-pro and mature B lymphocytes fiom the BM; early-pro B cells showed lesser, yet significant expression levels (Li et al., 1993; Merino et al., 1994). These and other experiments have indicated that normal in vitra apoptosis and dex-induced apoptosis is greatest in pre B cells (Garvy et al., 1993; Nunez et al., 1990). In transgenic mice and B cell lines with Bel-2 expression induced throughout the developmental stages, the pre B cells were resistant to apoptosis. This indicated that Bel-2 was likely important to normal B cell development and that its varied expression might be important in selection processes. Although some in vivo studies by our lab had looked at B lymphocyte sensitivity in IgM' verses lgM+ B cells (as described earlier), the more defined B lymphocyte populations had not been studied (Garvy et al., 1993). Therefore this thesis addresses the sensitivity the various specific B lymphocyte developmental substages to Cs elevation in viva, to determine if cell populations that express Bel-2 are more resistant to losses. Cytokine-Withdrawal Induced Apoptosis In multicellular organisms each cell depends on signals from its environment to survive. The context in which a cell is maintained provides for the homeostasis and identity of the cell. When a cell is removed from its environment, and functioning properly, it undergoes apoptosis. It has been proposed that each cell contains the necessary components to undergo apoptosis and that extracellular signals are critical for maintaining survival signals that repress the death process (Raff, 1992). This theory was 51 developed by observations that removing cells from their environment or by blocking gene transcription results in apoptosis. A classic model for this has been observed in immune cells, by withdrawal of cytokines that are necessary for survival (Boise et al., 1993; Rebollo et al., 1995; Sohur et al., 2000). This is a commonly observed form of apoptosis termed factor-withdrawal cell death. The recent discovery of IL-4 induced suppression of Bad activity via the P13 Kinase/AKT kinase pathway, indicated that cytokine receptors may function as survival signals by a common mechanism. Recent studies had also shown that both lL-3 and IL-7 could induce translocation of the pro- apoptotic Bax protein from the cytosol to the mitochondrial membrane (Khaled et al., 1999). It has been proposed that the removal of these cytokines in relevant cell types resulted in an intracellular increase in pH that caused a conformational change in Bax, allowing for its membrane binding and ultimately apoptosis. In addition to pro-apoptotic actions following withdrawal, the presence of several cytokines (including IL-7) had also been indicated in the upregulation of the anti-apoptotic Bc1-2 protein (Dancescu et al., 1992; Karawajew et al., 2000). These observations suggested that cytokine signals can result in direct activation of proteins involved in the control of apoptosis, yet the ability of cytokines to protect cells from apoptosis induction from other sources had not been studied in depth. This thesis will explore the potential of lL-7 and SCF to protect normal BM B lymphocytes fiom factor-withdrawal induced apoptosis and Cs-induced apoptosis. Another molecule that appeared to have significant anti-apoptotic function via cytokine signaling was STATS (signal transducers and activators of transcription) (Nosaka et al., 1999; Rosa Santos et al., 2000). In studies using STATS deficient B and T cell lines, the readdition of the molecule resulted in the upregulation of c-myc, Bel-2 and 52 Bcl-x (Lord et al., 2000). STATS activation is promoted by signaling through the janus kinase 3 (J ak3) phosphorylation pathway. The IL-2 common y receptor had been shown to recruit and activate J ak3 upon ligand binding, suggesting that the cytokines that share this domain (IL-2, IL-4, lL-7, IL-9 and IL-1 5) may send this common signal, promoting cell survival. This was also suggested by the lymphoid deficiencies observed in Jak3 knockout mice, since these mice are deficient in B cell lymphopoiesis. (Thomis et al., 1997). It therefore appeared that certain cytokines, especially those sharing the common y chain receptor, could potentially promote survival by inhibiting apoptotic factors and promoting anti-apoptotic factors. Glucocorticoid induced Apoptosis Although Gc-induced apoptosis has been studied for decades, the exact mechanism by which these steroids induce death in cells of the immune system has not been elucidated. This may be because it affects several cell survival and death processes within the cell. The biological effect of Go on immune cells was studied long before apoptosis became an active area of interest. It was observed that Go downregulated metabolism in lymphoid cells. Cells treated with Gc showed decreased glucose uptake, decreased intracellular ATP levels and down regulation in protein and nucleotide synthesis was observed (Montague and Cidlowski, 1995). Although a general decrease in transcription and translation was observed, Gc was also known to enhance the transcription of many genes (reviewed in (Bumstein and Cidlowski, 1989). Experiments that blocked Gc—induced transcription by coaddition of actinomycin D or protein synthesis with cycloheximide, eliminated Gc-induced death in thymocytes (McConkey et 53 al., 1989). This led to the hypothesis that Gc-induced apoptosis caused, at least in part, by inducing what were termed “death genes”. Many efforts were put into identifying and defining genes that were upregulated or downregulated in Gc-induced apoptosis. Further investigation was not able to indicate any one of these genes as sufficient or central to cell death. More recent efforts have been applied to determine how Gc may induce apoptosis by affecting specific signaling pathways. Using mouse thymocytes, studies have indicated that ceramide release, by sphingomyelinase (Smase) activity, may be one way that Go effects death (Cifone et al., 1999). One group used several indirect methods to inhibit Smase activity, thus inhibiting ceramide production and apoptosis. In addition, they showed that blocking phosphatidylinositol phospholipase C (PI-PLC) activity blocked the activation of Smase and ceramide production and apoptosis. These events were shown to be upstream of GcR mediated transcription. This data contrasts other observations indicating that Gc-induced apoptosis is dependent on GcR mediated transcription (McConkey et al., 1989). One transcriptionally regulated pathway that has been suggested in the mediation of Go induced apoptotic effects is the NFKB pathway. It has been shown that downregulation of NFKB activity and one of its gene targets, c-myc, by GcR mediated upregulation of IKB may be another mechanism by which Gc induces apoptosis. Overexpression of NFKB or c-myc in dex treated thymocytes resulted in a decrease in Gc-mediated apoptosis (Wang et al., 1999). Clearly, more research is needed to fully elucidate the mechanism of Gc-induced apoptosis, but it is likely that the mechanism will involve multiple intracellular processes. 54 Summary Over the past two decades the study of apoptosis has shown that cell death is critical in the homeostasis of multicellular organisms. Intensive research efforts were applied to discovering the mechanism of this process. In the currently proposed models, both caspases and mitochondrial dysregulation are the effectors of cell death. Mitochondrial release of cytochrome c can induce the activation of caspases due to death initiated by damage or neglect. Caspase activation is not required for death in this case, but normally occurs and provides the apoptotic phenotype. Death via death receptor ligation can directly initiate the caspase cascade and death, but mitochondrial release of cytochrome c enhances this process. There is also a family of intracellular proteins (the Bel-2 family) that have either pro-apoptotic or anti-apoptotic function. It is believed that they function as regulators of the mitochondrial membrane. Less is known about how cell death is initiated. Factor withdrawal induced death is one form of apoptosis commonly observed. Certain cytokines (such as IL-4) can promote survival by inhibiting the expression and activity of pro-apoptotic proteins. Their withdrawal then allows the activation of pro-apoptotic proteins and results in the downregulation of anti-apoptotic proteins. Glucocorticoid induced apoptosis induces death as seen in neglected or damaged cells. Developing B and T lymphocytes are two cell types that undergo massive apoptosis due to Go exposure. It is thought that this may play a role in regulating negative selection of these cell types during lymphopoiesis. Cmrent evidence suggests that Gc-induced apoptosis is a complex process that likely occurs via mediating transcription of genes directly involved in the death process, by downregulating survival pathways and by inducing other signaling pathways to induce death. Here Cs-induced 55 apoptosis in B lymphocytes will be thoroughly studied in vitra. Additionally, the effect of other factors, such as cytokine signaling by IL-7 and SCF or the effect of stromal cells as a whole, will be investigated for the ability to modify Cs action on the developing B cells. This is a whole new approach to studying Cs-induced death in normal B cells that explores how survival-inducing and death-inducing pathways affect B lymphocyte responses when they are co-initiated. 56 :8 83m “:88th a Boa Emacs: :8 e003 Soc? mo 5:358me —._ unsur— £00 82: .5. even: . . , =00 “ma—2 omega—9.32 :90 m =oU h .“wfiww, . c . . 6 .. ‘_ , 4.4.4-3.... chew: 9:562 :00 MZ .. . A. w neon—Hok— fi . * W 7 Eek—EH 8.58.28?on neufioweum causeway.— =oU m =oU goo MZ a. . . . i. . n v-. . . an“ ...fi 5.0: .l f \l...p s l I... ,,v .. on ...u 4 . . ... v . . o F... «A! l ...... ..u a: .c .o n .. z. .. . l-\‘-s|aa o h. {I a n pl I\ a. list all In \ II . a .. . on inn. . .. I H I.” I. .o . .. l Ia.— ... r». v... . i «5‘. 1 a . .. e... . . I. as... ,.. u. .. .. . .4 .‘.v.... . ....... I I :1 . I . . . .1 . .II . Ho Ah cahoeeegowhoeiaauw A heuauweum E29»: .. 44 saw 3285: 57 080:8 Eofimflgoe Sohoamfib m 38.88 28m NA 03w:— AEEE 25: a: t 32.9 23 m +23 €232 25m 2: a $2.8 2.8 m 2m O Al 365 E8 828 $5.50 see :45 Emma 838 55.8 age 3-5 3-: «New 5—38 8S8 $58 $222 28m 2.2.. {are 2.5 m 2m 3.9:: 42...: 24:: 3-: A5 9.8 83$ SEED sunny $.38 SSE Emcee 38.38 289 05 E £60 3883. 5:5 :oafioamw 383983 m mEQBoBQ Q— 95w:— Aoeem 2: we «828 .35: .836 :90 12:25 329.23 8 :5: 35m 3.5on 59 $85585 288805.80 v; 25E 8.580 888.8: 8m 89.883280 .. a o unencumoweua—axeaezluc: We: on e on Ilu O“ mofiu moan/U 088889389889; w 85883.51: 8388358 ~ ~ \ on” unsfigsubflun ~ N n O 85.89880: \ «run: 0 o8~§e§§~$u ~ u§§euuam~4N \' oeeuoumoweum \ o“ mm: o no”? EON—Wu 88883.. utv< unencueuuésu BeneuuBeuuaaun m. O “U 88.8.2.0 w... 9888805 on A chasmwk siege»? 0 “U / On! 60 A) Receptor Structure Transcription Activating (AF-l) AF-2 1 . . , 777 DNA Binding Ligand Bmdmg Domain Domain B) Receptor Activation @ _"> Nucleus Figure 1.5 The glucocorticoid receptor 61 E TNF .3 CC“ Death Receptor V Membrane (TN FRI) V Death Domain Adaptor Protein E M Procaspase 8 Mitochondria Caspase 8 cytochrome c Cell D ama g e cytochrome c APAF-l 7\:/ Procaspase 9\ Caspase Activation \C:spase\ 9 // Procaspase 3 Caspase 3 Apoptosis Figure 1.6 Apoptotic pathways 62 O Cytokine Cytokine A . Receptor ¢ . ........................................’ —J "OPActive PI3K JO \ P Inactive PI3K 0 ¥ Inactive Akt '\ ~.,. P “m. 19 Active Akt P / Mitochondria Ba‘bOBcI-XL Cell Survival 9 Apoptosis a-» = N0 Survival Promoting Cytokine = Survival Promoting Cytokine Figure 1.7 Cytokine survival via Bad inactivation ()3 CHAPTER 2: THE IN VI V0 EFFECT OF CORTICOSTERONE ON DEVELOPING B LYMPHOCYTES: FROM COMMITTED PROGENIT OR TO MATURITY 64 ABSTRACT Chronic increases in circulating glucocorticoids (Gc), resulting from activation of the stress axis, can cause decreased levels of circulating B lymphocytes and ultimately compromise host defense. Reduced B cell development in the bone marrow (BM) has been indicated as a mechanism contributing to Gc-induced lymphopenia, yet the affect of Go on each B cell developmental substage was not known. Utilizing multipararneter flow cytometry, the effect of the natural Gc, corticosterone (Cs), on B lymphocyte development, from progenitor cells to maturity, was investigated. The results showed that a modest elevation in plasma Cs concentrations, analogous to concentrations seen during physiological stress, caused a significant reduction in the BM B cell population as early as 12 hours; by 24 and 36 hours nearly 50% of the developing B lymphocytes were lost. Thirty-six hours after increased exposure to Cs, 80% of the surviving cells were IgMIgD+ mature B cells. The populations exhibiting the greatest losses were those that had been reported to be undergoing active lg gene rearrangement; the early-pro through the pre B cell stages of development. The earliest committed progenitors, the pre-pro B cells, with Ig genes in the germline configuration, were somewhat resistant to losses, decreasing by only 23% afier 24 hours and 34% after 36 hours. Immature B cells did not show significant losses after 24 hours, but were nearly eliminated by 36 hours. The remaining pro and pre B cells, normally cycling populations, decreased in cells in the S/GZ/M phases of the cell cycle by 74% and 85%, respectively. Therefore, Cs appeared to negatively affect developing BM B lymphocytes by severely depleting all but the earliest progenitors of developing B cells and by preventing their expansion by inhibiting 65 cell cycling. Additionally, analysis of major hematopoietic lineages of the bone marrow indicated that whereas Cs had a negative effect on BM B lymphocytes, it appeared to promote myelopoiesis by expanding the granulocyte compartment. Therefore modest increases in Cs, comparable to concentrations seen during the induction of the neuroendocrine stress axis, can result in the selective reduction of developing B lymphocytes in as little as 12 hours of exposure. Interestingly, after 36 hours the other developing blood cell lineages in the BM were either not affected or actually expanded in response to increased Cs. Thus, it appears that during stress the energy-expensive process of B cell development is downregulated, but perhaps compensated for by an upregulation in the production of cells that provide the first line of immune defense. 66 INTRODUCTION There are important physiological and pharmacological reasons for interest in the effect of glucocorticoids (Gc) on the immune system. The natural steroid hormones, like corticosterone (Cs) or cortisol, are produced and released from the adrenal gland at basal levels and can become elevated in response to stress (Selye, 1947). Zinc deficiency, trauma, burns and neuroendocrine diseases are examples of chronic stress that can cause enhanced production of Ge (DePasquale-Jardieu and F raker, 1980; Raber, 1998). Conversely, synthetic Gc are commonly used at pharmacological concentrations as anti- inflammatory drugs for diseases such as arthritis, asthma and autoimmune diseases. Natural or pharmacological increases in Go concentrations can cause a decline in the number of peripheral B and T cells. This lab has shown that a major cause of this decrease could be the downregulation of B and T cell development in the primary immune tissue (Garvy et al., 1993). Using Cs implants, Garvy et a1, produced concentrations of circulating Cs in mice analogous to those seen during stress, potentially resulting in a compromised host defense. This resulted in a dramatic decrease in thymus weight and a decrease in developing B lymphocytes in the bone marrow. It was also noted that early B lymphocytes, not expressing IgM, were very sensitive to Cs induced losses. However little was known about how Cs affected the various subpopulations of B lymphocytes as they matured from progenitor to maturity. Therefore it was important to determine whether the losses were specific to certain developmental stages or if Cs negatively affected all developing B cells to the same extent. 67 Such studies have been limited due to the complexity associated with the identification of subsets of cells within the B cell lineage, especially since they reside in a heterogeneous tissue, the bone marrow (BM). All cells of the hematopoietic lineage originate and most mature in the BM of adult mammals. One exception are the T lymphocytes, which are generated in the BM, then migrate to the thymus to mature. B lymphocytes develop entirely in the marrow in adults and comprise approximately 30% of murine BM cells. Over the last decade fluorochrome conjugated antibodies against B cell surface proteins has been utilized for the flow cytometric identification of distinct stages in B lymphocyte development. The B lymphocyte lineage is identified by an antibody that recognizes CD45RA (8220) (Coffinan and Weissman, 1981). Using the phenotypic scheme developed by Hardy et al. (Hardy et al., 1991) 8220+ B cells can be further subdivided, from the earliest progenitor to maturity, as follows: pre-pro (S7+HSA“BP1’), early-pro (S7+HSA+BP1'), late-pro (S7+HSA+BP1+), pre (S7'IgM'), immature (IgM+IgD') and mature (IgM+IgD+). The status of Ig gene rearrangement has been determined for each of these stages of development (Faust et al., 1993; Li et al., 1993). Pre-pro B cells have Ig genes in a germline configuration; recombination of the heavy chain begins in early-pro and is completed in late-pro B cells. Pre B cells contain a completed heavy chain and undergo light chain gene rearrangements. At the immature stage of development lgM rearrangement is complete and it is expressed on the surface of the cell. This well developed phenotypic system provides the tools needed to better determine the effect of Ge on cells of the B lineage. During the Ig rearrangement process it is estimated that nearly 80% of developing B cells are lost due to faulty or anti-self recombination events (Osmond, 1986). A 68 downregulation of the anti-apoptotic protein, Bcl-2, has been observed in cells undergoing Ig recombination, suggesting that induction of apoptosis may be a major mechanism of precursor B cell deletion of unwanted cells in the bone marrow (Li et al., 1993; Merino etal., 1994; Nunez et al., 1990). Only the pre-pro and mature B cells express substantial amounts of Bel-2, therefore, it was of interest to determine if increased endogenous production of Cs might adversely affect those stages in B cell development that do not express Bcl-2. This would suggest that apoptosis might play a role in the Gc mediated downregulation of B cell lymphopoiesis seen both in vivo and in vitro (Garvy et al., 1993; Garvy et al., 1993; Merino et al., 1994). Subdermal implantation of a Cs containing pellet can cause increased concentrations of circulating Cs being analogous to concentrations produced during physiological stress. Previous studies from this lab focussed on the effect of Cs on B lymphocyte after days or weeks of exposure. Beyond two days maximum depletion of B lymphocytes was apparent and remained virtually unchanged for two weeks. Thus the studies here were performed at 12, 24 and 36 hours after Cs implantation, to determine the initial effects of increased concentrations of circulating Cs on developing B lymphocytes and the thymus. A comprehensive investigation of the onset of Cs-induced suppression of B cell lymphopoiesis has not been investigated. These experiments demonstrate that the stages during B lymphocyte development where 1g gene rearrangement is occurring, undergo dramatic cell losses afier just a few hours of exposure to steroid. Indeed losses began as early as 12 hours in early-pro and pre B cell populations and maximum effects were seen in these and late-pro and immature populations by 36 hours. In contrast, the populations that had been shown to contain 69 significant levels of Bel-2 (pre-pro and mature B cells) were more resistant to losses due to chronic exposure to Cs. Therefore in the time period of 12 to 36 hours the negative effect of Cs on B cell lymphopoiesis in mice was dramatic. Previous in vitro studies have shown that Gc causes lymphoid cells to undergo cell cycle arrest in 60/61 (Andreau et al., 1998; King and Cidlowski, 1998). Pro and pre B cells normally undergo expansion, whereas the immature and mature B cells are quiescent in the bone marrow. Although a decrease in cycling lgM' precursors after prolonged Cs exposure was previously noted, the initial affect on the individual pro and pre B cells was not investigated. These experiments show a rapid and similar decrease in cycling of surviving pro and pre B cells. Therefore in addition to inducing B cell losses, Cs also rapidly reduced cell proliferation. Interestingly the BM cellularity did not change subsequent to exposure to Cs, yet there was a significant decrease in B lymphocytes. It seemed probable that other cell lineages were expanding during chronic exposure to Cs. Using a recently defined phenotypic labeling system, the cellular composition of key populations of cells of the marrow was determined (de Bruijn et al., 1994). Differential expression of the cell surface proteins ERMP12 (CD31) and ERMP20 (CD59), allows for the delineation of granulocytes, monocytes, hematopoietic progenitors, erythroid progenitors and lymphocytic cells. Interestingly, mice with Cs implants had a significant increase in the percent of granulocytic cells concomitant with a decrease in lymphocytic cells whereas the other lineages appeared unaffected. These observations suggested that Cs, while downregulating B cell development, might promote the generation of granulocytic cells. Therefore a reduction in B cell development can occur within 12 hours of modest 70 physiological increases in Cs. After 24-36 hours B cell development is nearly completely eliminated except for the earliest progenitors, yet the development of the other blood cell lineages is not downregulated and in some cases their development is upregulated. Thus it appears that very early after the induction of the stress axis the development of blood cells shifis fiom lymphopoiesis towards myelopoiesis. 71 MATERIALS AND METHODS Mice and Cs Implantation Young adult male Balbc/J mice were purchased from Jackson Laboratories (Bar Harbor, ME), and were used from 8 to 12 weeks of age. Mice were housed in the laboratory animal facility in the Biochemistry department and all protocols were approved by the University Animal Use Committee at Michigan State University, East Lansing, MI. The facility was maintained at 25°C with 12-hour light and dark cycles. Methoxyflurane inhalation was used to anesthetize the mice and tablets containing 20 mg corticosterone (Sigma, St. Louis, MO) and 20 mg cholesterol were implanted subcutaneously. Sham controls received tablets containing 40 mg of cholesterol. Post- surgery the mice were housed in the animal facility on sterile bedding and fed commercial rodent chow (Purina, St. Louis, MO) and acidified water. Tissue Harvest and Processing At 12, 24 or 36 hours afier tablet implantation, mice were bled under anesthesia and sacrificed by cervical dislocation. The plasma was separated from the blood and Cs concentrations were determined as previously described (DePasquale-Jardieu and Fraker, 1980). Thymuses were removed to determine the amount of thymic atrophy by weight. Bone marrow was flushed from femurs with approximately 1 ml harvest buffer (Hanks’ balanced salts, 1 mM HEPES pH 7.2 and 4% FBS) using a syringe and 22 gauge needle. The marrow was processed into a single cell suspension and red blood cells were removed by lysis. The cells were then washed in harvest buffer, resuspended in 0.5 ml label buffer (Hanks’ balanced salts, 1 mM HEPES pH 7.2, 0.1% sodimn azide and 2% FBS) and placed on ice. Cell counts were performed and trypan blue exclusion was used to determine bone marrow cell number and viability. Following cell counts, two million cells were aliquoted into 5 ml polystyrene tubes (Becton Dickinson, Franklin Lakes, NJ) for phenotypic analysis. lmmunophenotyping and DNA Staining Three separate phenotypic protocols were used to determine the various stages in B lymphocyte development. All antibodies were used at a dilution predetermined to provide optimum labeling. To identify pro, pre and lgM+ cells the following antibodies were used: phycoerythrin (PE) conjugated rat anti-mouse CD45RA (8220), fluorescein (F ITC) conjugated rat anti-mouse CD43 (S7) and biotinylated (biotin) goat anti-mouse lgM F (ab’); (lgM) were added simultaneously to cell aliquots. Antibodies against 8220 and S7 were purchased from Pharmingen (San Diego, CA) and anti-lgM was purchased from Jackson Immunoresearch Laboratories (West Grove, PA). Cells were incubated for thirty minutes, then washed two times with label buffer. The fluorochrome red 670 (R670) conjugated to streptavidin (Av) (Gibco, Grand Island, NY) was added to cells for conjugation to biotin-anti-IgM. Cells were incubated for 20 minutes, washed and fixed with 1 ml of 1.25% paraforrnaldehyde for 40 minutes at room temperature. For DNA staining cells were washed with label buffer two times and resuspended in 0.5 ml 1 ug/ml 4’, 6-diamidino-2-phenylindole (DAPI) (Sigma, St. Louis, MO) and incubated for at least one hour at room temperature. For the identification of immature and mature B cells PE conjugated anti-B220, biotin conjugated anti-lgM and FITC conjugated rat anti-mouse 73 IgD (Pharmingen, San Diego, CA) were added simultaneously. Samples were incubated for 25 minutes, washed and incubation with Av-R670 for 20 minutes. Following phenotyping cells were fixed with paraforrnaldehyde and stained with DAPI, as described above. To identify the pro B cell subsets (pre-pro, early-pro and late-pro) four-color phenotypic analysis was used. The antibodies (all purchased from Pharmingen) were as follows: PE conjugated anti-8220, FITC conjugated anti-S7, biotin rat-anti-mouse CD24 (HSA) and purified rat anti-mouse LY-51 (BP-l, 6C3 clone). Cells were first incubated with anti-BP-l for 30 minutes and washed. An arninomethylcoumarin (AMCA) conjugated goat anti-rat IgG (Jackson lrnmunoresearch Laboratories, West Grove, PA) was added for identification of the BP-l primary antibody. Cells were washed two times and 10 ug of Rat lg was added for 10 minutes to block any unbound anti-IgG antibody. The antibodies against 3220, S7 and HSA were then added at their appropriate dilutions and the cells were incubated for 25 minutes and washed. Av-R67O was then added for conjugation to HSA-biotin. Following labeling and washing, the cells were fixed with 0.5 ml of 1.25% paraformaldehyde and stored at 5°C until flow cytometric analysis. To determine the composition of the major cell lineages within the BM, additional antibodies were used to determine the percent of granulocytes, monocytes, lymphocytes (as a whole), hematopoietic progenitors and erythroid progenitors. For this purpose biotinylated rat anti-mouse CD31 (ERMP12) and F ITC conjugated rat anti-mouse CD59 (ERMP20), (Bachem, King of Prussia, PA) were utilized. Differential surface expression of CD31 and CD59 on the various cell lineages allowed for their identification as 74 follows: granulocytes(ERMP12'ERMP20MW‘3, monocytes (ERMplzmd‘mERMPmb’igh‘), lymphocytes (ERMPlzmwmmERMPZO'), hematopoietic progenitors (ERMP12W‘ERMP20WW) and erythroid progenitors (ERMPIZ‘ERMPZO' ). To label BM cells, ERMP12 and ERMP20 were added simultaneously to 2 X 106 cells on ice at a predetermined dilution. Samples were incubated for 30 minutes then washed with label buffer. Following phenotyping, cells were fixed by the addition of 0.5 ml 1.25% paraforrnaldehyde and stored at 5°C until F ACS analysis Flow Cytometry and Data Analysis Samples were analyzed on a Becton Dickinson FACS Vantage flow cytometer. F ITC, PE and R670 fluorochromes were excited at 488 nm, and emission was detected at 530, 575 and 670, respectively. DAPI and AMCA were excited at 365 nm and emission was detected at 470 nm and 450 nm, respectively. To reduce the spectral overlap of fluorochromes voltage compensation was performed by subtracting non-specific fluorescence fi'om each wavelength analyzed. For the 3-color phenotypic and DNA analysis, debris and cellular aggregates were excluded from analysis by gating using size and DNA content. Cell size was determined by forward and side light scatter and DNA content was determined by a DAPI pulse processed width versus area signal; cells with hypodiploid DNA were considered apoptotic, as previously described (Telford et al., 1991). A region was drawn around 8220+ cells to identify the B lymphocytes of the marrow and depending on the antibody combination used, the 3220+ subpopulations were determined as follows: pro B 75 (S7+IgM'), pre B (S7'IgM’), immature (IgM+IgD‘) and mature (IgM+IgD+) B cells. lsotype-matched samples were used as negative controls. To determine the subpopulations of pro B cells, the 4-color phenotypic samples were analyzed by flow cytometry. Cells were gated based on size to exclude debris and aggregates. To define pro B cells a region was drawn around 82204'87+ cells and then analyzed as follows: pre-pro (HSA'BPI'), early-pro (HSA+BP1') and late-pro (HSA+BP1+). The percent of each population was determined and isotype-matched control samples were used to determine negative fluorescence. Most samples were run in duplicate and averaged. Data was processed and analyzed using PC-LYSIS (Becton Dickinson) and WinList (Verity Software House, Inc., Topsham, ME) software. Microsoft Excel was used for statistical analysis and graphing. Statistics The Student’s t-test was used to determine significant differences (p < 0.05) between experimental and sham mice where mean data :1: the standard deviations were reported. 76 RESULTS Plasma Cs Concentrations, Thymic Weight and BM Cellularity Under conditions of chronic stress in mice the plasma Cs increases 2 to lO-fold above concentrations seen in non-stressed animals. Therefore the concentration of plasma Cs was determined for both sham-control mice and mice receiving a Cs-implant (Figure 2.1). Mice containing the Cs tablet showed increased concentrations of plasma Cs at 12 hours of approximately 95 ung. At 24 and 36 hours serum Cs plateaued between 60 and 70 ug/dl. These concentrations of Cs are consistent with those seen under conditions of stress (Osati-Ashtiani et al., 1998). The slightly higher concentration of Cs, apparent in sham mice at 12 hours, is likely due to some contribution of steroid due to surgical stress. Thymic atrophy is a common observation under conditions of chronic increases in circulating Cs (Garvy et al., 1993). The thymuses of mice that underwent a sham surgery with a cholesterol tablet implant or mice that had a Cs-containing tablet implant were weighed to determine the affects on thymic integrity. Much of the thymus is composed of developing T cells and this lab has shown that a decrease in weight is often correlated with a decrease in thymic cellularity (personal communication, King and Fraker). Figure 2.2 shows the thymic weights for each mouse. As early as 12 hours after surgery, the thymus of mice containing the Cs tablet implant had decreased by 25%. After 24 and 36 hours thymic weight had decreased 54% and 69%, respectively. This is of interest since BM, in contrast, did not decrease in total cell numbers. Table 2.1 shows the total cellularity of mouse femurs 12 and 36 hours after sham or Cs implantation. No 77 significant change in BM cellularity was seen and the absolute cell numbers did not change in the BM of mice exposed to Cs. Since the BM is composed of many different cell lineages, this suggested that the reduction in a population such as B lymphocytes might have been compensated for by an expansion in another cell lineages. Significant Changes in B-Lineage Cells of the Bone Marrow Induced by Cs Exposure Although distinct stages in B lymphocyte development have been defined based on surface expression of proteins, Ig gene rearrangement status and the expression of proteins relevant to B cell lymphopoiesis, the in viva effect of Cs on the various stages from committed progenitor to maturity had not been assessed. Figure 2.3 and Figure 2.4 show the dramatic changes in the composition of cells of the B lineage subsequent to only a few hours of exposure to Cs. Figure 2.3 shows actual flow cytometric profiles of the various B lymphocyte developmental populations 24 hours following implantation of a sham or Cs tablet. Chronic Cs exposure resulted in a decrease of almost 60% in pre B cells and a decrease of about 55% in pro B cells. Further analysis of the pro B cell substages showed that pro B cell response to Cs varied between the populations. Pre-pro B cells were relatively resistant to losses, decreasing by only 17%. Early-pro B cells, in contrast, showed extensive losses, decreasing 70% and late-pro B cells decreased 33%. Analysis of the immature and mature B cell populations indicated that, while immature cells were not substantially different from the negative control, mature cells actually increased to 164% of the proportion noted in sham mice. These profiles are from representative mice and indicate that the pre B cells (where lg light chain rearrangement 78 is initiated) and early-pro B cells (where heavy chain rearrangement is initiated) undergo dramatic losses in less than 24 hours of exposure to somewhat modest increases in Cs that were analogous to those produced by the stress axis. The effect of chronic exposure to Cs on the composition of the B lymphocyte developmental substages over the course of 36 hours is shown in Figure 2.4. The earliest committed B cell progenitors, the pre-pro B cells, were not significantly affected by enhanced exposure to Cs in the first 12 hours. Relative to the other precursor populations the decrease in these cells after 24 or 36 hours of 23% and 34%, respectively, were modest. Early-pro B cells, in contrast, were dramatically decreased by 60% after only 12 hours of heightened Cs exposure. These cells were nearly eliminated, at 24 and 36 hours, decreasing by approximately 80%. The next developmental stage, the late-pro B cells, did not show significant decreases until 24 hours post tablet implant. At 24 a decrease of 62% was observed and they continued to decrease by 72% through 36 hours. The pre B cells, like the early-pro B cells, were significantly decreased after only 12 hours of Cs elevation. They continued to decrease by 70% and 94% after 24 and 36 hours, respectively. Immature B cells were somewhat resistant to losses during early exposure, not showing significant decreases until 36 hours post-implantation. At 36 hours, these cells were nearly eliminated, they decreased fiom approximately 7% of the BM population to less than 1% of the BM. Interestingly, mature B cells were not adversely affected by Cs and as early as 12 hours after heightened exposure to Cs these cells actually increased in the BM. They continued to increase through 36 hours, where experimental mice displayed nearly 3-fold more mature B cells in their BM compared to sham controls. Mature B cells in sham mice made up around 4% of the BM, whereas 79 they made up 12% of the marrow in Cs-treated mice. Clearly, Cs severely affected the development of B lymphocytes in the EM. Early-pro, late-pro, pre and immature B lymphocytes were so dramatically decreased that together they made up less than 3% of the BM as compared to 24% of the BM in sham mice. In contrast the pre-pro B cells were somewhat resistant to losses and mature B cells actually increased dramatically in the BM. Since the total cellularity of the bone marrow of sham and experimental mice did not change these increases represent actual increases in the overall cell numbers for mature B cells in the bone marrow. Therefore Cs exposure can begin to negatively affect B cell lymphopoiesis as early as 12 hours and maximum losses were noted by 36 hours. Cs Induced Changes in Cell Cycle Status of Pro and Pre B Lymphocytes The reduction in the B cell compartment of the marrow was largely due to losses in numbers of developing cells, but a decrease in proliferation of the remaining cells might have also played a role. To investigate this, the cycling status of pro and pre B cells (lgM+ B cells have very few cells in a cycling state) after Cs exposure was determined. For this purpose the BM was phenotyped with DAPI to determine the DNA content. A dramatic decrease in the percent of pro and pre B cells in the S/G2/M phases of the cycle was observed 24 hours after Cs implantation. Figure 2.5 shows both the change in the cell cycle distribution of these cells as compared to sham mice as well as the overall percentage of cells in the S/Gz/M phases for both treatment groups. The average percent of pro B cells in the S/Gz/M phases of the cell cycle in sham controls was 25.5 :1: 2.8% whereas mice exposed to increased Cs had only 6.2 i 0.2% cells in the cycling phases (a 75% decrease). The remaining pre B cells displayed a dramatic 84% 80 decrease in cells in the S/G2/M phases of the cell cycle 24 hours after Cs exposure. Therefore chronic Cs exposure, analogous to that produced by the stress axis, resulted in dramatic decreases in cell cycling in the remaining pro and pre B cells in the BM. This suggested that at 24 hours the remaining cells were not proliferating and could not reconstitute the B cell compartment. Taken together this data shows that chronically elevated Cs can not only cause dramatic decreases in pro and pre B cells, but also substantially reduced the proliferative capacity of the surviving cells during exposure. The Effect of Cs on Major Hematopoietic Cell Lineages of the BM Considering that approximately half of the B lymphocytes in the BM are eliminated due to Cs exposure, an analysis of the composition of the key hematopoietic cell lineages of the BM was performed to determine why the overall BM cellularity was unchanged. Using the ERMP12 and ERMP20 labeling system granulocytes, monocytes, lymphocytes, hematopoietic progenitors and erythroid progenitors could be differentiated (de Bruijn et al., 1994). The composition of the BM of sham mice and mice containing a Cs tablet insert alter 36 hours is shown in Figure 2.6. The lymphocyte compartment that represented 32% of the cells of the marrow of sham mice had been reduced by 40%, to 19% of the BM, in mice with heightened exposure to Cs. Interestingly, the granulocyte population increased by 31%, comprising 42% of the marrow in sham mice and 55% of the marrow in mice with a Cs tablet implant. These results strongly suggested that the overall BM cellularity did not change due to Cs, because the granulocyte population expanded almost proportionately to the decrease in lymphocytes. Thus Cs can have two different effects on hematopoiesis by negatively affecting the generation of B 81 lymphocytes while positively affecting the generation of granulocytes, thereby potentially skewing the immune system towards myeloid defense as lymphoid mediated immune responses are decreased. 82 DISCUSSION The experiments herein showed that increased concentrations of circulating Cs caused substantial losses of developing B lymphocytes and reduced proliferation among surviving cells. The early-pro through pre B cell stages of development, where lg gene rearrangement has been shown to occur, were the most sensitive to Cs-induced losses. These populations are likely intrinsically susceptible to apoptosis, due to the natural elimination of up to 80% of developing B cells that occurs during selection for correct lg gene rearrangement and against faulty rearrangements. These are also the developmental populations where the expression of the anti-apoptotic protein, Bel-2, had been shown to be downregulated (Merino et al., 1994). This suggests that Cs might, at least in part, elicit its negative effect on developing B cells by inducing apoptosis in cells that are sensitive to cell death, due to low levels of Bcl-2. In fact Chapters three and four of this thesis will show in vitro studies where Cs directly induced apoptosis in bone marrow B lymphocytes with pre B cells exhibiting the greatest susceptibility to cell death. The loss of B lymphocyte cellularity also appears to be due to a reduction in cell cycling. These data showed that the remaining pro and pre B cells, normally cycling populations, displayed dramatically reduced cycling after 24 hours of Cs exposure. Considering that newly generated cells would likely be eliminated by apoptosis due to heightened Cs, proliferation of these cells would be a futile process. The data presented later in this thesis will also demonstrate that, in vitro, Cs caused reduced cell cycling in the non-apoptotic pro and pre B cells. It is not known whether the decrease in proliferating cells in vivo was due to a selective elimination of cycling cells and/or as a 83 result of a direct reduction in cycling by Cs. The latter scenario would be congruent with the literature that reports a direct link between Cs and cell cycle downregulation (King and Cidlowski, 1998). The negative effect of Cs on the BM, appeared to be specific for the B lymphocyte lineage. Results here suggest that increased Cs exposure might actually have a positive effect on the generation of other cell lineages, specifically granulocytes. This generates an intriguing hypothesis on how the immune system might respond to stress. Heightened Cs might selectively reduce the highly error-prone B cell development process while promoting granulopoiesis to maintain and enhance a first line of defense to try to compensate for a decreased second line of defense. Clearly, this area of research needs to be studied further to determine the effect of Cs on granulopoiesis for longer time periods and to determine the mechanism whereby Cs enhances the granulocyte population. This might simply be due to a non-responsiveness of granulocytes to Cs, resulting in expansion since loss of B lymphocytes would allow for more physical space in the BM. Expansion may also result if Cs has a proliferative and/or cell survival effect on BM granulocytes. Relatively recently studies have shown that Gc, in contrast to its negative effect on lymphocytes, promotes the survival of neutrophils (Liles et al., 1995). Therefore further investigation of Ge effects on the myeloid lineage is clearly warranted. The experimental system used here has proven to be a reliable and reproducible model for inducing physiological-like increases in serum Cs. Surprisingly little was known of the effects that the induction of the stress axis could have on developing cells of the immune system. Most studies had focussed on the effects of Gc on peripheral, mature immune cells. Here it was shown that the development of B cells is reduced, thus 84 likely contributing to the lymphopenia often seen during chronic stress. Considering the early onset of BM B cell downregulation and the modest increases in Cs, it appears that lymphocyte development is one of the initial immune functions to be negatively affected by the steroid. It therefore seems likely that somehow protecting these cells from heightened Cs exposure might provide immune resilience to stress. With the advent of B lymphopoiesis supporting long term bone marrow cultures and more advanced technology for studying cells of the marrow, it has become clear that stromal cells can play a major role in both lymphopoiesis and myelopoiesis by the production of critical cytokines (Whitlock and Witte, 1982). It is also clear that the addition of Gc to long term cultures, where stromal cells are present, selectively promotes the development of the myelopoietic lineage. Considering the results presented herein, it might be that Cs modifies the production of key molecules produced by stromal cells to support a more myelopoietic-like environment. It may also be that the exogenous addition of lymphopoiesis supporting factors could override Cs negative effects on B cell development. Interleukin-7 is one cytokine produced by stromal cells that can act as a critical proliferative, survival and differentiation signal in early stages of B cell development (Namen et al., 1988). It might be that this cytokine could potentially act as an immunotheraputic agent to protect B cell development against increases in Ge. Therefore studies investigating the effect of IL-7 and other factors, critical to lymphopoiesis, on Cs- induced downregulation of B cells could have important implications in the treatment of stress-related immunodeficiencies. Later chapters in this thesis investigate the potential of stromal cells and/or exogenous lL-7 to modulate the negative effects of Cs on B 85 lymphocytes and also investigate the direct effect of Cs on the production of two key cytokines involved in blood cell development, IL-7 and SCF. Further understanding of the effect of Go on immune cell development could provide important advances in health studies, since both physiological stress and pharmacological administration of Go is widespread. In vivo studies to determine the effect of Cs on the development of the myeloid lineage is warranted, as are studies investigating possible immunotheraputic agents on B lymphocyte development. Additionally, in vitro experiments could provide insight into the mechanism whereby Gc downregulates B cell development and how various factors might modulate that mechanism. Clearly the work presented here has provided some key insight into the effect of Go on blood cell development and likely could provide a foundation for further investigations. 86 Figure 2.1 The plasma Cs levels for mice receiving a 40 mg cholesterol implant (sham) (solid bars) or a 20 mg Cs plus 20 mg cholesterol implant (dotted bars) at 12, 24 and 36 hours are shown. The data are expressed as ug/dl and standard deviations are shown (n=4). Significant differences between control and experimental mice were established using the Student’s t-test (p < 0.05) and are indicated (*). 87 1m m m m m o €33 20228380 mo 36 hours S 3CW mo, hm” 48 2." S I, w u 0 h. on A L A l L Figure 2.1 Plasma corticosterone concentrations 88 Figure 2.2 The individual thymic weights for sham control mice (X) and mice receiving a Cs implants (O) for 12, 24 and 36 hours are shown. Six mice were analyzed per group where some individual data points are hard to distinguish due to overlap. The mean of each group is indicated and an asterisk indicates significant differences between sham control and experimental mice as determined by the Student’s t-test (p < 0.05). 89 35.4 ’5. 3% X E sot X r .. r x X ‘9 j 9 x X g 20 ; Us: 0 .2 ~ E, 1511 8 §* 10 : +5 i o g* 5 1 0-; ____ z" 0 m 12 Hour 24 Hour 36 Hour X Sham 0 Cs Figure 2.2 The effect of Cs or cholesterol implants on thymic weights 90 Table 2.1 Bone Marrow cellularity of Cs or sham implanted mice ”Implant " mores more . , Time/Tm _, , , ' , aren‘t“) , 12 Hour Sham 2.58 i 0.19 X 107 12 Hour Cs 2.75 i 0.46 X 107 36 Hour Sham 2.20 i 0.29 x 107 36 Hour Cs 2.32 i 0.46 X 107 ' There was no significant difference between sham or experimental mice and data shown are mean :1: standard deviation where n = 6 mice per treatment group. 9] Figure 2.3 F low cytometric data for pre-pro through mature B lymphocytes where representative results from a sham control mouse and a Cs exposed mouse 24 hours post- implantation are shown. Panel (A) shows the 8220+ gated cells and gives the percent of pro (B220+S7'IgM’), pre (B220+S7'IgM') and lgM“ (B220+S7‘IgM+) cells in the BM of sham controls and Cs-treated mice. Regions were drawn around each population and the cell type and percent are shown. Panel (B) shows the B220+ gated cells and the percentage of immature (B220+IgM+IgD') and mature (B220+IgM+IgD‘) B cells in the BM. Panel (C) shows BZZO+S7+ cells and the percent of pre-pro (B220+S7+HSA'BP'1'), early-pro (3220*S7+HSA+BP-l') and late-pro (3220*s7+HSA*BP-1*) B cells. Percentages were based on total nucleated BM cells. Data is representative of at least six mice per treatment group. 92 Sham Cs A 77.7 ‘7 7 i Immature . ' ‘ +Mature 8.9% C l‘ 0 °F 2 £9 | Pro 2.9% S7-FITC B M ature ‘ M ature IgD-FITC i C Late-Pro ‘ i Late-Pro \ anrly-Pro , 1.2% Early-Pro . _ - 0.8% : . 1,132. j '..." :54 HSA-R670 BPl-AMCA Figure 2.3 Flow cytometric profile of developing bone marrow B lymphocytes 93 Figure 2.4 The percentage of pre-pro, early-pro, late-pro, pre, immature and mature B cells in the BM are shown for sham (solid lines) and Cs-implanted mice (dashed line) at 12, 24 and 36 hours after implantation. Six mice were averaged per treatment group and the mean i standard deviations are shown. 94 c: on b: «N .1 NF @. h: on hI «N h: N . l . l r l n ...... n. I I I I ID =mO m 0.5.33 ..h ...... e “m”. h =8 m A68 2:5: cat .5>o 338339 ofioofifib m 38.88 25p 3 mowawno 326:va «.N 0.53..— w -1e 3. .fi 3 s f 3 _ 3 H. 3 2.8.9:. «o . .D . . Eusmlol h_._ on b_.._ «N .1 NF r. _ . ad . - 3 - °.« . 9.0 - ad . r 0.2. :00 m SBaEE. .1 on a: «N .1 Nw . ad D ..... U. . .. . . no 0 oé ”r m... e ad _ m4. e c6 :8 m oi->..em h: on .1 «N 5: NF inn»... Ill! !. e... U. ad {4%. 2: m 3.: i 98 0....“ :00 m 05n— i 8 i «N i a. lilLfii 1i. .- I'll; coo . «d . ed . ed , ed - a . .Do . r O.—. . . - u; e 3 e 3 :00 m Gully—Q mouew auoa lo % mouew auoa #0 °/. 95 Figure 2.5 The cell cycle distribution for pro and pre B cells of sham and Cs-treated mice at 24 hours was determined. The main panel shows the change in the percentage of cells found in the S/GZ/M phases of the cell cycle for pro and pre B cells. The actual percentages of cells in S/szM are shown in the panel insert. The data are for six mice from both sham and Cs-implanted mice where the mean i stande deviations are shown and significance is indicated with an asterisk. 96 % Cells in S/Gle 25% « I 20% ~' i 15% J l 10% .. i 5% i . 0% Pro Pre 'ISlanDCs Figure 2.5 Cs-induced cell cycle changes in pro and pre B cells 97 Figure 2.6 The cellular composition of the BM from sham and Cs-treated mice after 36 hours is shown. Percentages were determined by flow cytometric analysis of the phenotypic distribution of granulocyte, lymphocyte, hematopoietic progenitors, monocyte and erythroid progenitors. These results are based on the expression of the ERMP12 and ERMP20 cell surface antigens. The means for sham control mice (n=5) and Cs- implanted mice (n=3) are shown. 98 l Sham Implant l 36 Hours/Bone Marrow Cellularity Monocyte j 5% Erythrocyte 16% Progenltor 5% ‘ Lymphocyte 32% I Granulocyte 42% 'Cortfééfeifié Implaintffi 36 Hours/Bone Marrow Cellularity ‘ Margo/cm Erythrocyte ° 16% 1 Lymphocyte i 1 9% Granulocyte 55% Progenltor 4% Figure 2.6 Cs effect on major hematopoietic lineages in the bone marrow 99 CHAPTER 3: STROMAL CELL PROTECTION AGAINST CORTICOSTERONE-INDUCED APOPTOSIS IN BONE MARROW B LYMPHOCYTES 100 ABSTRACT Glucocorticoids (Gc) can induce apoptosis and cell cycle arrest in developing bone marrow (BM) B cells both in vitro and in viva. Stromal cells are critical to B lymphocyte development in the BM and in long-term BM cultures. They can provide support to the developing B cells through direct cell-cell interactions and by the production of lymphopoietic cytokines. Here it is shown that soluble factors produced by and direct interactions with stromal cells can protect BM B lymphocytes from short-term exposure to a natural Gc, corticosterone (Cs). Indeed, stromal cells could reduce the ability of the steroid to initiate apoptosis by up to 30-50% among pro, pre and IgM+ B cells. However, neither soluble factors nor direct interactions with the stromal cells restored cell cycling in developing B cells exposed to Cs. However, the exogenous addition of the lymphopoiesis promoting cytokine, interleukin-7 (IL-7), augmented the stromal cell protection against Cs-induced apoptosis in pro and pre B cells and also restored cycling to normal levels among pro B cells. In addition the potential of Ge to directly modulate the expression of genes important in hematopoiesis was investigated. Stromal cell expression of IL-7 and stem cell factor (SCF), a cytokine that supports early blood cell progenitors from many lineages, was therefore determined. After one day, Cs caused SCF mRN A to increase 2-3 fold, but it did not have an affect on IL-7 expression. Therefore the direct affect of Cs on BM B lymphocytes can be influenced by other cells normally found in the BM microenvironment, namely stromal cells. Additionally Cs appeared to also directly effect stromal cells by causing the upregulation of a general hematopoiesis supporting cytokine, SCF. Taken together these experiments demonstrate 101 the complex nature of studying BM B cells, since cells and factors that would normally be present in vivo can modulate the responses of developing cells presented with a death inducing molecule such as Cs. Thus, some of the negative effects of Gc on B cell development during physiological stress could potentially be alleviated by factors produced by stromal cells and/or exogenous IL—7. 102 INTRODUCTION The successful generation of mature B lymphocytes is a complex multi-stage process consisting of differentiation and proliferation. It is accepted that during B cell development apoptosis plays an important role by eliminating the high population of precursor cells with faulty immunoglobulin (lg) gene rearrangements. Detection of apoptosis in vivo has been very difficult due to the rapid phagocytosis of dying cells and the difficulties associated with studying subsets of cells in heterogeneous tissues like the bone marrow. Short-term in vitro culture systems and flow cytometry have allowed for the phenotypic analysis of the various stages in B cell development to include DNA analysis for cell cycle status and apoptosis (Garvy et al., 1993; Hardy et al., 1991; Telford et al., 1994). Our lab and others have shown that, in vitro, precursor B cells normally undergo low but significant levels of apoptosis in short-term cultures (less than 24 hours) (Garvy et al., 1993; Merino et al., 1994). Additionally, glucocorticoids (Gc), such as dexamethasone or corticosterone (Cs), have been shown to cause a dramatic increase in apoptosis among these cells. Therefore in vitro models have proven valuable in studies that otherwise could not be performed in vivo. A major concern when studying B lymphocyte apoptosis and cell cycle in vitro is that in vivo B lymphocytes normally develop in close association with fibroblast-like cells, termed stromal cells, that support their development (Osmond, 1990). In the mid- 1980’s Whittlock and Whitte developed a long-term B lymphopoiesis supporting culture system (Whitlock and Witte, 1982). Stromal cells were key to the successful long-term maintenance of lymphopoiesis in this culture system. Later a variety of stromal cell lines 103 were generated that could help support B cell lymphopoiesis for long periods of time (Collins and Dorshkind, 1987; Dorshkind et al., 1986). Without stromal cells, lymphocytes do not survive beyond a few days (Borghesi et al., 1997). Interestingly, with the modification of culture conditions B lymphopoietic long-term bone marrow cultures and stromal cell lines could be converted from the promotion of B cell lymphopoiesis to the production of granulocytes (Dorshkind et al., 1986). Myelopoiesis supporting cultures were first described by Dexter (Dexter et al., 1977) and vary from lymphopoietic cultures in that exogenous Go is added to the media. This suggested that the presence of Ge aided in changing the environment from the support of lymphocytes to the support of granulocytes in long-term cultures. Therefore Cs, in a period of weeks, could dramatically modify the culture environment. However, few studies have investigated the more immediate effect of stromal cells on lymphoid cells treated with an apoptosis inducing factor such as Cs. One study by Borghesi, et al showed that a stromal cell line, BMSZ, could protect precursor B lymphocytes (IgM') from dexamethasone induced apoptosis, but effects on the specific pro, pre and lgM+ populations was not investigated. In addition to directly inducing apoptosis in developing B cells, it may be that Cs directly modulates stromal cell gene transcription. Many effects of Ge on their target cells are mediated by transcription modulation of certain genes via Gc receptor binding to glucocorticoid response elements. Under B lymphopoiesis supporting conditions, stromal cells produce certain cytokines, such as interleukin 7 (IL-7) and stem cell factor (SCF) that promote B cell development. Exogenous IL-7 can promote the proliferation and survival of pro and pre B cells without stromal cell support (N amen et al., 1988; 104 Namen et al., 1988). Elimination of IL-7 in culture resulted in the loss of B cells. Moreover it was determined that IL-7 deficient transgenic mice do not produce B cells beyond the pro-B cell stage (von Freeden-Jeffry et al., 1995). SCF, in combination with IL-7, promoted commitment to the B lineage and enhanced proliferation of the pro B cell population (McNiece et al., 1991). Whereas IL-7 was shown to be specific and critical for lymphocyte development, SCF has been shown to promote the development of progenitors of several cell lineages (Ashman, 1999). Therefore it might be that IL-7 plays an important role in Whitlock-type cultures whereas SCF supports both B lymphopoietic and myelopoietic Dexter-type cultures. Since the presence of CC is the main difference between these culture systems it is possible that the steroid can modulate these cytokines that are involved blood cell development. The stromal cell lines 810 and $17 and stromal-like cells sorted from long term bone marrow cultures were used to determine the effect of stromal cells on Cs-induced apoptosis and cell cycle status in pro, pre and IgM+ B cells. Additionally, the effect of 810 and 817 stromal cell derived soluble factors, alone, on Cs-induced apoptosis was determined by culturing the BM B cells suspended over stroma with a membrane insert. These soluble factors modestly protected each B lymphocyte population fiom Cs-induced apoptosis while direct interactions with the stromal cells afforded slightly more protection. Although stromal cells demonstrated some protection against Cs-induced apoptosis they did not protect against the Cs-induced cell cycle arrest. The exogenous addition of IL-7 augmented stromal cell protection against Cs-induced apoptosis in pro and pre B cells and restored cycling among pro B cells. Additionally, the direct effect of Cs on the expression of SCF and IL-7 mRNA was determined. In the time frame studied 105 here, Cs caused a significant increase in SCF mRNA levels; the cytokine that promotes both lymphoid and myeloid development. Interestingly, the cytokine that is specific for lymphoid development, IL-7, did not increase. Therefore Cs might promote myelopoietic cultures by upregulating cytokines involved in this process. 106 MATERIALS AND METHODS Harvesting of Bone Marrow from Mice Young adult male Balb c/J mice were purchased from Jackson Laboratories (Bar Harbor, ME) and housed in the animal facility in the Biochemistry department at Michigan State University, East Lansing, MI; protocols were approved by the All- University Committee on Animal Use and Care. Mice were used from 6 to 14 weeks of age and marrow was flushed from femurs with approximately 1 ml harvest buffer per bone (Hanks’ balanced salts, 1 mM HEPES pH 7.2 and 4% F BS) using a 22 gauge needle. The marrow was processed into a single cell suspension and red blood cells were removed by lysis. Cell Culture/Cell Lines Bone marrow derived murine stromal cell lines 310 and S l 7 were generous gifts from the laboratory of Dr. Kenneth Dorshkind. These lines were grown at 37°C with 7.5% C02 and maintained in RPMI 1640 containing 5 x 10'5 M 2-mercaptoethanol, 1 mM HEPES pH 7.2, 1000 units/ml penicillin, 0.1 mg/ml streptomycin, 2 mM glutamine and 5% fetal bovine serum (PBS). PBS was purchased from HyClone (Logan, UT) and was tested for optimal performance; the same lot was used throughout. $10, $17 or sorted stromal cells were plated at 1x105 cells/ml in 24-well (lml/well) or 12-well (2ml/well) Corning Costar (Corning, NY) tissue culture plates and were incubated overnight to insure complete adherence. To obtain stromal cells sorted from long-term bone marrow cultures, cultures were initiated and maintained as previously described (Whitlock and 107 Witte, 1982) and VCAM-1+ (Southern Biotechnology Associates, Inc., Birmingham, AL) cells were sorted via fluorescence activated cell sorting (F ACS). FACS sorted cells were maintained under the conditions used for the stromal cell lines; the sorted VCAM-1+ cells had typical stromal cell morphology and were MAC—1'(Pharmingen, San Diego, CA). Bone marrow was plated at 1-2 x 106 cells/well either in 24-well plates in media only, in 24 well plates directly onto confluent stroma or onto 0.4 uM transwell inserts suspended over stroma in 12-well plates. Corticosterone (Cs), purchased from Sigma (St. Louis, MO), and/or recombinant murine interleukin 7 (IL-7), purchased fiom R&D Systems (Minneapolis, MN), were added to cultures at 0.1 11M Cs and at 0.1 ng/ml IL-7. BM cells were dislodged from culture by the addition of 0.02% EDTA. lmmunophenotyping/DNA staining Antibodies to B cell surface antigens were added to samples at a predetermined dilution to phenotypically label different stages involved in B cell lymphopoiesis. phycoerythrin (PE) conjugated anti-CD45R (B220), fluorescein (F lTC) conjugated anti- CD43 (S7) and biotinylated anti-lgM F(ab’)2 (lgM) were added to samples simultaneously and incubated for 25 minutes. Antibodies against S7 and B220 were purchased from Pharmingen (San Diego, CA) and anti-lgM was purchased fiom Jackson Immunoresearch Laboratories (West Grove, PA). Following primary staining, streptavidin-red670 (R670), purchased from Gibco (Grand Island, NY), was added at a predetermined dilution for conjugation to biotinylated anti-lgM. Samples were incubated with R670 for 20 minutes and then washed two times with label buffer. Following phenotyping, cells were suspended in 50% FBS and fixed with the slow addition 1.2 ml 108 of ice cold 70% ethanol. To determine DNA content for cell cycle and apoptosis, samples were stained with 0.5 to 1.0 ml of 4’, 6-diamidino-2-phenylindole (DAPI) (Sigma, St. Louis, M0) for at least one hour prior to flow cytometric analysis. Flow Cytometry Samples were analyzed on a Becton Dickinson FACS Vantage flow cytometer. F ITC, PE and R670 fluorochromes were excited at 488 nm, and emission was detected at 530, 575 and 670, respectively. DAPI was excited at 365 nm and emission was detected at 470 nm. Debris and cellular aggregates were excluded from analysis by gating based on size and DNA content. Cell size was determined by forward and side light scatter. DNA content was determined by a DAPI fluorescence and cells with hypodiploid DNA were considered apoptotic, as previously described (Telford et al., 1994). 8220' B lymphocyte subsets were defined as follows: pro B cells were S7+IgM', pre B cells were S7'IgM' and IgM B cell were 87'. Flow cytometric data was processed using either PC-LYSIS (Becton Dickinson) or WinList (Verity Software House, Inc.) software Ribonuclease protection assay RNA was isolated fi'om 810, 817 or sorted stromal cells using Pharrningen’s Total RNA Isolation Kit and protocol (San Diego, CA). In general 5 X 106 to 1 X 107 cells were used per isolation yielding approximately 80-150 ug total RNA. RNA was resuspended in 0.6 ml RNase-free water and stored at -80°C until analysis. 109 The IL-7 and SCF probe templates were purchased from Pharmingen. Probe synthesis was performed as described in Pharmingen’s Riboquant Multi-Probe RNase Protection Assay (RPA) System. Briefly, the probe template was transcribed with 32F labeled uridine triphosphate (New England Biolabs). The radiolabeled probe was then added to 50 ug stromal cell RNA and hybridized overnight. For positive control samples 10 pg of RNA from IL-7 overexpressing cell lines Psi 5-20 and N59 and the SCF overexpressing BHK-MKI cell line were used. Control lines were very generous gifts from the laboratory of Dr. Richard Schwartz. Negative control samples contained 2 ug of tRN A. RNA samples and radiolabeled probe were hybridized overnight at 56°C. Following hybridization, ribonuclease treatment was performed as described by Pharmingen. Samples were electrophoresed on a 5% denaturing polyacrylamide gel and visualized by phosphorimaging. To obtain a standard curve for fragment mobility, the probe was run at 2000 counts per minute per lane of the gel. The housekeeping genes glyceraldehyde 3-phosphate dehydrogenase (GAPDH) and L32 were used to normalize variances in the amount of sample applied to the gels. Phosphorimaging was used for quantitation and Excel was used to analyze the data. Data Analysis and Statistics Data analysis was performed using Microsoft Excel software and mean i standard errors are reported. Statistical significance was determined using the Student’s t-test (p < 0.05). 110 RESULTS Pro, Pre and lgM+ B Lymphocytes Undergo Spontaneous and Cs-Induced Apoptosis To determine the effect of stromal cells on spontaneous verses Cs-induced apoptosis it was important to assess the degree of apoptosis in each cell type. To choose an appropriate culture time frame where significant levels of apoptosis were induced but where necrosis or cell losses were minimal, the degree of spontaneous and Cs-induced apoptosis at 15, 24, and 48 hours was determined. Figure 3.1 shows a time course of apoptotic pro, pre and lgM+ B lymphocytes incubated in media alone (Panel A) or with 0.1 M Cs (Panel B). The data show that over time, the percent of spontaneous apoptosis increased in each developmental population from 8%, 14% and 9% in pro, pre and lgM cells at 15 hours to 29%, 36% and 23% in the pro, pre and lgM+ populations at 48 hours, respectively. Freshly isolated marrow contained fewer than 2% apoptotic B lymphocytes (data not shown). The addition of Cs to culture caused a significant increase in apoptosis. At 15 hours Cs caused a modest induction of apoptosis in pro, pre and IgM+ of 24%, 44% and 36%. Greater levels of death were apparent at 24 hours; 32%, 64% and 48% in the pro, pre and IgM+ cells, respectively. Afier 48 hours nearly all of the pre and IgM+ cells were apoptotic. Therefore a 15 hour culture period resulted in a significant induction of spontaneous and Cs-induced apoptosis, yet these levels were more modest compared to later time points, thus making the modulation of B cell death by stromal cells more apparent. lll Protective Effect of Stromal Cell Lines on Spontaneous and Cs-Induced Apoptosis in B Lymphocytes Pro, pre and IgM+ B lymphocytes underwent low levels of spontaneous apoptosis at 15 hours and the addition of Cs caused increased apoptosis, as shown above. Therefore, whether stromal cells and/or the soluble factors they produce could modulate this apoptosis in the B cell populations was determined. BM cells were cultured as a single cell suspension either in media alone, directly on confluent stromal cells or suspended above the stromal cells using a 0.4 11M permeable membrane insert. Samples were incubated for 15-16 hours after which the composition of cells of the B lineage and degree of apoptosis were determined by flow cytometry. Figure 3.2 (A) shows the effect of either soluble factors from and/or direct interactions of BM B lymphocytes with the 817 stromal cell line on spontaneous apoptosis. Direct incubation with $17 resulted in a significant decrease in spontaneous apoptosis of 40-50% in each pro, pre and lgM+ B cells. However soluble factors, alone, did not reduce spontaneous apoptosis in any stage of development. The 810 stromal cell line also had similar effects on the BM B cells (data not shown). Therefore, direct interaction with the stromal cell lines, but not soluble support alone, could reduce spontaneous apoptosis in pro, pre and IgM B cells. Figure 3.2 (B) shows the effect of soluble factors from and/or direct interactions with S17 on Cs-induced apoptosis in pro, pre and lgM+ B cells. Direct interaction with 817 resulted in the greatest protection against Cs-induced apoptosis in pro B cells (44% reduction). lgM+ cells were also significantly protected with a 39% decrease and pre B cells showed a more modest 29% decrease. Soluble factors produced by S17 also protected B lymphocytes from Cs-induced apoptosis. A 31%, 20% and 28% reduction in 112 apoptosis by soluble factors alone was observed in pro, pre and IgM+ cells, respectively. Although this trend suggested lesser protection by soluble factors, the reduction was not statistically different from those obtained from direct stromal contact. S 10 stromal cells had similar protective effects on the B lymphocytes (data will be presented below). The data suggests that soluble factors produced by these stromal cell lines provided significant protection to each stage of B lymphocyte development from Cs-induced cell death in short-term cultures. Interestingly, soluble factors from the stromal cell lines did not appear to protect B lymphocytes from spontaneous apoptosis, suggesting that either one or more soluble molecules produced by stromal cells could specifically protect against Cs-induced apoptosis without affecting basal levels of apoptosis. Stromal Cell Effect on Cs-Induced Cell Cycle Arrest in Pro and Pre B Lymphocytes Data here and in Chapter two showed that along with causing apoptosis in B lymphocytes, Cs also caused a reduction in cycling pro and pre B cells. The extent to which these two Gc-induced phenomenon are interrelated is not fully understood. Here experiments were performed to determine the effect of S17 on Cs-induced cell cycle arrest. BM B lymphocytes were incubated with or without S17 as previously described and in media alone. Table 3.1 shows the percent of pro, pre or lgM+ cells in the S/Gle phases of the cell cycle. Samples incubated without Cs showed that neither soluble nor direct interaction with S17 changed the proportion of cycling cells in control cultures. Addition of 0.1 M Cs, on the other hand, resulted in a significant decrease in the percent of cycling pro and pre B cells similar to the decreases seen in in viva studies. Cells in S/Gz/M decreased by 54% in pro B and 71% in pre B cells when cultured with Cs. 113 Neither soluble factors nor direct interaction with S17 were able to restore the cycling population in the pro or pre B cells. S10 similarly did not effect the cell cycle status and these results will be shown below. Therefore, although stromal cells could protect B lymphocytes from Cs-induced apoptosis they were not able to protect the developing pro and pre cells from the Cs-induced cycling decrease. The Effect of Exogenous IL-7 on BM B Lymphocytes Cultured with Stromal Cells Chapter 4 in this thesis shows the effect of exogenous addition of IL-7 alone on B lymphocytes exposed to Cs. The cytokine protected pro B cells from Cs-induced apoptosis and cell cycle arrest and also modestly protected pre B cells fiom the Cs- induced apoptosis, but not the cell cycle arrest. It was therefore important to determine if the exogenous addition of IL—7 to cells incubated with either S10 or S l 7 could reproduce the effect on B lymphocytes seen with IL-7 alone. Figure 3.3 shows actual flow cytometric data for apoptosis and S/Gz/M in pro, pre and IgM cells cultured with no treatment, 0.1 uM Cs, Cs and $10 or Cs, S10 and IL—7 for 16 hours. As compared to background levels, Cs clearly induced substantial induction of apoptosis in pro, pre and IgM cells. Additionally significant reductions in pro and pre B cell cycling were apparent. Direct incubation with S10 resulted in a 50%, 51% and 55% decrease in Cs- induced apoptosis in pro, pre and lgM+ cells, respectively. The exogenous addition of IL-7 further reduced the percent of apoptotic pro B cells by an additional 62%, bringing the percent apoptosis to below background levels. IL-7 caused a more modest decrease in apoptosis in pre B cells of 43% as compared to pre B cells cultured with S10 alone. The cytokine did not cause a further reduction in apoptosis in lgM cells. These data are 114 consistent with the independent effect of IL-7 on these cell types. IL-7 addition to B lymphocytes cultured with S 17 resulted in similar responses to those seen with S10 (data not shown). As seen with the S l 7 data above, S 10 alone did not affect the decrease in pro and pre B cells in S/G2/M induced by treatment with Cs. The addition of IL-7 to cells incubated with S 1 0, on the other hand, resulted in a restoration in pro B cell cycling to that seen in background cultures. Again the effect of exogenous IL-7 addition to these cultures reproduced the effect of IL-7 alone on pro B cell cycling. Therefore recombinant IL-7 has effects on Cs-induced cycle arrest that S10 alone could not provide and S10 alone appears to protect lgM cells from Cs-induced apoptosis whereas IL-7 did not augment this response. The Effect of Cs on Stromal Cell Expression of SCF and IL-7 The major mechanism whereby Cs elicits its effects in cells is by transcription regulation via the Ge receptor. Although the protective effect of 810 and 817 on B lymphocytes could be attributable to soluble factors other than IL-7, it was important to determine whether this cytokine was expressed in each stromal cell type and whether that expression was modified by Cs-treatment. Additionally the cytokine SCF could be a target for regulation by Cs, since it generally promotes hematopoietic progenitors. In contrast to IL-7, SCF likely could play a role in long term myelopoietic cultures, where Cs is part of the culture medium. Figure 3.4 shows the relative RNA expression of IL-7 and SCF in 817, S10, sorted stroma and from positive control cell lines. Clearly message for both cytokines was present in each stromal cell type, with sorted stromal cells 115 expressing higher levels of IL-7. Relative to IL-7, SCF was expressed at dramatically higher levels. Figure 3.5 (A) shows the change in SCF mRNA levels in 810, S17 and sorted stromal cells treated with 0.1 M or 10 1.1M Cs for 20-24 hours. Cs treatment (10 1.1M) resulted in a 2-3 fold elevation in SCF mRNA expression in all stromal cell types tested. Although the increase caused by 0.1 uM Cs was less overall it was not statistically different fi'om the increases induced by 10 11M treatment. These data show that Cs caused an upregulation of SCF as early as one day after treatment. IL-7 mRNA was detectable in SI 7, S10 and sorted stroma. $17 and 810 showed very low but similar expression levels. Sorted stroma, in contrast, had approximately 5- 10 fold higher production of IL-7 mRNA (as shown in Figure 3.4). Figure 3.5 (B) shows that there was no significant change in IL-7 mRN A expression with Cs treatment. Therefore, Cs does not appear to regulate IL-7 expression after 20-24 hours exposure to Cs, suggesting that the modest protection provided by stromal cells was not due to upregulated IL—7 expression. 116 DISCUSSION The experiments presented here have shown that in short-term primary bone marrow cultures significant but low levels of spontaneous apoptosis occurs in pro, pre and IgM+ B cells. The addition of Cs at concentrations analogous to physiological rather than pharmacological concentrations resulted in substantial increases in apoptosis, with pre B cells being the most susceptible to death. Additionally the steroid dramatically reduced cycling pro and pre B cell numbers. Soluble factors produced by, or direct interactions with, the $10 or S 17 stromal cell lines caused a modest reduction in apoptosis in each B cell population, but in contrast the stromal cells did not affect the Cs- induced reduction in cell cycling. This suggests that factors produced by stromal cells could potentially protect developing B cells from Cs-induced apoptosis over a short period of time. At least one soluble factor produced by the stroma appeared to be responsible for the decrease in Cs-induced apoptosis. IL—7 would be surmised to be one likely candidate for this protective function, since it is critical and specific for lymphopoiesis rather than myelopoiesis in the bone marrow (Namen et al., 1988; von Freeden-Jeffry et al., 1995). The S 1 7 cell line had been reported to not produce biologically active levels of this cytokine but here these cells protected the B lymphocytes to the same extent as $10, which has been reported to produce biologically active IL-7 (Henderson et al., 1990). However, 8 1 7 was also reported to not produce IL—7 mRN A by RT-PCR but the experiments herein show that S17 in fact displayed low levels of IL-7 gene expression being comparable to that seen in S 10 (Faust et al., 1993). Therefore these 817 stromal 117 cells did express IL-7 mRN A in contrast to previous publications. The difference could potentially arise from a reversion of the cell line to an IL-7 producing stromal cell. Both 810 and S17 were derived from multiple passages of long term bone marrow cultures to obtain a self-renewing clonal cell type, therefore there is no enforced selection of cells to maintain characteristics such as the lack of expression of certain cytokines. The independent protective effect of exogenous IL—7 on B lymphocytes is thoroughly presented in Chapter 4 of this dissertation. The cytokine clearly protected pro and pre B cells from Cs-induced apoptosis and eliminated the cell cycle arrest in pro B cells. Those experiments suggested that 1L-7 could potentially have immunotheraputic value in lymphocyte development during stress. Here the exogenous co-addition of this cytokine to stromal cell supported cultures, reproduced the protective effects seen by the cytokine alone in addition to stromal cell protection. Therefore the immunotheraputic potential of IL-7 was further supported by these experiments. Regardless of the potential effects that low levels of lL-7 produced by stromal cells might have, it seems likely that other soluble factors produced by the stromal cells have a protective effect on B cells exposed to Cs. Evidence for this comes from the protection fiom Cs-induced apoptosis observed in IgM+ cells. IL-7 alone or in combination with stromal cells did not protect this more mature cell population and previous studies have indicated that lgM+ cells in the marrow are mostly unresponsive to IL-7 (Sudo et al., 1993). One stromal cell derived cytokine that could potentially elicit this effect is thymic stromal lymphopoietin (TSLP). Recent studies on the positive effect of TSLP on developing B lymphocytes indicated that this cytokine could potentially support B cells from Cs—induced apoptosis, because it supports B cell 118 development throughout I gM+ stages (Levin et al., 1999). Further studies to identify active factors that can protect all stages of B lymphocyte development from Cs-induced apoptosis would therefore be of value. The studies on the expression of SCF message levels have shown that Cs can directly affect bone marrow stromal cells and potentially can modulate the hematopoietic environment by transcriptionally regulating cytokines important to blood cell development. Cs treatment for one day significantly upregulated SCF gene expression. Since SCF is a cytokine that positively affects a variety of progenitor blood cell types and that Cs promotes myelopoiesis, it may be that the upregulation of SCF acts to promote the initiation of a myelopoietic-like environment. IL-7 on the other hand, which does not promote myelopoiesis, was not upregulated in response to Cs. Since long-term exposure to Cs causes the stromal cell environment to promote myelopoiesis it is logical that the steroid would cause the upregulation of a general cytokine involved in blood cell development, but would not upregulate a cytokine specific for lymphopoiesis which the environment does not support. It could be that after extended culture with Cs, stromal cells might actually downregulate IL—7 expression. Advances in technology have created the potential to screen the expression of a variety of cytokine genes simultaneously and Cs treatment of stromal cells could be a valuable area of investigation. It is clear that the study of B lymphocyte responses to physiological factors is a very complex area of study with several variables to take into account when studying cells in vitra. Nevertheless, the complete understanding of B cell development and how that development can be modified remains a very important area in health and disease. 119 Therefore it is important to study these cells and attempt to understand the potential roles of factors that would normally be found in the BM on them. 120 Figure 3.1 The percent of apoptotic cells in pro, pre and lgM+ B lymphocytes from BM are shown. Apoptosis was determined after culturing for 15, 24 or 48 hours, by DNA staining and analysis for hypodiploid DNA. Panel A shows apoptosis among cells cultured in media only and Panel B shows the percent apoptosis in the B cell populations cultured with 0.1 uM Cs. Data are the average of duplicate samples i the standard deviations and is representative of several experiments. Error bars not seen are smaller than the data symbols. 121 ,-s from 811 by DNA cells populations ldard '6 smaller 50% " 40% 30% ‘ Apoptosis 20% *' 10% * 0% .L - . _.__ 10096. 80% 60% l Apoptosis 40% 20% 0% 7‘ Figure 3.1 Spontaneous and Cs-induced apoptosis over time in pro, pre and IgM+ B cells in vitro. Cs Treatment Time (Hours) 122 Figure 3.2 Apoptosis was measured among pro, pre and IgM+ B cells with and without Cs to determine potential protective effects of stromal cells (direct contact) or factors produced by stromal cells (membrane insert). BM was processed into a single cell suspension and plated in media alone (solid bars), on a 0.4 uM transwell insert suspended over S l 7 stromal cells (dots) or directly onto a confluent layer of S 1 7 stromal cells (stripes). Following 15-16 hours in culture the BM cells were harvested and phenotyped for pro, pre and lgM+ B cells, fixed and stained with DAPI, a DNA dye, for analysis of apoptosis. Panel (A) shows the change in apoptosis in cells cultured alone or with or without direct stromal cell contact in media and (B) shows the change in apoptosis among cells exposed to 0.1 M Cs alone or when in direct or indirect contact with stromal cells. All data were normalized to the percent of apoptosis among cells cultured without S17 and are the averages of at least three experiments. The actual percent apoptosis for pro, pre and IgM cells in control cultures without Cs or stromal cells were 7%, 14% and 12%, respectively. Control samples treated with Cs were 21%, 51% and 41% apoptotic in pro, pre and IgM+ cells, respectively. Standard error bars are shown and significant differences from control samples was determined using the Student’s t-test (p < 0.05) and indicated by an asterisk. 123 wt 1 A No Treatment i 140% “ 120% ‘ 1 100% * 1 ' 80% - s * l 60% l 40% a j 20% l 0% , Pro Pre IgM+ l I No s17 [:1 Soluble s17 IDirect S17 Apoptosrs 1 B Cs Treatment ‘ 120%? 1 100% . s * é a 80% i * * * * l 1. ‘ a 1 40% j 1 20% 1 l ‘ ‘ i 0% . , 7, , Pro Pre IgM+ I No 817 El Soluble Sl7 IDirect 817 l l. _ hm, , _ Figure 3.2 Stromal cell modulation of spontaneous and Cs-induced apoptosis 124 Table 3.] Bone marrow B cells in S/Gz/M phases of the cell cycle Cell Type . N o Stroma Soluble Factors‘l Direct Contact” 1"" 16.6 i- 8.5%* 153 i 6.8%“ 13.1 i- 4.6% c N” C’ P“ 12.6 i 1.2% 10.6 i 1.9% 11.9 i 1.8% lgM 4.8 3: 2.0% 4.8 i 0.8% 4.2 i 1.7% Pro 7.7 i 0.5% 8.0 i 1.2% 6.25 i 2.1% 0.1 M Pre 3.6 :1: 2.0% 3.8 i 1.5% 3.4 i 1.0% Cs IgM+ 3.6 i 1.4% 3.6 i 03% 3.3 i 1.1% * Data are presented as the averages of three experiments i standard deviations a B lymphocytes cultured suspended over S 1 7 stromal cells by a 0.4 11M membrane insert b B lymphocytes cultured directly on S17 stromal cells c There were no statistical differences between cells incubated in media alone, with soluble factors from the S17 stromal cell line or with direct contact with the stromal cells 125 Figure 3.3 Flow cytometric DNA histograms show the Cs-induced change in apoptosis and cell cycle status for pro, pre and IgM B cells and modulation of these Cs-induced changes by stromal cells or a combination of stromal cells and IL-7. BM was processed into a single cell suspension and incubated for 16 hours in culture media only (A), with 0.1 11M Cs (B), directly on a confluent S10 stromal cell layer and with 0.1 uM Cs (C) or with S l 0, Cs and 0.1 ng/ml IL-7 (D). Following culture BM cells were harvested and phenotyped for pro, pre and IgM+ cell types, the samples were fixed with ethanol and the DNA stained with DAPI. Cells were analyzed by flow cytometry as follows: pro B (B220+S7+IgM’), pre B (B220+S7'lgM') and IgM cells (B220+S7'IgM+). Debris and cell aggregates were excluded from analysis by gating based on cell size and each population was analyzed for apoptosis (A0) and cell cycling (S/Gz/M) as indicated. These data are representative of several experiments. 126 Relative Cell Number DNA Content Figure 3.3 Flow cytometry of 810 and/or lL-7 effects on apoptosis and cycling of pro B cells 127 \ Figure 3.4 The relative RNA expression between stromal cell types is shown as the actual phosphoirnage obtained from the ribonuclease protection assay. RNA was extracted and the ribonuclease protection assay were performed as described by Pharmingen using 50 1.1g stromal cell RNA and 10 pg positive control cell line RNA. The samples are shown as follows: Lane (A) unprotected probe, (B) S17 stromal cells, (C) 810 stromal cells, (D) sorted stromal cells, (E) IL-7 positive control cell line N59, (F) SCF positive control cell line BHK-MKI. The approximate electrophoretic mobility for each is indicated on the left hand side of the figure. A B C D E F IL-7 Probe—p g IL-7 —> SCF Probe—-p w SCF—p L32 Probe/v GAPDH Probe—p L32 —D GAPDH Figure 3.4 Relative expression of IL-7 and SCF message 129 Figure 3.5 The relative change in messenger RNA expression levels of SCF (A) and IL- 7 (B) for stromal cells treated with 0.1 11M or 10 11M Cs for 20-24 hours. Stromal cells tested were the S10 and S17 cell lines and stromal-like cells sorted from long term bone marrow cultures. Cells were sorted based on large size and VCAM-1 surface expression. These cells had fibroblast like morphology typical of stromal cells and did not express MAC-l on their surface. The stromal cells were grown just to confluence and spent media was replaced with fresh media for approximately one day before Cs addition. Following culture, whole cell RNA was extracted from approximately 5 X 106 - 1 X 107 stromal cells using Pharmingen’s Total Cell RNA Isolation Kit. RNA analysis was performed using the ribonuclease protection assay and quantified using phosphorimaging. The results shown are the average changes relative to the corresponding non-treated cells for at least two experiments and standard deviation bars are shown. Asterisks represent significant differences as determined by the Student’s t-test (p < 0.05) from untreated samples. 130 ’ A SCF gene expression 1 440% \ 300% T 250% : 20096 < 15096 i 100% 1 50% 1' 0% 1 RNA Expresssion l INoCs EJ'0.1|.1M'Cs I110uMCs 1L7 gene expression 25096 1 20096 ; 15096 ~ 100% *1 RNA Expresssion 50% 2 096 ‘ INo Cs 110.1 11M Cs I110 11M Cs Figure 3.5 Cs effect on stromal cell expression of SCF and IL-7 RNA 131 CHAPTER 4: INTERLEUKIN-7 PROTECTS EARLY STAGES IN B LYMPHOYCTE DEVELOPMENT FROM CORTICOSTERONE-INDUCED APOPTOSIS AND CELL CYCLE ARREST 132 ABSTRACT Independent and combined effects of the cytokines interleukin-7 (IL-7) and stem cell factor (SCF) on spontaneous verses glucocorticoid (Gc) induced apoptosis in developing B lymphocytes was investigated. These cytokines could potentially protect developing B cells from the negative effects of cultming and Ge exposure, since they act as positive factors during the early stages of normal B cell lymphopoiesis. Murine bone marrow was cultured as a single cell suspension, without stromal cell support, and the various B lymphocyte developmental stages, levels of apoptosis and cell cycle status were determined by flow cytometry. Exogenous addition of IL-7 reduced spontaneous apoptosis in the pro B cell population by nearly half, but had little effect on the survival of the more mature pre and IgM+ B cells. This cytokine also dramatically protected the pro B cell population from apoptosis induced by the natural murine Gc, corticosterone (Cs), decreasing levels of apoptosis by 60%. IL-7 also afforded modest protection to pre B cells from Cs-induced apoptosis but only minimal protection was provided to lgM bearing cells. Cs exposure resulted in a profound decrease in cycling among pro and pre B cells but once again IL-7 protected pro B cells by eliminating this cell cycle arrest while providing little cell cycle protection to pre B cells. SCF, which can augment lymphopoietic development, similarly caused a 40% decrease in background apoptosis in pro B cells when added alone to short-term cultures. However, in contrast to IL-7, it did not appear to have a significant affect on apoptosis induced by Cs. When SCF was used in combination with IL-7, it did not augment the protection against Cs-induced apoptosis in pro B cells. Therefore although both cytokines could protect pro B lymphocytes fi'om 133 Spontaneous apoptosis, only IL-7 protected early B lymphocytes from Cs-induced apoptosis. Additionally, I L-7 also provided substantial protection of pro B cells from Cs- induced cell cycle arrest. Clearly this cytokine has the potential to dramatically protect early stages in B cell development from agents that alter proliferation or induce apoptosis. It reinforces the importance of IL-7 to the early phases of lymphopoiesis and indicates it has immunotheraputic potential. 134 INTRODUCTION Chronic increases in circulating levels of glucocorticoids (Gc) are evident during physiological stresses such as in burn victims, trauma patients, the aged and the malnourished and can cause impaired immune function and lymphopenia (Schleirner et al., 1989). In vitra, dexamethasone (a synthetic Gc) has been shown to induce apoptosis and decrease cycling in developing lymphocytes in the marrow, with precursor cells (IgM') showing greater sensitivity to these negative effects (Borghesi et al., 1997; Garvy et al., 1993; Merino et al., 1994). Our lab has also demonstrated, in viva, that chronically elevating corticosterone (Cs), resulted in a selective reduction in early B lymphocytes that had not begun to express surface irmnunoglobulin (Garvy et al., 1993). Therefore Cs might elicit its lymphopenic effect in viva, at least in part, by inducing apoptosis in the early stages of B cell development. Recent data also showed that Cs could dramatically reduce cycling among both pro and pre B cells, suggesting this as another mechanism whereby the steroid might downregulate B cell development (Laakko and Fraker, unpublished observation; King and Cidlowski, 1998). Blocking Gc-induced cell death and cell cycle reduction in B lymphocytes could potentially help protect against the lymphopenia that can occur under conditions of stress. Cytokines that are critical to B lymphocyte development might be of benefit in protecting these cells from Gc induced apoptosis. IL-7 and SCF are cytokines that could potentially protect early stages in B lymphocyte development from apoptosis. IL-7 was originally identified as a factor produced by stromal cells in long term bone marrow cultures that, independently, promoted proliferation of precursor B cells in vitra (N amen et al., 1988). 135 When added to B lymphocyte cultures, with no other apparent cytokine or stromal cell support, I L-7 promoted the survival of pro B cells that would otherwise rapidly undergo apoptosis in culture (Gibson et al., 1996; Lee et al., 1989). In addition, IL-7 caused dramatic increases in cycling in early B cell progenitors, not expressing surface lgM. Recent evidence suggested that IL-7 might promote survival via phosphatidyl inositol 3 kinase (PI3K) activation of the AKT pathway and/or by upregulation of the anti-apoptotic Bel-2 protein (Lavagna-Sevenier et al., 1998; Lu et al., 1999). Another stromal cell derived cytokine, SCF, also can promote B cell lymphopoiesis. In combination with IL-7 it promoted the commitment of progenitor cells to the B lineage and it increased IL-7 induced proliferation among pro B cells (McNiece etal., 1991). In addition to amplifying lymphopoiesis, SCF had also been shown to positively affect many progenitors from various lineages when in combination with growth factors specific for those lineages (Ashman, 1999). PI3K mediated activation of the AKT cell survival pathway has also been indicated in SCF signaling via its receptor c-kit (Blurne-Jensen et al., 1998). Although both IL-7 and SCF have shown positive effects on precursor B cell development and survival, what effects these cytokines might have on apoptosis among early B cells has not been determined. Moreover the effect of these cytokines on developing B lymphocytes exposed to a potent death-inducing factor, such as Cs, are not known. The potential of IL-7 or SCF to suppress spontaneous or Cs-induced apoptosis in a short-term (16 hour) culture system was determined here. Murine bone marrow was used as the source for developing B lymphocytes and the degree of apoptosis and cell cycle status was determined using flow cytometry. An established phenotypic scheme 136 using fluorochrome conjugated monoclonal antibodies to cell surface proteins, developed by Hardy et a1 (Hardy et al., 1991), was used to identify three stages in B lymphocyte development. The stages are listed in order of least to most mature as follows: pro B cells, pre B cells and IgM+ cells consisting of immature and mature B cells. The degree of apoptosis and cell cycle status was also determined for each population by measuring the DNA content. The results show that IL—7 and SCF individually provided substantial protection against background apoptosis among pro B cells prior to the induction of cell proliferation, but there was no appreciable protection provided by either cytokine to pre or IgM+ cells undergoing spontaneous death. Interestingly, IL-7 and SCF did not have the same effect on Cs-induced apoptosis. IL-7 protected pro B cells, and to a lesser extent pre B cells, from Cs-induced apoptosis. In the pro B cell population IL—7 also inhibited the cell cycle arrest induced by Cs, but this was not the case for the pre B cell population. SCF did not appear to have a significant affect on Cs-induced apoptosis either alone or in combination with IL-7. Therefore, both of these cytokines appear to promote survival of cells early in B lymphocyte development but only IL-7 can protect early B cells from Cs-induced apoptosis. This is the first direct evidence that a cytokine may be able to help protect early stages in B cell lymphopoiesis from Cs exposure. 137 MATERIALS AND METHODS Bone Marrow Preparation and Cell Culture Young adult male Balbc/J mice were purchased from Jackson Laboratories (Bar Harbor, ME) and housed in a temperature and light controlled facility. All protocols used herein were approved by the All-University Committee on Animal Use and Care at Michigan State University. Bone marrow was flushed from femurs with approximately 1 ml of harvest buffer (Hanks’ balanced salts, 1 mM HEPES pH 7.2 and 4% F BS) per bone using a 22 gauge needle. The marrow was processed into a single cell suspension and red blood cells were removed by lysis. Bone marrow cells (1-2 x 106 cells/ml) were cultured at 37°C in 7.5% C02 using RPMI 1640 containing 5 x 10'5 M 2-mercaptoethanol, 1 mM HEPES pH 7.2, 1000 units/ml penicillin, 0.1 mg/ml streptomycin, 2 mM glutamine and 5% fetal bovine serum (F BS). FBS was purchased from HyClone (Logan, UT) and the same lot of serum was used throughout. Corticosterone was purchased from Sigma (St. Louis, MO) and both recombinant murine IL-7 (rrnIL-7) and recombinant murine SCF (rmSCF) were purchased from R&D Systems (Minneapolis, MN). Cs was used at 0.1 11M, unless otherwise stated. Both SCF and IL-7 were titered for maximum activity and used at 100 ng/ml and 0.1 ng/ml, respectively, unless otherwise stated. The standard culture time was 15 to 16 hours. lmmunophenotyping/DNA staining Antibodies to B cell surface antigens were added to samples at optimum concentrations and samples were maintained at approximately 5°C throughout 138 phenotyping. Cells were incubated for 25 minutes, simultaneously, with phycoerythrin (PE) conjugated anti-CD45R (8220), fluorescein (F ITC) conjugated anti-CD43 (S7) and biotinylated anti-lgM F (ab’)2 (IgM) in label buffer (HBSS containing 2% FBS, lmM HEPES and 0.1% sodium azide). Antibodies against S7 and B220 were purchased from Pharmingen (San Diego, CA) and anti-lgM was purchased fi'om Jackson Immunoresearch Laboratories (West Grove, PA). Following primary staining, Streptavidin-Red670 (R67 0), purchased from Gibco (Grand Island, NY), was added for conjugation to biotinylated anti-lgM for 20 minutes and then washed two times with label buffer. Following phenotyping, cell pellets were resuspended in 50% FBS and fixed by slow addition of 1.2 ml of ice cold 70% ethanol with gentle mixing. Cells were lefi in the fixative for at least one hour and for as long as three days. At least one hour before FACS analysis, fixed samples were incubated at room temperature in 0.5-1.0 ml of DAPI staining solution (1 [lg/m1 DAPI and 0.01 mM EDTA in phosphate buffered saline) for DNA analysis of apoptosis and cell cycle. Flow Cytometry Samples were analyzed on a Becton Dickinson FACS Vantage flow cytometer. F ITC, PE and R670 fluorochromes were excited at 488 nm, and emission was detected at 530, 575 and 670 nm, respectively. DAPI was excited at 365 nm and emission was detected at 470 nm. Debris and cellular aggregates were excluded fiom analysis by gating based on size and DNA content and fluorochrome conjugated isotype antibodies were used as 139 negative controls. Cell size was determined by forward and sideward light scatter and DNA content was determined by DAPI width versus area. According to the phenotypic scheme developed by Hardy (Hardy et al., 1991), B lymphocyte subsets were defined as follows: pro B cells were B220+S7+IgM', pre B cells were 8220+S7'lgM' and immature and mature cells were defined as B220+IgM. Each population was gated and analyzed for apoptosis (hypodiploid DNA) and for cell cycle status. PClysis (Becton Dickinson) or WinList (Verity Software House, Inc., Topsham, ME) sofiware were used to process flow cytometric data. Statistics Microsoft Excel software was used to analyze data and perform statistics. Data are presented as the mean i the standard deviation and significance was established using the Student’s t-test (p < 0.05). 140 RESULTS Analysis of the Effect of IL-7 and Cs on Short-Term Marrow Cultures Cytokine signals are often required for the survival of cells of the immune system. Early stages in B cell lymphopoiesis depend on IL-7 for proliferation, differentiation and long-term survival (Lee et al., 1989; Mertsching etal., 1996; Sudo et al., 1989; Valenzona et al., 1998). Here a short-term culture system where stromal cells and their effects were negligible was used to directly analyze the effect of IL-7 on developing B cells. For this type of culture system it was also important to use a rather short time period where necrosis and cell losses due to culturing and Cs addition were minimal. This was to ensure accurate comparisons between treatment groups and to determine IL-7 effects independent of its proliferative potential. Thus the majority of studies were performed in 16 hours. Table 4.1 shows the recovery and the viability of nucleated cells from bone marrow after 16 hours in culture with media alone, 0.1 ng/ml lL-7, 0.1 uM Cs or IL-7 in combination with Cs. Microscopic cell counts were performed to determine recovery and viability by trypan blue exclusion; apoptotic cells exclude trypan blue and were considered viable in cell counts. The viability at the initiation of cultures was approximately 90% (data not shown). At the end of a typical 16 hour period there was excellent viability in control cultures (93%) with some expansion of cells (126% recovery). Cells incubated with Cs alone had only a slight reduction in viability, decreasing from 90% viable at culture initiation to 82% post-culture. Therefore under these culture conditions there was a modest expansion in B cells and the viability did not 141 differ by more than 10% of the viability of freshly plated cells. The effect of IL-7, alone, on recovery and viability was analogous to untreated cultures, however it provided good protection to Cs-treated cells keeping proliferation and survival to levels analogous to media alone. Protective Effects of IL-7 on Spontaneous and Cs-Induced Apoptosis Among Pro B Cells To determine the specific effect of IL-7 on the survival of cells of the B lineage, the marrow was phenotyped and DNA analysis of the pro, pre and lgM+ cells was assessed by flow cytometry. Figure 4.1 shows that the addition of IL-7 resulted in a 40% decrease in spontaneous apoptosis among pro B cells (Panel A). However, this protection from apoptosis was not evident in either the pre B cell or lgM+ cell populations. Panel B of Figure 4.1 shows IL-7 also provided dramatic protection from Cs-induced apoptosis causing a 60% decrease in apoptosis among pro B lymphocytes. The IL-7 reduction in Cs-induced apoptosis in pre and IgM B lymphocytes was much less than that seen in pro B cells; there was a 30% reduction in pre B cell apoptosis and only a 15% decrease in lgM)r cell apoptosis. Therefore, IL-7 protected pro B lymphocytes from both spontaneous apoptosis and Cs-induced apoptosis dining a 16 hour culture period. Interestingly, although IL-7 had been shown to act as a growth factor for pre B cells (Mertsching et al., 1996; Sudo et al., 1993), it did not appear to inhibit the background apoptosis that occurred when these cells were incubated in culture media alone, but did modestly protect the pre B cells from Cs—induced apoptosis. Therefore the mechanism of IL-7 protection varies between protection from spontaneous apoptosis 142 verses protection against Cs-induced apoptosis, since the cytokine reduced Cs-induced apoptosis without affecting spontaneous apoptosis. The Ability of IL-7 to Maintain Cell Cycling of B cells in the Presence of Cs While IL-7 can enhance proliferation in early B lymphocytes, Cs can induce cell cycle arrest in Go/Gl in some cell types, especially early B cells (King and Cidlowski, 1998). Since the above data showed that IL-7 dramatically protected pro B cells, and to a lesser extent pre B cells, from Cs-induced apoptosis it was important to determine if it could protect against changes in cell cycle induced by Cs. Data in Figure 4.2 shows the change in cycling status of pro and pre B cells when IL—7 was added alone, Cs was added alone or Cs and IL-7 were added in combination to short-term BM B lymphocyte cultures. Data was normalized to control levels of pro and pre B cells in the S/G2/M phases of the cell cycle cultured in media alone. During 16 hours of culture, the addition of IL-7 did not enhance the number of pro or pre B cells in the S/Gz/M phases of the cell cycle. A significant increase in IL-7 induced cycling in pro and pre B cells was not observed until 24 hours after culture; by 48 hours there was nearly a 4-fold increase in pro B cells in S/Gz/M and a 2.7-fold increase in pre B cells in S/G2/M in cells treated with IL-7 (data not shown). The latter showed that the IL-7 used was biologically active on precursor B cells. Addition of Cs to cultures resulted in a substantial decrease in pro and pre B cells in the S/Gz/M phases of the cell cycle. Cells in a cycling phase were decreased by nearly 50% in pro B cells and by 57% in pre B cells. Interestingly, when IL-7 and Cs were added to cultures simultaneously, IL-7 was able to completely override the cell cycle arrest induced by Cs in pro B cells, maintaining the level of cycling to that 143 ”A. seen in media or IL-7 only treated cultures. By contrast, IL-7 had no effect on the Cs- induced decrease in cycling observed in pre B cells. These results show that IL-7 totally protected pro B cells from the Cs-induced cell cycle arrest. In contrast, IL-7 did not inhibit the cell cycle arrest induced by Cs in pre B cells, suggesting a variable effect by the cytokine on changes in cycling in pro and pre B cells caused by Cs. DNA Histograms of IL-7 Protected Pro B cells The above data indicate that IL-7 elicited its greatest effect on the earliest committed B cell progenitors, the pro B cell, both by protecting the cells fi'om spontaneous and Cs-induced apoptosis and by eliminating the Cs-induced cell cycle arrest. Figure 4.3 shows the DNA distribution for representative pro B cells (B220+S7+lgM‘) as determined by flow cytometry. The percent of apoptotic cells (hypodiploid DNA content) and cells in the S/Gz/M phases of the cell cycle (hyperdiploid DNA content) are indicated for each treatment group. Following 16 hours in culture in media alone, pro B lymphocytes underwent low levels of spontaneous apoptosis (less than 10% of pro B cells) and maintained a significant percentage of cells in the S/G2/M phases of the cell cycle (18.5% of pro B cells). The addition of 1 ng/ml IL-7 resulted in a reduction in background apoptosis of approximately 40%, but it did not dramatically affect the percent of pro B cells in S/G2/M phases. Therefore, IL-7 induced survival does not appear to be due to any enhancement of proliferation at 16 hours. As discussed, the addition of Cs at 0.1 11M to short-term BM cultures resulted in an increase in apoptosis in pro B cells from 10% in background cultures to 25% apoptosis. Cs also dramatically reduced the percentage of pro B cells in the S/G2/M 144 phases of the cell cycle by more than 50%. Whether Cs resulted in selective induction of apoptosis in cycling cells or induced an arrest of cycling cells in Go/Gl was not determined here, but previous studies have shown that Cs induces cell cycle arrest prior to induction of apoptosis. However as before, the addition of IL-7 to Cs-treated cultures resulted in a dramatic decrease in the percent of apoptotic pro B cells (Figure 4.3 D). In fact, the protection from Cs-induced apoptosis in this cell type appeared to be complete, since IL-7 reduced levels of apoptosis to less than levels seen in background cultures (Figure 4.3 A). Interestingly, the percent of cycling pro B cells was unchanged by IL-7 at the 16 hour time point but as before IL—7 offset the reduction in cycling created by Cs in pro B cells (Figure 4.3 D). Therefore, IL-7 appeared to inhibit both the cell cycle arrest and apoptosis induction by Cs in the pro B cell population after 16 hours in culture. Continued Protection of Pro B cells by IL-7 at Higher Concentrations of Cs The culture conditions presented in the above experiments were used to ascertain the potential of IL-7 to protect against the effects of Cs on developing B lymphocytes at concentrations analogous to that observed during physiological stress. In the next set of experiments higher concentrations of Cs, some that were analogous to pharmacological levels, were used to determine whether IL-7 continued to protect B cells against apoptosis while maintaining cell cycle status. Phenotypic distribution, degree of apoptosis, and cell cycle status were determined following 15 hours exposure to 0.1 11M, 1 11M or 10 11M Cs (Figure 4.4). At 1 11M Cs, apoptosis appeared to plateau and levels at 10 11M Cs were 32%, 73% and 46% apoptosis for pro, pre and lgM+ B cells, respectively. IL-7 addition to cultures containing 0.1 11M or 1 uM Cs resulted in similar protection against apoptosis 145 among the B cell subsets. Apoptosis among pro B cells was decreased by approximately 60% with the addition of IL-7 at either 0.1 11M or 1 11M Cs. Pre and lgM+ B cells treated with 0.1 1.1M or 1 11M Cs exhibited IL-7 induced decreases in apoptosis of 35% and 19%, respectively. The degree of protection afforded by IL-7 to cells treated with 10 11M Cs for pro and pre B-lymphocytes was not as extensive as for lower concentrations of steroid. IL-7 caused a 43% reduction in pro B cell apoptosis and a 27% decrease in pre B cell apoptosis of cell treated with 10 11M Cs. However 10 11M is a pharmacological level of this steroid. Consistent with the previous cell cycle data presented, IL-7 overrode the decrease in pro B cell cycling caused by Cs at each concentration. However pre B cycling, as before was substantially reduced by Cs and remained unprotected by the addition of IL—7 (data not shown). These results show that IL-7, even at high concentrations of Cs, maintained its ability to substantially protect pro B lymphocytes from apoptosis and decreases in cell cycling. The Effect of SCF on Background and Cs-Induced Apoptosis in BM B Lymphocytes Since SCF also has beneficial effects on early B cell development, its effects on background and Cs-induced apoptosis in B lymphocytes were also investigated. Figure 4.5 shows the effect of SCF on background and Cs-induced apoptosis in pro, pre and IgM+ cells. SCF was added to cultures at predetermined concentration of 100 ng/ml and Cs was used at a concentration of 0.1 M. The addition of SCF to cultures, resulted in a decrease in spontaneous pro B cell apoptosis of 40% but pre and IgM bearing cells were not significantly protected. In contrast to the dramatic effect of SCF on background apoptosis in pro B cells, only a small protective effect was seen on Cs-induced apoptosis 146 in these cells. In fact, the decrease of 17% in pro B cell apoptosis might be attributed to SCF protection from basal apoptosis as described above. Additionally, pretreatment of cells with SCF for 20 hours did not result in decreased apoptosis beyond that presented here (data not shown). Therefore, although SCF appeared to protect pro B cells from basal apoptosis in culture it did not necessarily protect these cells fi'om Cs-induced apoptosis, in contrast to the effect of IL-7. This suggests that SCF might promote in vitra survival, similar to IL-7, but it does not protect these cells from the negative effect of Cs. Since SCF acts synergistically with IL-7 to promote lymphopoiesis, it might synergize with IL-7 to protect pro B cells from Cs-induced apoptosis. To determine this, SCF together with IL-7 was added to cultures and their effects on apoptosis in pro B cells analyzed. Table 4.2 shows the percent of apoptosis in pro B cells cultured alone or with SCF, IL-7 and Cs in various combinations. IL-7 was also added at more dilute concentrations to obtain a lower level of protection from Cs-induced apoptosis, since at 0.1 ng/ml it reduced apoptosis to background levels in pro B cells therefore any effect by SCF may not have been apparent. The data indicated that SCF did not appear to synergize with IL-7 to protect pro B cells from Cs-induced apoptosis. Additionally the combination of SCF and IL-7 did not result in more protection from background apoptosis. Pretreatment with SCF for 20 hours also did not provide enhancement of the protective effect by IL—7 (data not shown). Clearly, although both SCF and IL-7 promote the survival of pro B cells in culture these cytokines do not synergize in the protection of be. these cells fi'om Cs-induced apoptosis. 147 DISCUSSION The survival of hematopoietic cells often depends on signals sent from cytokines. The limiting amounts of cytokines present in short-term marrow cultures can result in enhanced cell death. Here we have shown that both SCF and IL-7 appear to act as survival factors specifically for the pro B cell stage during B cell development by reducing the degree of background or spontaneous apoptosis in these cells afier approximately 16 hours in culture. This reduction in spontaneous apoptosis preceded the enhancement of proliferation by either cytokine, which did not occur until 24 hours. This suggests separate mechanisms for cell cycle control and cell survival. The promotion of pro B cell survival in culture did not necessarily mean that either IL-7 or SCF could protect these cells from the induction of apoptosis by Cs. Here it was shown that IL-7 dramatically reduced Cs-induced cell death in pro B lymphocytes and it also inhibited the cell cycle arrest caused by the steroid. Therefore IL-7 appears to block Cs signaling of apoptosis and/or to promote cell survival signals that overrides Cs death promoting and cycle blocking activity. Even though IL-7 prevented the cycle block it did not cause increased proliferation, above background levels, until after 24 hours in cultm'e, suggesting that protection from apoptosis occurred earlier than the enhancement of cell cycling. Gc-induced apoptosis has been studied for many years, yet its exact mechanism of eliciting death has not been established. It is becoming clear that Gc signaling through the GcR can affect many other intracellular signaling pathways. For example, Gc has been shown to negatively affect the nuclear factor kappa B (N FKB) and AP-l 148 transcription factors, both of which can promote survival and proliferation of lymphocytes (Auphan et al., 1995; Heck etal., 1994). Additionally, some evidence exists that Gc may promote apoptosis by the activation of phosphatidyl inositol dependent phospholipase C and subsequent activation of sphingomyelinase and cerarrride release (Cifone et al., 1999). It is apparent that much more investigation will be needed to determine the exact mechanism of Gc-induced cell death and it is likely that different cell types may use different pathways to induce death. The recently described AKT pathway, which promotes survival by causing the cytosolic sequestration of the pro-apoptotic protein Bad, may be a viable mechanism for the survival induced by both lL-7 and SCF. The receptors for both IL—7 (IL-7R) and SCF (c-kit) have been previously reported to have phosphatidyl inositol 3-kinase (PI3K) activity upon ligand binding (Blume-Jensen et al., 1998). AKT has been shown to be phosphorylated by the PI3K pathway, thus rendering it active to phosphorylate the cytosolic protein 14-3-3 that sequesters Bad (Datta et al., 1997). Reports on the expression of the IL-7R on B lymphocytes have varied, but it now seems clear that the high affinity form of the receptor is present on IgM' cells (pro and pre B cells) (Sudo et al., 1993). If receptor activity was the same in pro and pre B cells, the differences in IL-7 mediated protection from apoptosis and cell cycle arrest must be attributed to variations in intracellular conditions and signaling. Clearly, the use of phenotypic and apoptotic analysis provided great insight into the response of primary developing B lymphocytes from the BM to specific cytokines. The use of primary cells for these studies was very important in determining responses such as survival, since 149 normal cell survival, cell proliferation and cell differentiation are often compromised in cell lines. Interestingly, IL-7 elicited a modest, yet significant protective effect against Cs- induced apoptosis in pre B lymphocytes even though it did not promote background survival in this cell type. This suggests that IL-7 may be utilizing different mechanisms to promote background survival (perhaps the AKT pathway) than that used in promoting protection from Cs-induced apoptosis. Since Bel-2 upregulation by IL-7 has been previously reported, it could be that protection from Cs is mediated by increased expression of this anti-apoptotic factor (Hemandez-Caselles et al., 1995; Karawajew et al., 2000). Also of interest, although apoptosis was decreased, IL-7 did not inhibit the Cs-induced cycle decrease in pre B cells. Therefore it appears from the pre B cell results, that the protective effect of IL-7 on Cs-induced apoptosis can be separate from its ability to inhibit cell cycle arrest. Considering the similar ability of IL—7 and SCF to provide protection of pro B cells from spontaneous cell death, it would be reasonable to assume similar effects on Cs- induced apoptosis. This was not the case. SCF provided no appreciable protection fiom Cs treatment, above that seen in background cultures, in any cell type. Also when used in combination with lL-7 it did not appear to augment the protective effects of IL—7. Poor activity of the recombinant SCF could be one explanation for the lack of apparent effects, but this is unlikely since titrations were performed to determine the maximum effect of SCF and SCF elicited protective effects on background apoptosis. Had there been no activity or low activity, it would be likely that SCF-induced protection from spontaneous apoptosis would not have been apparent. 150 By using methodology that allowed for the direct analysis of apoptosis in primary developing B lymphocytes it was determined that IL-7, in fact, could protect early B lymphocytes from Cs—induced apoptosis. Considering the devastating affect that Cs elevation can have on the immune system, IL-7 may warrant further investigation as an immunotheraputic agent. The mechanisms by which Cs elicits its effect or whereby IL-7 can inhibit those effects does not appear to be a simple process, the differential effects on apoptosis and cell cycle status between cell types suggests a complex interplay between these two signaling systems. It does appear that IL-7 mediated protection from background apoptosis and Cs-induced apoptosis utilizes at least two different pathways. Additionally, protection from apoptosis clearly occurs prior to induction of proliferation and the two processes appear to be regulated separately. SCF, while similar to IL-7 in protection from background apoptosis, does not appear to protect B lymphocytes from Cs-induced apoptosis. This suggests that either these cytokines initiate different modes of survival or that a different mechanism exists whereby IL-7 can also promote protection from Cs-induced death. A combinatorial approach between the analysis of biological responses and cell signaling analysis may aid in the elucidation of some of the interactions between Cs-mediated and IL—7-mediated responses. 151 Table 4.1 Recovery and viability of bone marrow cultured for 15 to 16 hours Media only 0.1 M Cs 0.1 rig/m1 IL-7 Cs and IL-7 Recovery‘ 126 i 9%c 106 i 8% 127 :1: 8% 123 i 18% Viability*" 93 :1: 3% 82 :1: 3% 90 i 3% 86 :1: 3% " Recovery was determined by dividing the viable and apoptotic cells obtained following culture by the number of viable cells added to the original culture b Viability was determined by dividing the viable and apoptotic cells obtained following culture by the total of viable, apoptotic and trypan blue+ cells (dead) ° Cells were cultured for 15 to 16 hours, the data are averaged from five separate experiments and standard deviations are shown * Cell viability was approximately 90% at the initiation of cell cultures 152 Figure 4.1 The effect of IL-7 on spontaneous levels of apoptosis verses Cs-induced apoptosis in pro, pre and IgM+ B lymphocytes. BM was cultured for 16 hours either with 0.1 ng/ml lL-7 only, 0.1 1.1M Cs only or with a combination of IL-7 and Cs. Phenotypic and DNA analysis were performed using flow cytometry. Panel (A) shows the change in spontaneous apoptosis elicited by the addition of IL-7. Data was normalized to the percent of apoptosis in cells cultured in media alone, with actual percent apoptosis of 7%, 16% and 11% for the pro, pre and IgM+. Panel (B) shows the change in Cs-induced apoptosis caused by IL-7. Data was normalized to the percent apoptosis in cells incubated with Cs alone, with actual percentages of 23%, 53% and 39% for pro, pre and lgM+ cells, respectively. lL-7 induced changes in apoptosis were determined for four separate experiments and were averaged with standard error bars shown. Significant differences were determined by the Student’s t-test (p < 0.05) and indicated with an asterisk. 153 No Cs 120% 100% ,2 80% .8 g. 60% 9- < 40% 20% 0% , ProB PreB 120% . Cs Treated 100% U) ‘5 80% 8 Q. a 60% < 40% 20% 0% i ProB PreB INo IL7 El 111ng IL7 Figure 4.1 IL-7 modulation of spontaneous and Cs-induced apoptosis 154 Figure 4.2 Data show the effect of added IL-7 on the S/Gz/M phases of the pro and pre B lymphocyte cell cycle of cells cultured in media alone or treated with Cs. BM was cultured for 16 hours in media alone (solid bars), with 0.1 ng/ml IL-7 alone (sparse dots), with 0.1 11M Cs alone (stripes) or with a combination of both IL-7 and Cs (dense dots). Data was normalized to 100% percent for pro or pre B cells in the S/GZ/M phase of the cell cycle when cultured in media alone; actual percentages were 17% for pro B cells and 15% for pre B cells (controls). The percentage S/G2/M was determined via flow cytometric analysis of the DNA content and pro and pre B cells were identified by phenotyping. The data shown are the average of two separate experiments and stande error bars are shown. Significant differences were determined by the Student’s t-test (p < 0.05) and indicated with an asterisk. 155 160% I IControl r a 1 E] IL-7 1 120% J ICs and IL-7 : E 100% ~ 1 N l 1 Q 80% « t . m , l i 60% — 1 are . 40% i DEL % ‘ ’ it. 3 20% 1 1c , and ' 0% 1 rd ~7 4 , # ——e _, J ‘p < Figure 4.2 IL—7 and/or Cs effect on cell cycle distribution among pro and pre B cells 156 Figure 4.3 A representative DNA flow cytometric profile of pro B lymphocytes cultured for 16 hours was determined by DAPI staining of phenotyped BM B cells. The hypodiploid, or apoptotic, populations are indicated by “A0”. Panel (A) shows the DNA content of pro B cells incubated in media alone. Panel (B) shows the percent A0 and S/Gz/M for cells incubated with 0.1 ng/ml lL-7 alone and panel (C) shows pro B cells incubated with 0.1 uM Cs. Panel (D) is the DNA profile of pro B cells incubated with both IL-7 and Cs. These data are representative of several experiments. 157 Relative Cell Number I llilll 1111111 1 D 1 06/6. 1 1 ji A0 S/szM ' 7.3% 17.7% FL 1 1 1 DAPI fluorescent (DNA Content) Figure 4.3 Flow cytometry of lL-7 and/or Cs effect on pro B cell apoptosis and cell cycle 158 Figure 4.4 Data show that IL7 protected B lymphocytes from Cs-induced apoptosis even at high concentrations of the steroid. The percent apOptosis, as determined flow cytometrically by DAPI stained DNA, is shown for (A) pro B (B) pre B and (C) lgM+ B cells. BM was cultured for 16 hours either with no Cs, 0.1 11M, 1 11M or 10 11M Cs (solid line) and the percent apoptosis was plotted. The percent apoptosis for each population incubated with the above concentrations of Cs and with 0.1 ng/ml IL7 (broken line) is also shown. Standard error bars are plotted and the data shown are representative of at least two separate experiments. 159 ttptosis even ow '0 13211.3 .\1 C s 150111 )pU1311011 1 11116115 nveofm % Apoptosis N 6 Pro B Cell 10 01“ *_,1____ N0 CS 0.1 pM Cs 1 ”M Cs 10 “M Cs 80 Pre B Cell % Apoptosis & G No Cs 0.1 ”M Cs 1.0 ”M Cs 10 11M Cs IgM+ B Cell % Apoptosis No Cs 0.1 “M Cs 1.0 “M Cs 10 “M Cs ° N0 IL7 "a" lug/m1 IL7 Figure 4.4 IL-7 effect on apoptosis induced by varying concentrations of Cs among B cells 160 Figure 4.5 The effect of SCF on spontaneous and Cs-induced apoptosis in bone marrow B cells. BM was cultured for 15 to 16 hours, cells were phenotyped for B cell subsets and DNA was stained with DAPI to determine the amount of apoptosis by flow cytometry. SCF was used at 0.1 rig/ml and Cs was used at 0.1 uM. Panel (A) shows the effect of SCF on spontaneous apoptosis in media alone. Data were normalized to the amount of apoptosis in cells cultured without SCF addition with actual percentages of 12%, 16% and 16% for pro, pre and lgM+ cells, respectively. Panel (B) shows the effect of SCF on Cs-induced apoptosis. Data were normalized to the level of apoptosis in cells cultured with Cs only with actual control percentages of 22%, 49% and 40%, respectively. Standard error bars are shown for the average of three separate experiments where significant differences were determined by the Student’s t-test (p < 0.05) and indicated by an asterisk. 161 tone marrow :11 subsets 0w \) shows the :ed 10 1116 :ntagCS 01 .vs the 9113“ tests in 65115 t'expt’fimm 051 and 125% ‘ 10096 1 75% j * 50% j 25% 3 0% 1 Pro Pre I Apoptosis 13M+ 1 1 1 I No treatment B SCF only ‘ B j 1 1 125% 4 I 1 m 100% § 75% 1 ~ 8 1 2- 50% 1 1 25% i 0% 1 1 Pro Pre IgM+ 3 CS C ”CS, 3995‘??? Figure 4.5 SCF effect on spontaneous and Cs—induced apoptosis 162 Table 4.2 Effects of IL—7 and SCF on apoptosis in pro B cells No Treatment _- SCF‘ ’ CSF— CS and SCF N" "’7 8.7 :1: 0.5% ° 43 :t 1.2%* 23.4 :1: 4.1% 16.3 i 1.9%" "-1 “g’m' "’7 3.6 :1: 2.2% 3.6 :1: 0.9% 7.3 :1: 2.8% 7.1 i 1.1% 0.01 ng/ml IL-7 N/D N/D 14.1 :1: 0.7% 13.8 i- 23% 0.001 118/ml 1L'7 M) MD 18.1 :1: 2.8% 19.2 i 1.0% ' SCF was used at 100 ng/ml b Cs was used at 0.1 MM ° Data are the averages of duplicate experiment :1: the standard deviations and represent at least two experiments . Significantly different from control sample (p < 0.05) " Significantly different from Cs sample (p < 0.1) 163 REFERENCES 164 "L Aiuti, A., Webb, 1. J., Bleul, C., Springer, T., and Gutierrez-Ramos, J. C. (1997). The chemokine SDF -1 is a chemoattractant for human CD34+ hematopoietic progenitor cells and provides a new mechanism to explain the mobilization of CD34+ progenitors to peripheral blood. J Exp Med 185, 111-20. Akashi, K., Reya, T., Dalma-Weiszhausz, D., and Weissman, l. L. (2000). Lymphoid precursors. Ctur Opin Immunol 12, 144-50. Almlof, T., Gustafsson, J. A., and Wright, A. P. (1997). Role of hydrophobic amino acid clusters in the transactivation activity of the human glucocorticoid receptor. Mol Cell Biol 1 7, 934-45. Andreau, K., Lemaire, C., Souvannavong, V., and Adam, A. (1998). Induction of apoptosis by dexamethasone in the B cell lineage [In Process Citation]. Irnmunopharrnacology 40, 67-76. Ashbumer, M., Chihara, C., Meltzer, P., and Richards, G. (1974). Temporal control of puffing activity in polytene chromosomes. Cold Spring Harb Symp Quant Biol 38, 655- 62. Ashman, L. K. (1999). The biology of stem cell factor and its receptor C-kit. Int J Biochem Cell Biol 31, 1037-51. Auphan, N., DiDonato, J. A., Rosette, C., Hehnberg, A., and Karin, M. (1995). Immunosuppression by glucocorticoids: inhibition of NF -kappa B activity through induction of l kappa B synthesis [see comments]. Science 270, 286-90. Bain, G., Maandag, E. C., Izon, D. J., Amsen, D., Kruisbeek, A. M., Weintraub, B. C., Krop, 1., Schlissel, M. S., Feeney, A. J ., van Roon, M., and et al. (1994). E2A proteins are required for proper B cell development and initiation of immunoglobulin gene rearrangements [see comments]. Cell 79, 885-92. Bain, G., and Murre, C. (1998). The role of E-proteins in B- and T-lymphocyte development. Semin Immunol 10, 143-53. Bakhshi, A., Jensen, J. P., Goldman, P., Wright, J. J., McBride, 0. W., Epstein, A. L., and Korsmeyer, S. J. (1985). Cloning the chromosomal breakpoint of t(14;18) human lymphomas: clustering around JH on chromosome 14 and near a transcriptional unit on 18. Cell 41, 899-906. Ballard, P. L., Baxter, J. D., Higgins, S. J., Rousseau, G. G., and Tomkins, G. M. (1974). General presence of glucocorticoid receptors in mammalian tissues. Endocrinology 94, 998-1002. 165 Bamberger, C. M., Bamberger, A. M., de Castro, M., and Chrousos, G. P. (1995). Glucocorticoid receptor beta, a potential endogenous inhibitor of glucocorticoid action in humans. J Clin Invest 95, 2435-41. Baxter, J. D., and Tomkins, G. M. (1971). Specific cytoplasmic glucocorticoid hormone receptors in hepatoma tissue culture cells. Proc Natl Acad Sci U S A 68, 932-7. Beato, M., Chalepakis, G., Schauer, M., and Slater, E. P. (1989). DNA regulatory elements for steroid hormones. J Steroid Biochem 32, 737-47. Bicknell, G. R., and Cohen, G. M. (1995). Cleavage of DNA to large kilobase pair fragments occurs in some forms of necrosis as well as apoptosis. Biochem Biophys Res Commun 207, 40—7. Blume-Jensen, P., Janknecht, R., and Hunter, T. (1998). The kit receptor promotes cell survival via activation of PI 3-kinase and subsequent Akt-mediated phosphorylation of Bad on Serl36. Curr Biol 8, 779-82. Bodwell, J. E., Webster, J. C., Jewell, C. M., Cidlowski, J. A., Hu, J. M., and Munck, A. (1998). Glucocorticoid receptor phosphorylation: overview, function and cell cycle- dependence. J Steroid Biochem Mol Biol 65, 91-9. Boise, L. H., Gonzalez-Garcia, M., Postema, C. E., Ding, L., Lindsten, T., Turka, L. A., Mao, X., Nunez, G., and Thompson, C. B. (1993). bcl-x, a bcl-2-related gene that firnctions as a dominant regulator of apoptotic cell death. Cell 74, 597-608. Boldin, M. P., Goncharov, T. M., Goltsev, Y. V., and Wallach, D. (1996). Involvement of MACH, a novel MORTl/FADD-interacting protease, in Fas/APO-l- and TNF receptor- induced cell death. Cell 85, 803—15. Borghesi, L. A., Smithson, G., and Kincade, P. W. (1997). Stromal cell modulation of negative regulatory signals that influence apoptosis and proliferation of B lineage lymphocytes. J Immunol 159, 4171-9. Bosma, G. C., Chang, Y., Karasuyama, H., and Bosma, M. J. (1999). Differential effect of an Ig mu transgene on development of pre-B cells in fetal and adult SCID mice. Proc Natl Acad Sci U S A 96, 11952-7. Bosma, G. C., Custer, R. P., and Bosma, M. J. (1983). A severe combined immunodeficiency mutation in the mouse. Nature 301, 527-30. Bright, G. M. (1995). Corticosteroid-binding globulin influences kinetic parameters of plasma cortisol transport and clearance. J Clin Endocrinol Metab 80, 770-5. 166 Burkly, L., Hession, C., Ogata, L., Reilly, C., Marconi, L. A., Olson, D., Tizard, R., Cate, R., and Lo, D. (1995). Expression of relB is required for the development of thymic medulla and dendritic cells. Nature 373, 531-6. Bumstein, K. L., and Cidlowski, J. A. (1989). Regulation of gene expression by glucocorticoids. Annu Rev Physiol 51, 683-99. Bumstein, K. L., Jewell, C. M., Sar, M., and Cidlowski, J. A. (1994). Intragenic sequences of the human glucocorticoid receptor complementary DNA mediate honnone- inducible receptor messenger RNA down-regulation through multiple mechanisms. Mol Endocrinol 8, 1764-73. Cerezo, A., Martinez, A. C., Lanzarot, D., Fischer, S., Franke, T. F ., and Rebollo, A. (1998). Role of Akt and c-Jun N-terrninal kinase 2 in apoptosis induced by interleukin-4 deprivation. Mol Biol Cell 9, 3107-18. Chakraborti, P. K., Garabedian, M. J ., Yamamoto, K. R., and Simons, S. S., Jr. (1992). Role of cysteines 640, 656, and 661 in steroid binding to rat glucocorticoid receptors. J Biol Chem 267, 11366-73. Chandler, V. L., Maler, B. A., and Yamamoto, K. R. (1983). DNA sequences bound specifically by glucocorticoid receptor in vitro render a heterologous promoter hormone responsive in vivo. Cell 33, 489-99. Chang, Y., Bosma, G. C., and Bosma, M. J. (1995). Development of B cells in scid mice with immunoglobulin transgenes: implications for the control of V(D)J recombination. Immunity 2, 607-16. Cheng, E. H., Kirsch, D. G., Clem, R. J., Ravi, R., Kastan, M. B., Bedi, A., Ueno, K., and Hardwick, J. M. (1997). Conversion of Bel-2 to a Bax-like death effector by caspases. Science 278, 1966-8. Cidlowski, J. A., and Cidlowski, N. B. (1981). Regulation of glucocorticoid receptors by glucocorticoids in cultured HeLa S3 cells. Endocrinology 109, 1975-82. Cifone, M. G., Migliorati, G., Parroni, R., Marchetti, C., Millirnaggi, D., Santoni, A., and Riccardi, C. (1999). Dexamethasone-induced thymocyte apoptosis: apoptotic signal involves the sequential activation of phosphoinositide-specific phospholipase C, acidic sphingomyelinase, and caspases. Blood 93, 2282-96. Cleary, M. L., Smith, S. D., and Sklar, J. (1986). Cloning and structural analysis of cDNAs for bcl-2 and a hybrid bcl-2/immunoglobulin transcript resulting from the t(14;18) translocation. Cell 47, 19-28. Coffinan, R. L. (1982). Surface antigen expression and immunoglobulin gene rearrangement during mouse pre-B cell development. hnmunol Rev 69, 5-23. 167 Coffinan, R. L., and Weissman, I. L. (1981). B220: a B cell-specific member of th T200 glycoprotein family. Nature 289, 681-3. Cohen, J. J. (1992). Glucocorticoid-induced apoptosis in the thymus. Semin Immunol 4, 363-9. Collins, L. S., and Dorshkind, K. (1987). A stromal cell line fiom myeloid long-term bone marrow cultures can support myelopoiesis and B lymphopoiesis. J Immunol 138, 1082-7. Corcoran, A. E., Riddell, A., Krooshoop, D., and Venkitaraman, A. R. (1998). Impaired immunoglobulin gene rearrangement in mice lacking the IL-7 receptor. Nature 391, 904- 7. Corcoran, A. E., Smart, F. M., Cowling, R. J., Crompton, T., Owen, M. J ., and Venkitaraman, A. R. (1996). The interleukin-7 receptor alpha chain transmits distinct signals for proliferation and differentiation during B lymphopoiesis. Embo J 15, 1924-32. Dadi, H. K., Ke, S., and Roifman, C. M. (1993). Interleukin 7 receptor mediates the activation of phosphatidylinositol-3 kinase in human B-cell precursors. Biochem Biophys Res Commun 192, 459-64. Dancescu, M., Rubio-Trujillo, M., Biron, G., Bron, D., Delespesse, G., and Sarfati, M. (1992). Interleukin 4 protects chronic lymphocytic leukemic B cells from death by apoptosis and upregulates Bel-2 expression. J Exp Med 176, 1319-26. Danielian, P. S., White, R., Lees, J. A., and Parker, M. G. (1992). Identification of a conserved region required for hormone dependent transcriptional activation by steroid hormone receptors [published erratum appears in EMBO J 1992 Jun;11(6):2366]. Embo J I 1, 1025-33. Datta, S. R., Dudek, H., Tao, X., Masters, S., Fu, H., Gotoh, Y., and Greenberg, M. E. (1997). Akt phosphorylation of BAD couples survival signals to the cell-intrinsic death machinery. Cell 91, 231-41. de Bruijn, M. F., Slieker, W. A., van der Loo, J. C., Voerman, J. S., van Ewijk, W., and Leenen, P. J. (1994). Distinct mouse bone marrow macrophage precursors identified by differential expression of ER-MP12 and ER-MP20 antigens. Eur J Immunol 24, 2279-84. DePasquale-Jardieu, P., and Fraker, P. J. (1980). Further characterization of the role of corticosterone in the loss of humoral immunity in zinc-deficient A/J mice as determined by adrenalectomy. J Immunol 124, 2650-5. 168 DePasquale-Jardieu, P., and Fraker, P. J. (1984). Interference in the development of a secondary immune response in mice by zinc deprivation: persistence of effects. J Nutr 114, 1762-9. Desiderio, S. V., Yancopoulos, G. D., Paskind, M., Thomas, E., Boss, M. A., Landau, N., Alt, F. W., and Baltimore, D. (1984). Insertion of N regions into heavy-chain genes is correlated with expression of terminal deoxytransferase in B cells. Nature 311, 752-5. Dexter, T. M., Allen, T. D., and Lajtha, L. G. (1977). Conditions controlling the proliferation of haemopoietic stem cells in vitro. J Cell Physiol 91, 335-44. Dittel, B. N., McCarthy, J. B., Wayner, E. A., and LeBien, T. W. (1993). Regulation of human B-cell precursor adhesion to bone marrow stromal cells by cytokines that exert opposing effects on the expression of vascular cell adhesion molecule-1 (VCAM-1). Blood 81, 2272-82. Dorshkind, K., Johnson, A., Collins, L., Keller, G. M., and Phillips, R. A. (1986). Generation of purified stromal cell cultures that support lymphoid and myeloid precursors. J Immunol Methods 89, 37-47. Engler, D., Redei, E., and Kola, I. (1999). The corticotropin-release inhibitory factor hypothesis: a review of the evidence for the existence of inhibitory as well as stimulatory hypophysiotropic regulation of adrenocorticotropin secretion and biosynthesis. Endocr Rev 20, 460-500. Fadok, V. A., and Henson, P. M. (1998). Apoptosis: getting rid of the bodies. Curr Biol 8, R693-5. Fadok, V. A., Voelker, D. R., Campbell, P. A., Cohen, J. J., Bratton, D. L., and Henson, P. M. (1992). Exposure of phosphatidylserine on the surface of apoptotic lymphocytes triggers specific recognition and removal by macrophages. J Immunol 148, 2207-16. Faust, E. A., Saffran, D. C., Toksoz, D., Williams, D. A., and Witte, O. N. (1993). Distinctive growth requirements and gene expression patterns distinguish progenitor B cells from pre-B cells. J Exp Med 1 7 7, 915-23. Feng, W., Ribeiro, R. C., Wagner, R. L., Nguyen, H., Apriletti, J. W., Fletterick, R. J ., Baxter, J. D., Kushner, P. J ., and West, B. L. (1998). Hormone-dependent coactivator binding to a hydrophobic cleft on nuclear receptors. Science 280, 1747-9. Fraker, P. J ., Haas, S. M., and Luecke, R. W. (1977). Effect of zinc deficiency on the immune response of the young adult A/J mouse. J Nutr [07, 1889-95. Frasch, S. C., Henson, P. M., Kailey, J. M., Richter, D. A., Janes, M. S., Fadok, V. A., and Bratton, D. L. (2000). Regulation of phospholipid scramblase activity during apoptosis and cell activation by PKC {delta}. J Biol Chem 275, 23065-73. I69 Freedman, L. P., and Luisi, B. F. (1993). On the mechanism of DNA binding by nuclear hormone receptors: a structural and functional perspective. J Cell Biochem 51, 140-50. Friend, S. L., Hosier, S., Nelson, A., Foxworthe, D., Williams, D. E., and Farr, A. (1994). A thymic stromal cell line supports in vitro development of surface IgM+ B cells and produces a novel growth factor affecting B and T lineage cells. Exp Hematol 22, 321-8. Funk, P. E., Stephan, R. P., and Witte, P. L. (1995). Vascular cell adhesion molecule 1- positive reticular cells express interleukin-7 and stem cell factor in the bone marrow. Blood 86, 2661-71. Garvy, B. A., King, L. E., Telford, W. G., Morford, L. A., and Fraker, P. J. (1993). Chronic elevation of plasma corticosterone causes reductions in the number of cycling cells of the B lineage in murine bone marrow and induces apoptosis. Immunology 80, 587-92. Garvy, B. A., Telford, W. G., King, L. E., and Fraker, P. J. (1993). Glucocorticoids and irradiation-induced apoptosis in normal murine bone marrow B-lineage lymphocytes as determined by flow cytometry. Immunology 79, 270-7. Gibson, L. F., Piktel, D., Narayanan, R., Nunez, G., and Landreth, K. S. (1996). Stromal cells regulate bcl-2 and bax expression in pro-B cells. Exp Hematol 24, 628-37. Goodwin, R. 6., Friend, D., Ziegler, S. F., Jerzy, R., Falk, B. A., Gimpel, S., Cosman, D., Dower, S. K., March, C. J ., Namen, A. E., and Park, L. S. (1990). Cloning of the Human and Murine Interleukin-7 Receptors: Demonstration of a Soluble Form and Homology to a New Receptor Superfamily. Cell 60, 941-951. Griffiths, S. D., Goodhead, D. T., Marsden, S. J ., Wright, E. G., Krajewski, S., Reed, J. C., Korsmeyer, S. J ., and Greaves, M. (1994). Interleukin 7-dependent B lymphocyte precursor cells are ultrasensitive to apoptosis. J Exp Med 1 79, 1789-97. Gustafsson, J. A., Wikstrom, A. C., and Denis, M. (1989). The non-activated glucocorticoid receptor: structure and activation. J Steroid Biochem 34, 53-62. Hagman, J ., Wheat, W., Fitzsimmons, D., Hodsdon, W., Negri, J ., and Dizon, F. (2000). Pax-S/BSAP: regulator of specific gene expression and differentiation in B lymphocytes. Curr Top Microbiol Immunol 245, 169-94. Hakem, R., Hakem, A., Duncan, G. S., Henderson, J. T., Woo, M., Soengas, M. S., Elia, A., de la Pompa, J. L., Kagi, D., Khoo, W., Potter, J., Yoshida, R., Kaufman, S. A., Lowe, S. W., Penninger, J. M., and Mak, T. W. (1998). Differential requirement for caspase 9 in apoptotic pathways in vivo. Cell 94, 339-52. 170 Hardy, R. R., Carmack, C. E., Shinton, S. A., Kemp, J. D., and Hayakawa, K. (1991). Resolution and characterization of pro-B and pre-pro-B cell stages in normal mouse bone marrow. J Exp Med 173, 1213-25. Hardy, R. R., Kemp, J. D., and Hayakawa, K. (1989). Analysis of lymphoid population in scid mice; detection of a potential B lymphocyte progenitor population present at normal levels in scid mice by three color flow cytometry with B220 and S7. Curr Top Microbiol Immunol 152, 19-25. Heck, S., Kullmann, M., Gast, A., Ponta, H., Rahmsdorf, H. J ., Herrlich, P., and Cato, A. C. (1994). A distinct modulating domain in glucocorticoid receptor monomers in the repression of activity of the transcription factor AP-l. Embo J 13, 4087-95. Hench, P. S., Kendall E.C., Slocumb, CH. and Polley HF. (1949). The effect of a hormone of the adrenal cortex (17-hydroxy-ll-dehydroxycorticosterone: compound E) and of pituitary adrenocorticotropic hormone on rheumatoid arthritis. Proceedings of the Staff Meeting at Mayo Clinic 24, 181-197. Henderson, A. J ., Johnson, A., and Dorshkind, K. (1990). Functional characterization of two stromal cell lines that support B lymphopoiesis. J Immunol 145, 423-8. Hengartner, M. 0., and Horvitz, H. R. (1994). C. elegans cell survival gene ced-9 encodes a functional homolog of the mammalian proto-oncogene bcl-2. Cell 76, 665-76. Higgins, S. J., Rousseau, G. G., Baxter, J. D., and Tomkins, G. M. (1973). Early events in glucocorticoid action. Activation of the steroid receptor and its subsequent specific nuclear binding studied in a cell-free system. J Biol Chem 248, 5866-72. Hirayarna, F., Lyman, S. D., Clark, S. C., and Ogawa, M. (1995). The flt3 ligand supports proliferation of lymphohematopoietic progenitors and early B-lymphoid progenitors. Blood 85, 1762-8. Hollenberg, S. M., Weinberger, C., Ong, E. S., Cerelli, G., Oro, A., Lebo, R., Thompson, E. B., Rosenfeld, M. G., and Evans, R. M. (1985). Primary structure and expression of a functional human glucocorticoid receptor cDNA. Nature 318, 635-41. Hombach, J ., Tsubata, T., Leclercq, L., Stappert, H., and Reth, M. (1990). Molecular components of the B-cell antigen receptor complex of the IgM class. Nature 343, 760-2. Hong, H., Kohli, K., Trivedi, A., Johnson, D. L., and Stallcup, M. R. (1996). GRIP], a novel mouse protein that serves as a transcriptional coactivator in yeast for the hormone binding domains of steroid receptors. Proc Natl Acad Sci U S A 93, 4948-52. Hough, M. R., Chappel, M. S., Sauvageau, G., Takei, F., Kay, R., and Humphries, R. K. (1996). Reduction of early B lymphocyte precursors in transgenic mice overexpressing the murine heat-stable antigen. J Immunol 156, 479-88. 171 Howard, A. D., Kostura, M. J ., Thomberry, N., Ding, G. J ., Limjuco, G., Weidner, J ., Salley, J. P., Hogquist, K. A., Chaplin, D. D., Mumford, R. A., and et al. (1991). IL-l- converting enzyme requires aspartic acid residues for processing of the IL-1 beta precursor at two distinct sites and does not cleave 3l-kDa IL-l alpha. J Immunol 14 7, 2964-9. Hu, J. M., Bodwell, J. E., and Munck, A. (1997). Control by basal phosphorylation of cell cycle-dependent, hormone-induced glucocorticoid receptor hyperphosphorylation. Mol Endocrinol 11, 305-1 1. Huang, S. T., and Cidlowski, J. A. (1999). Glucocorticoids inhibit serum depletion- induced apoptosis in T lymphocytes expressing Bel-2. F aseb J 13, 467-76. Jacobsen, K., and Osmond, D. G. (1990). Microenvironmental organization and stromal cell associations of B lymphocyte precursor cells in mouse bone marrow. Eur J Immunol 20, 2395-404. Jacobsen, K., Tepper, J ., and Osmond, D. G. (1990). Early B-lymphocyte precursor cells in mouse bone marrow: subosteal localization of B220+ cells during postirradiation regeneration. Exp Hematol 18, 304-10. Jarvis, L. J ., and LeBien, T. W. (1995). Stimulation of human bone marrow stromal cell tyrosine kinases and IL-6 production by contact with B lymphocytes. J Immunol 155, 2359-68. Jarvis, L. J ., Maguire, J. E., and LeBien, T. W. (1997). Contact between human bone marrow stromal cells and B lymphocytes enhances very late anti gen-4/vascular cell adhesion molecule-1 -independent tyrosine phosphorylation of focal adhesion kinase, paxillin, and ERK2 in stromal cells. Blood 90, 1626-35. Jurgensrneier, J. M., Xie, Z., Deveraux, Q., Ellerby, L., Bredesen, D., and Reed, J. C. (1998). Bax directly induces release of cytochrome c from isolated mitochondria. Proc Natl Acad Sci U S A 95, 4997-5002. Justement, L. B. (2000). Signal transduction via the B-cell antigen receptor: the role of protein tyrosine kinases and protein tyrosine phosphatases. Curr Top Microbiol Immunol 245 , 1-51. Karnada, S., Shimono, A., Shinto, Y., Tsujimura, T., Takahashi, T., Noda, T., Kitarnura, Y., Kondoh, H., and Tsujimoto, Y. (1995). bcl-2 deficiency in mice leads to pleiotropic abnormalities: accelerated lymphoid cell death in thymus and spleen, polycystic kidney, hair hypopigrnentation, and distorted small intestine. Cancer Res 55, 354-9. 172 Karasuyama, H., Kudo, A., and Melchers, F. (1990). The proteins encoded by the VpreB and lambda 5 pre-B cell-specific genes can associate with each other and with mu heavy chain. J Exp Med 172, 969-72. Karasuyama, H., Rolink, A., Shinkai, Y., Young, F ., Alt, F. W., and Melchers, F. (1994). The expression of Vpre-B/lambda 5 surrogate light chain in early bone marrow precursor B cells of normal and B cell-deficient mutant mice. Cell 7 7, 133-43. Karawajew, L., Ruppert, V., Wuchter, C., Kosser, A., Schrappe, M., Dorken, B., and Ludwig, W. D. (2000). Inhibition of in vitro spontaneous apoptosis by IL-7 correlates with bcl-2 up-regulation, cortical/mature irnmunophenotype, and better early cytoreduction of childhood T-cell acute lymphoblastic leukemia. Blood 96, 297-306. Karin, M., Haslinger, A., Holtgreve, H., Richards, R. I., Krauter, P., Westphal, H. M., and Beato, M. (1984). Characterization of DNA sequences through which cadmium and glucocorticoid hormones induce human metallothionein-IIA gene. Nature 308, 513-9. Kashiwamura, S., Koyama, T., Matsuo, T., Steinmetz, M., Kirnoto, M., and Sakaguchi, N. (1990). Structure of the murine mb-l gene encoding a putative sIgM-associated molecule. J Immunol 145, 337-43. Katagiri, T., Ogimoto, M., Hasegawa, K., Arimura, Y., Mitomo, K., Okada, M., Clark, M. R., Mizuno, K., and Yakura, H. (1999). CD45 negatively regulates 1yn activity by dephosphorylating both positive and negative regulatory tyrosine residues in immature B cells. J Immunol 163, 1321-6. Kayalar, C., Ord, T., Testa, M. P., Zhong, L. T., and Bredesen, D. E. (1996). Cleavage of actin by interleukin 1 beta-converting enzyme to reverse DNase I inhibition. Proc Natl Acad Sci U S A 93, 2234-8. Kerr, J. F., Wyllie, A. H., and Currie, A. R. (1972). Apoptosis: a basic biological phenomenon with wide-ranging implications in tissue kinetics. Br J Cancer 26, 239-57. Khaled, A. R., Kim, K., Hofineister, R., Muegge, K., and Durum, S. K. (1999). Withdrawal of IL-7 induces Bax translocation from cytosol to mitochondria through a rise in intracellular pH [see comments]. Proc Natl Acad Sci U S A 96, 14476-81. Kierney, P. C., and Dorshkind, K. (1987). B lymphocyte precursors and myeloid progenitors survive in diffusion chamber cultures but B cell differentiation requires close association with stromal cells. Blood 70, 1418-24. Kim, K., Lee, C. K., Sayers, T. J ., Muegge, K., and Durum, S. K. (1998). The trophic action of IL-7 on pro-T cells: inhibition of apoptosis of pro-T1 , -T2, and -T3 cells correlates with Bcl-2 and Bax levels and is independent of Fas and p53 pathways. J Immunol 160, 5735-41. 173 King, K. L., and Cidlowski, J. A. (1998). Cell cycle regulation and apoptosis. Annu Rev Physiol 60, 601-17. King, L. E., Osati-Ashtiani, F ., and Fraker, P. J. (1995). Depletion of cells of the B lineage in the bone marrow of zinc-deficient mice. Immunology 85, 69-73. Kitarnura, D., Kudo, A., Schaal, S., Muller, W., Melchers, F., and Rajewsky, K. (1992). A critical role of lambda 5 protein in B cell development. Cell 69, 823-31. Kondo, M., Takeshita, T., Higuchi, M., Nakamura, M., Sudo, T., Nishikawa, S., and Sugamura, K. (1994). Functional participation of the IL—2 receptor gamma chain in IL-7 receptor complexes. Science 263, 1453-4. Kudo, A., and Melchers, F. (1987). A second gene, VpreB in the lambda 5 locus of the mouse, which appears to be selectively expressed in pre-B lymphocytes. Embo J 6, 2267- 72. Kuo, M. H., and Allis, C. D. (1998). Roles of histone acetyltransferases and deacetylases in gene regulation. Bioessays 20, 615-26. Laferte, S., and Dennis, J. W. (1988). Glycosylation-dependent collagen-binding activities of two membrane glycoproteins in MDAY-D2 tumor cells. Cancer Res 48, 4743-8. Landau, N. R., Schatz, D. G., Rosa, M., and Baltimore, D. (1987). Increased frequency of N-region insertion in a murine pre-B-cell line infected with a terminal deoxynucleotidyl transferase retroviral expression vector. Mol Cell Biol 7, 3237-43. Lavagna-Sevenier, C., Marchetto, S., Birnbaum, D., and Rosnet, O. (1998). The CBL- related protein CBLB participates in FLT3 and interleukin-7 receptor signal transduction in pro-B cells. J Biol Chem 273, 14962-7. LeBien, T. W. (1998). B-cell lymphopoiesis in mouse and man [comment]. Curr Opin Immunol 10, 188-95. Lechner, O., Wiegers, G. J ., Oliveira-Dos-Santos, A. J ., Dietrich, H., Recheis, H., Waterman, M., Boyd, R., and Wick, G. (2000). Glucocorticoid production in the murine thymus. Eur J Immunol 30, 337-46. Lee, G., Namen, A. E., Gillis, S., Ellingsworth, L. R., and Kincade, P. W. (1989). Normal B cell precursors responsive to recombinant murine IL-7 and inhibition of IL-7 activity by transforming growth factor-beta. J Immunol 142, 3875-83. Lee, G., Namen, A. E., Gillis, S., and Kincade, P. W. (1988). Recombinant interleukin-7 supports the growth of normal B lymphocyte precursors. Curr Top Microbiol Immunol 141, 16-8. 174 Lee, S. H., Fujita, N., Mashima, T., and Tsuruo, T. (1996). Interleukin-7 inhibits apoptosis of mouse malignant T-lymphoma cells by both suppressing the CPP32-1ike protease activation and inducing the Bcl-2 expression. Oncogene 13, 2131-9. Levin, S. D., Koelling, R. M., Friend, S. L., Isaksen, D. E., Ziegler, S. F ., Perlrnutter, R. M., and Farr, A. G. (1999). Thymic stromal lymphopoietin: a cytokine that promotes the development of IgM+ B cells in vitro and signals via a novel mechanism. J Immunol 162, 677-83. Li, Y. S., Hayakawa, K., and Hardy, R. R. (1993). The regulated expression of B lineage associated genes during B cell differentiation in bone marrow and fetal liver. J Exp Med 1 78, 951-60. Liles, W. C., Dale, D. C., and Klebanoff, S. J. (1995). Glucocorticoids inhibit apoptosis of human neutrophils. Blood 86, 3181-8. Lin, Q., Taniuchi, 1., Kitamura, D., Wang, J ., Kearney, J. F., Watanabe, T., and Cooper, M. D. (1998). T and B cell development in BP-l/6C3/aminopeptidase A-deficient mice. J Immunol 160, 4681-7. Linnekin, D. (1999). Early signaling pathways activated by c-Kit in hematopoietic cells. Int J Biochem Cell Biol 31, 1053-74. Lord, J. D., McIntosh, B. C., Greenberg, P. D., and Nelson, B. H. (2000). The IL-2 receptor promotes lymphocyte proliferation and induction of the c-myc, bcl-2, and bcl-x genes through the trans-activation domain of Stat5. J Immunol 164, 2533-41. Los, M., Van de Craen, M., Penning, L. C., Schenk, H., Westendorp, M., Baeuerle, P. A., Droge, W., Krammer, P. H., Fiers, W., and Schulze-Osthoff, K. (1995). Requirement of an ICE/CED-3 protease for Fas/APO-l-mediated apoptosis. Nature 375, 81-3. Lu, L., Chaudhury, P., and Osmond, D. G. (1999). Regulation of cell survival during B lymphopoiesis: apoptosis and Bcl-2/Bax content of precursor B cells in bone marrow of mice with altered expression of IL-7 and recombinase-activating gene-2. J Immunol 162, 1931-40. Lu, L., and Osmond, D. G. (1997 ). Apoptosis during B lymphopoiesis in mouse bone marrow. J Immunol 158, 5136-45. Lu, L., Smithson, G., Kincade, P. W., and Osmond, D. G. (1998). Two models of murine B lymphopoiesis: a correlation. Eur J Immunol 28, 1755-61. Lucibello, F. C., Slater, E. P., Jooss, K. U., Beato, M., and Muller, R. (1990). Mutual transrepression of Fos and the glucocorticoid receptor: involvement of a functional domain in Fos which is absent in F088. Embo J 9, 2827-34. 175 Ma, A., Pena, J. C., Chang, 8., Margosian, E., Davidson, L., Alt, F. W., and Thompson, C. B. (1995). Bclx regulates the survival of double-positive thymocytes. Proc Natl Acad Sci U S A 92, 4763-7. Maeurer, M. J ., Trinder, P., Homrnel, G., Walter, W., Freitag, K., Atkins, D., and Storkel, S. (2000). Interleukin-7 or interleukin-15 enhances survival of Mycobacterium tuberculosis-infected mice. Infect Irnmun 68, 2962-70. Malabarba, M. G., Kirken, R. A., Rui, H., Koettnitz, K., Kawamura, M., O'Shea, J. J ., Kalthoff, F. S., and Farrar, W. L. (1995). Activation of JAK3, but not JAKl, is critical to interleukin-4 (1L4) stimulated proliferation and requires a membrane-proxirnal region of 1L4 receptor alpha. J Biol Chem 270, 9630-7. Manabe, A., Murti, K. G., Coustan-Smith, E., Kurnagai, M., Behm, F. G., Raimondi, S. C., and Campana, D. (1994). Adhesion-dependent survival of normal and leukemic human B lymphoblasts on bone marrow stromal cells. Blood 83, 758-66. Maraskovsky, E., Peschon, J. J ., McKenna, H., Teepe, M., and Strasser, A. (1998). Overexpression of Bcl-2 does not rescue impaired B lymphopoiesis in IL-7 receptor- deficient mice but can enhance survival of mature B cells. Int Immunol 10, 1367-75. Marchetti, P., Hirsch, T., Zamzami, N., Castedo, M., Decaudin, D., Susin, S. A., Masse, B., and Kroemer, G. (1996). Mitochondrial permeability transition triggers lymphocyte apoptosis. J Immunol 15 7, 4830-6. Matsuo, T., Nomura, J ., Kuwahara, K., Igarashi, H., Inui, S., Hamaguchi, M., Kimoto, M., and Sakaguchi, N. (1993). Cross-linking of B cell receptor-related MB-l molecule induces protein tyrosine phosphorylation in early B lineage cells. J Immunol 150, 3766- 75. Matthews, W., Jordan, C. T., Wiegand, G. W., Pardoll, D., and Lemischka, I. R. (1991). A receptor tyrosine kinase specific to hematopoietic stem and progenitor cell-enriched populations. Cell 65, 1143-52. Maundrell, K., Antonsson, B., Magnenat, E., Camps, M., Muda, M., Chabert, C., Gillieron, C., Boschert, U., Vial-Knecht, E., Martinou, J. C., and Arkinstall, S. (1997). Bel-2 undergoes phosphorylation by c-Jun N-terrninal kinase/stress-activated protein kinases in the presence of the constitutively active GTP-binding protein Racl. J Biol Chem 272, 25238-42. McCarthy, N. J ., Whyte, M. K., Gilbert, C. S., and Evan, G. I. (1997). Inhibition of Ced- 3/ICE-related proteases does not prevent cell death induced by oncogenes, DNA damage, or the Bcl-2 homologue Bak. J Cell Biol 136, 215-27. 176 McConkey, D. J ., Nicotera, P., Hartzell, P., Bellomo, G., Wyllie, A. H., and Orrenius, S. (1989). Glucocorticoids activate a suicide process in thymocytes through an elevation of cytosolic Ca2+ concentration. Arch Biochem Biophys 269, 365-70. McKercher, S. R., Torbett, B. E., Anderson, K. L., Henkel, G. W., Vestal, D. J ., Baribault, H., Klemsz, M., Feeney, A. J., Wu, G. E., Paige, C. J., and Maki, R. A. (1996). Targeted disruption of the PU.1 gene results in multiple hematopoietic abnormalities. Embo J 15, 5647-58. McNiece, I. K., Langley, K. E., and Zsebo, K. M. (1991). The role of recombinant stem cell factor in early B cell development. Synergistic interaction with IL-7. J Immunol 146, 3785-90. Mendel, C. M., Kuhn, R. W., Weisiger, R. A., Cavalieri, R. R., Siiteri, P. K., Cunha, G. R., and Murai, J. T. (1989). Uptake of cortisol by the perfused rat liver: validity of the fi'ee hormone hypothesis applied to cortisol. Endocrinology 124, 468-76. Merino, R., Ding, L., Veis, D. J ., Korsmeyer, S. J ., and Nunez, G. (1994). Developmental regulation of the Bel-2 protein and susceptibility to cell death in B lymphocytes. Embo J 13, 683-91. Mertsching, E., Grawunder, U., Meyer, V., Rolink, T., and Ceredig, R. (1996). Phenotypic and functional analysis of B lymphopoiesis in interleukin-7- transgenic mice: expansion of pro/pre-B cell number and persistence of B lymphocyte development in lymph nodes and spleen. Eur J Immunol 26, 28-33. Meyaard, L., Otto, S. A., Jonker, R. R., Mijnster, M. J., Keet, R. P., and Miedema, F. (1992). Programmed death of T cells in HIV-1 infection. Science 257, 217-9. Mills, J. C., Stone, N. L., Erhardt, J ., and Pittman, R. N. (1998). Apoptotic membrane blebbing is regulated by myosin light chain phosphorylation. J Cell Biol 140, 627-36. Minn, A. J ., Velez, P., Schendel, S. L., Liang, H., Muchmore, S. W., Fesik, S. W., Fill, M., and Thompson, C. B. (1997). Bcl-x(L) forms an ion channel in synthetic lipid membranes. Nature 385, 353-7. Miura, M., Zhu, H., Rotello, R., Hartwieg, E. A., and Yuan, J. (1993). Induction of apoptosis in fibroblasts by IL-1 beta-converting enzyme, a mammalian homolog of the C. elegans cell death gene ced-3. Cell 75, 653-60. Miyake, K., Medina, K., Ishihara, K., Kirnoto, M., Auerbach, R., and Kincade, P. W. (1991). A VCAM-like adhesion molecule on murine bone marrow stromal cells mediates binding of lymphocyte precursors in culture. J Cell Biol 114, 557-65. Miyake, K., Weissman, I. L., Greenberger, J. S., and Kincade, P. W. (1991). Evidence for a role of the integrin VLA-4 in lympho-hemopoiesis. J Exp Med 1 73, 599-607. 177 Miyazaki, T., Kawahara, A., Fujii, H., Nakagawa, Y., Minami, Y., Liu, Z. J ., Oishi, I., Silvennoinen, O., Witthuhn, B. A., Ihle, J. N., and et al. (1994). Functional activation of Jakl and Jak3 by selective association with IL-2 receptor subunits. Science 266, 1045-7. Miyazawa, K., Williams, D. A., Gotoh, A., Nishimaki, J ., Broxmeyer, H. E., and Toyama, K. (1995). Membrane-bound Steel factor induces more persistent tyrosine kinase activation and longer life span of c-kit gene-encoded protein than its soluble form. Blood 85, 641-9. Mombaerts, P., Iacomini, J ., Johnson, R. S., Herrup, K., Tonegawa, S., and Papaioannou, V. E. (1992). RAG-l-deficient mice have no mature B and T lymphocytes. Cell 68, 869- 77. Montague, J. W., and Cidlowski, J. A. (1995). Glucocorticoid-induced death of immune cells: mechanisms of action. Om Top Microbiol Immunol 200, 51-65. Motoyama, N., Wang, F ., Roth, K. A., Sawa, H., Nakayama, K., Negishi, 1., Senju, S., Zhang, Q., Fujii, S., and et a1. (1995). Massive cell death of immature hematopoietic cells and neurons in Bcl-x-deficient mice. Science 267, 1506-10. Muchmore, S. W., Sattler, M., Liang, H., Meadows, R. P., Harlan, J. E., Yoon, H. S., Nettesheim, D., Chang, B. S., Thompson, C. B., Wong, S. L., Ng, S. L., and Fesik, S. W. (1996). X-ray and NMR structure of human Bcl-xL, an inhibitor of programmed cell death. Nature 381, 335-41. Murti, K. G., Brown, P. S., Kurnagai, M., and Campana, D. (1996). Molecular interactions between human B-cell progenitors and the bone marrow microenvironment. Exp Cell Res 226, 47-58. Muzio, M., Chinnaiyan, A. M., Kischkel, F. C., O'Rourke, K., Shevchenko, A., Ni, J ., Scaffidi, C., Bretz, J. D., Zhang, M., Gentz, R., Mann, M., Krammer, P. H., Peter, M. E., and Dixit, V. M. (1996). FLICE, a novel FADD-homologous ICE/CED-3-like protease, is recruited to the CD95 (Fas/APO-l) death--inducing signaling complex. Cell 85, 817- 27. Nagasawa, T., Kikutani, H., and Kishimoto, T. (1994). Molecular cloning and structure of a pre-B-cell growth-stimulating factor. Proc Natl Acad Sci U S A 91 , 2305-9. Nagata, S. (1997). Apoptosis by death factor. Cell 88, 355-65. Nakayama, K., Negishi, I., Kuida, K., Shinkai, Y., Louie, M. C., Fields, L. E., Lucas, P. J ., Stewart, V., Alt, F. W., and et al. (1993). Disappearance of the lymphoid system in Bel-2 homozygous mutant chimeric mice. Science 261, 1584-8. 178 Namen, A. E., Lupton, S., Hjerrild, K., Wignall, J., Mochizuki, D. Y., Schmierer, A., Mosley, B., March, C. J ., Urdal, D., and Gillis, S. (1988). Stimulation of B-cell progenitors by cloned murine interleukin-7. Nature 333, 571-3. Namen, A. E., Schmierer, A. E., March, C. J., Overell, R. W., Park, L. S., Urdal, D. L., and Mochizuki, D. Y. (1988). B cell precursor growth-promoting activity. Purification and characterization of a growth factor active on lymphocyte precursors. J Exp Med 1 6 7, 988-1002. Nguyen, M., Millar, D. G., Yong, V. W., Korsmeyer, S. J ., and Shore, G. C. (1993). Targeting of Bcl-2 to the mitochondrial outer membrane by a COOH-terrninal signal anchor sequence. J Biol Chem 268, 25265-8. Nichols, C. T., and Tyler, F. H. (1967). Diurnal variation in adrenal cortical function. Annu Rev Med 18, 313-24. Nielsen, P. J ., Lorenz, B., Muller, A. M., Wenger, R. H., Brombacher, F ., Simon, M., von der Weid, T., Langhome, W. J ., Mossmann, H., and Kohler, G. (1997). Altered erythrocytes and a leaky block in B-cell development in CD24/I-ISA-deficient mice. Blood 89, 1058-67. Noguchi, M., Nakamura, Y., Russell, S. M., Ziegler, S. F., Tsang, M., Cao, X., and Leonard, W. J. (1993). Interleukin-2 receptor gamma chain: a functional component of the interleukin-7 receptor [see comments]. Science 262, 1877-80. Nomura, J ., Matsuo, T., Kubota, E., Kimoto, M., and Sakaguchi, N. (1991). Signal transmission through the B cell-specific MB-l molecule at the pre-B cell stage. Int Immunol 3, 117-26. Nosaka, T., Kawashima, T., Misawa, K., Ikuta, K., Mui, A. L., and Kitarnura, T. (1999). STATS as a molecular regulator of proliferation, differentiation and apoptosis in hematopoietic cells. Embo J 18, 4754-65. Nunez, G., London, L., Hockenbery, D., Alexander, M., McKeam, J. P., and Korsmeyer, S. J. (1990). Deregulated Bcl-2 gene expression selectively prolongs survival of growth factor-deprived hemopoietic cell lines. J Immunol 144, 3602-10. Nutt, S. L., Urbanek, P., Rolink, A., and Busslinger, M. (1997). Essential functions of Pax5 (BSAP) in pro-B cell development: difference between fetal and adult B lymphopoiesis and reduced V-to-DJ recombination at the IgH locus. Genes Dev 11, 476- 91. Okazaki, H., Ito, M., Sudo, T., Hattori, M., Kano, S., Katsura, Y., and Minato, N. (1989). IL-7 promotes thymocyte proliferation and maintains immunocompetent thymocytes bearing alpha beta or gamma delta T-cell receptors in vitro: synergism with IL-2. J Immunol 143, 2917-22. 179 Onate, S. A., Tsai, S. Y., Tsai, M. J., and O'Malley, B. W. (1995). Sequence and characterization of a coactivator for the steroid hormone receptor superfamily. Science 270, 1354-7. Opstelten, D., Deenen, G. J ., Rozing, J ., and Hunt, S. V. (1986). B lymphocyte-associated antigens on terminal deoxynucleotidyl transferase-positive cells and pre-B cells in bone marrow of the rat. J Immunol 13 7, 76-84. Opstelten, D., and Osmond, D. G. (1983). Pre-B cells in mouse bone marrow: immunofluorescence stathrnokinetic studies of the proliferation of cytoplasmic mu-chain- bearing cells in normal mice. J Immunol 131, 2635-40. O'Reilly, L. A., and Strasser, A. (1999). Apoptosis and autoimmune disease. Inflamm Res 48, 5-21. Osati-Ashtiani, F., King, L. E., and Fraker, P. J. (1998). Variance in the resistance of murine early bone marrow B cells to a deficiency in zinc. Immunology 94, 94-100. Osmond, D. G. (1990). B cell development in the bone marrow. Semin Immunol 2, 173- 80. Osmond, D. G. (1986). Population dynamics of bone marrow B lymphocytes. Immunol Rev 93, 103-24. Osmond, D. G., Kim, N., Manoukian, R., Phillips, R. A., Rico-Vargas, S. A., and Jacobsen, K. (1992). Dynamics and localization of early B-lymphocyte precursor cells (pro-B cells) in the bone marrow of scid mice. Blood 79, 1695-703. Osmond, D. G., Rolink, A., and Melchers, F. (1998). Murine B lymphopoiesis: towards a unified model. Immunol Today 19, 65-8. Pallard, C., Stegrnann, A. P., van Kleffens, T ., Smart, F., Venkitaraman, A., and Spits, H. (1999). Distinct roles of the phosphatidylinositol 3-kinase and STAT5 pathways in IL-7- mediated development of human thymocyte precursors. Immunity 10, 525-35. Pemberton, P. A., Stein, P. E., Pepys, M. B., Potter, J. M., and Carrel], R. W. (1988). Hormone binding globulins undergo serpin conformational change in inflammation. Nature 336, 257-8. Picard, D., Salser, S. J., and Yamamoto, K. R. (1988). A movable and regulable inactivation function within the steroid binding domain of the glucocorticoid receptor. Cell 54, 1073-80. Raff, M. C. (1992). Social controls on cell survival and cell death. Nature 356, 397-400. 180 Rast, J. P., and Litrnan, G. W. (1998). Towards understanding the evolutionary origins and early diversification of rearranging antigen receptors. Immunol Rev 166, 79-86. Rebollo, A., Gomez, J ., Martinez de Aragon, A., Lastres, P., Silva, A., and Perez-Sala, D. (1995). Apoptosis induced by IL-2 withdrawal is associated with an intracellular acidification. Exp Cell Res 218, 581-5. Reichman-Fried, M., Hardy, R. R., and Bosma, M. J. (1990). Development of B-lineage cells in the bone marrow of scid/scid mice following the introduction of functionally rearranged immunoglobulin transgenes. Proc Natl Acad Sci U S A 87, 2730-4. Rexin, M., Busch, W., Segnitz, B., and Gehring, U. (1992). Structure of the glucocorticoid receptor in intact cells in the absence of hormone. J Biol Chem 26 7, 9619- 21. Rico-Vargas, S. A., Weiskopf, B., Nishikawa, S., and Osmond, D. G. (1994). c-kit expression by B cell precursors in mouse bone marrow. Stimulation of B cell genesis by in vivo treatment with anti-c-kit antibody. J Immunol 152, 2845-52. Rolink, A., Karasuyama, H., Grawunder, U., Haasner, D., Kudo, A., and Melchers, F. (1993). B cell development in mice with a defective lambda 5 gene. Eur J Immunol 23, 1284-8. Rolink, A. G., Winkler, T., Melchers, F., and Andersson, J. (2000). Precursor B cell receptor-dependent B cell proliferation and differentiation does not require the bone marrow or fetal liver environment [see comments]. J Exp Med 191, 23-32. Rosa Santos, S. C., Dumon, S., Mayeux, P., Gisselbrecht, S., and Gouilleux, F. (2000). Cooperation between STATS and phosphatidylinositol 3-kinase in the IL-3-dependent survival of a bone marrow derived cell line. Oncogene 19, 1164-72. Sakaguchi, N., and Melchers, F. (1986). Lambda 5, a new light-chain-related locus selectively expressed 1n pre-B lymphocytes. Nature 324, 579- 82. Satterthwaite, A., and Witte, O. (1996). Genetic analysis of tyrosine kinase firnction in B cell development. Annu Rev Immunol 14, 131-54. Schatz, D. G., Oettinger, M. A., and Baltimore, D. (1989). The V(D)J recombination activating gene, RAG-l. Cell 59, 1035-48. Scherer, D. C., Brockrnan, J. A., Bendall, H. H., Zhang, G. M., Ballard, D. W., and Oltz, E. M. (1996). Corepression of RelA and c-rel inhibits immunoglobulin kappa gene transcription and rearrangement in precursor B lymphocytes. Immunity 5, 563-74. Schleirner, R. P., Claman, H. N., and Oronsky, A. (1989). Anti-inflammatory Steroid Action, lst Edition (San Diego: Academic Press, Inc.). 181 Schmidt, T. J ., Bollum, F. J ., and Litwack, G. (1982). Correlations between the activities of DNA polymerase alpha and the glucocorticoid receptor. Proc Natl Acad Sci U S A 79, 4555-9. Schuler, W., Weiler, I. J ., Schuler, A., Phillips, R. A., Rosenberg, N., Mak, T. W., Kearney, J. F ., Perry, R. P., and Bosma, M. J. (1986). Rearrangement of antigen receptor genes is defective in mice with severe combined immune deficiency. Cell 46, 963-72. Schwartzman, R. A., and Cidlowski, J. A. (1993). Apoptosis: the biochemistry and molecular biology of programmed cell death. Endocr Rev 14, 133-51. Scott, E. W., Simon, M. C., Anastasi, J., and Singh, H. (1994). Requirement of transcription factor PU.l in the development of multiple hematopoietic lineages. Science 265, 1573-7. Selye. (1947). Textbook of Endocrinology. In Acta Endocrinology, U. 0. Montreal, ed. (Montreal, pp. 837-866). Sen, R., and Baltimore, D. (1986). Inducibility of kappa immunoglobulin enhancer- binding protein Nf-kappa B by a posttranslational mechanism. Cell 4 7, 921-8. Sha, W. C., Liou, H. C., Tuomanen, E. 1., and Baltimore, D. (1995). Targeted disruption of the p50 subunit of NF-kappa B leads to multifocal defects in immune responses. Cell 80, 321-30. Shinkai, Y., Rathbun, G., Lam, K. P., Oltz, E. M., Stewart, V., Mendelsohn, M., Charron, J., Datta, M., Young, F., Stall, A. M., and et al. (1992). RAG-Z-deficient mice lack mature lymphocytes owing to inability to initiate V(D)J rearrangement. Cell 68, 855-67. Smithson, G., Medina, K., Ponting, 1., and Kincade, P. W. (1995). Estrogen suppresses stromal cell-dependent lymphopoiesis in culture. J Immunol 155, 3409-17. Sohur, U. S., Chen, C. L., Hicks, D. J., Yull, F. E., and Kerr, L. D. (2000). Nuclear factor-kappaB/Rel is apoptogenic in cytokine withdrawal-induced programmed cell death. Cancer Res 60, 1202-5. Srinivasula, S. M., Ahmad, M., Femandes-Alnemri, T., and Alnemri, E. S. (1998). Autoactivation of procaspase-9 by Apaf-l-mediated oligomerization. Mol Cell 1, 949-57. Strasser, A., Whittingharn, S., Vaux, D. L., Bath, M. L., Adams, J. M., Cory, S., and Harris, A. W. (1991). Enforced BCL2 expression in B-lymphoid cells prolongs antibody responses and elicits autoimmune disease. Proc Natl Acad Sci U S A 88, 8661-5. 182 Sudo, T., Ito, M., Ogawa, Y., Iizuka, M., Kodarna, H., Kunisada, T., Hayashi, S., Ogawa, M., Sakai, K., and Nishikawa, S. (1989). Interleukin 7 production and function in stromal cell-dependent B cell development. J Exp Med 170, 333-8. Sudo, T., Nishikawa, S., Ohno, N., Akiyama, N., Tamakoshi, M., and Yoshida, H. (1993). Expression and fiinction of the interleukin 7 receptor in murine lymphocytes. Proc Natl Acad Sci U S A 90, 9125-9. Tannenbaum, B., Rowe, W., Sharma, S., Diorio, J ., Steverman, A., Walker, M., and Meaney, M. J. (1997). Dynamic variations in plasma corticosteroid-binding globulin and basal HPA activity following acute stress in adult rats. J Neuroendocrinol 9, 163-8. Tartaglia, L. A., Ayres, T. M., Wong, G. H., and Goeddel, D. V. (1993). A novel domain within the SS kd TNF receptor signals cell death. Cell 74, 845-53. Telford, W. G., King, L. E., and Fraker, P. J. (1991). Evaluation of glucocorticoid- induced DNA fragmentation in mouse thymocytes by flow cytometry. Cell Prolif 24, 447-59. Telford, W. G., King, L. E., and Fraker, P. J. (1994). Rapid quantitation of apoptosis in pure and heterogeneous cell populations using flow cytometry. J Immunol Methods 1 72, l-16. Thomis, D. C., Lee, W., and Berg, L. J. (1997). T cells from Jak3-deficient mice have intact TCR signaling, but increased apoptosis. J Immunol 159, 4708-19. Torchia, J., Rose, D. W., Inostroza, J., Karnei, Y., Westin, S., Glass, C. K., and Rosenfeld, M. G. (1997). The transcriptional co-activator p/CIP binds CBP and mediates nuclear-receptor function [see comments]. Nature 387, 677-84. Touitou, Y., Auzeby, A., and Bogdan, A. (1990). Cortisol and cortisone production in rat and mouse adrenal incubations. J Steroid Biochem Mol Biol 37, 279-84. Tsai, M. J ., and O'Malley, B. W. (1994). Molecular mechanisms of action of steroid/thyroid receptor superfamily members. Annu Rev Biochem 63, 451-86. Urbanek, P., Wang, Z. Q., F etka, I., Wagner, E. F., and Busslinger, M. (1994). Complete block of early B cell differentiation and altered patterning of the posterior midbrain in mice lacking PaxS/BSAP [see comments]. Cell 79, 901-12. Vacchio, M. S., and Ashwell, J. D. (1997). Thymus-derived glucocorticoids regulate antigen-specific positive selection. J Exp Med 185, 2033-8. Vacchio, M. S., Ashwell, J. D., and King, L. B. (1998). A positive role for thyrnus- derived steroids in formation of the T-cell repertoire. Ann N Y Acad Sci 840, 317-27. 183 Vacchio, M. S., Papadopoulos, V., and Ashwell, J. D. (1994). Steroid production in the thymus: implications for thymocyte selection. J Exp Med 179, 1835-46. Valenzona, H. O., Dhanoa, S., Finkelrnan, F. D., and Osmond, D. G. (1998). Exogenous interleukin 7 as a proliferative stimulant of early precursor B cells in mouse bone marrow: efficacy of IL—7 injection, IL-7 infusion and IL-7-anti-IL-7 antibody complexes. Cytokine 10, 404-12. Vander Heiden, M. G., Chandel, N. S., Williamson, E. K., Schurnacker, P. T., and Thompson, C. B. (1997). Bcl-xL regulates the membrane potential and volume homeostasis of mitochondria [see comments]. Cell 91 , 627-37. Vaux, D. L., Weissman, I. L., and Kim, S. K. (1992). Prevention of programmed cell death in Caenorhabditis elegans by human bcl-2. Science 258, 1955-7. Veis, D. J., Sorenson, C. M., Shutter, J. R., and Korsmeyer, S. J. (1993). Bcl-2-deficient mice demonstrate fulminant lymphoid apoptosis, polycystic kidneys, and hypopigrnented hair. Cell 75, 229-40. Vinson, G. P., Laird, S. M., Hinson, J. P., and Teja, R. (1992). Origin of aldosterone in trypsin-stirnulated rat adrenal zona glomerulosa incubations. J Endocrinol 135, 125-33. Vogeser, M., Felbinger, T. W., Kilger, E., Roll, W., Fraunberger, P., and Jacob, K. (1999). Corticosteroid-binding globulin and free cortisol in the early postoperative period after cardiac surgery. Clin Biochem 32, 213-6. von Freeden-Jeffry, U., Solvason, N., Howard, M., and Murray, R. (1997). The earliest T lineage-committed cells depend on IL-7 for Bel-2 expression and normal cell cycle progression. Immunity 7, 147-54. von Freeden-Jeffry, U., Vieira, P., Lucian, L. A., McNeil, T., Burdach, S. E., and Murray, R. (1995). Lyrnphopenia in interleukin (IL)-7 gene-deleted mice identifies IL-7 as a nonredundant cytokine. J Exp Med 181, 1519-26. Waggie, K. S., Kahle, P. J ., and Tolwani, R. J. (1999). Neurons and mechanisms of neuronal death in neurodegenerative diseases: a brief review. Lab Anim Sci 49, 358-62. Wagner, R. L., Apriletti, J. W., McGrath, M. E., West, B. L., Baxter, J. D., and Fletterick, R. J. (1995). A structural role for hormone in the thyroid hormone receptor. Nature 378, 690-7. Wang, H. G., Miyashita, T., Takayama, S., Sato, T., Torigoe, T., Krajewski, S., Tanaka, S., Hovey, L., 3rd, Troppmair, J., Rapp, U. R, and et al. (1994). Apoptosis regulation by interaction of Bel-2 protein and Raf-1 kinase. Oncogene 9, 2751-6. 184 Wang, J ., Lin, Q., Wu, Q., and Cooper, M. D. (1998). The enigmatic role of glutamyl arninopeptidase (BP-1/6C3 antigen) in immune system development. Immunol Rev 161, 71-7. Wang, J. H., Nichogiarmopoulou, A., Wu, L., Sun, L., Sharpe, A. H., Bigby, M., and Georgopoulos, K. (1996). Selective defects in the development of the fetal and adult lymphoid system in mice with an Ikaros null mutation. Immunity 5, 537-49. Wang, L., Chen, J. J ., Gelman, B. B., Konig, R., and Cloyd, M. W. (1999). A novel mechanism of CD4 lymphocyte depletion involves effects of HIV on resting lymphocytes: induction of lymph node homing and apoptosis upon secondary signaling through homing receptors. J Immunol 162, 268-76. Wang, W., Wykrzykowska, J ., Johnson, T., Sen, R., and Sen, J. (1999). A NF-kappa B/c- myc-dependent survival pathway is targeted by corticosteroids in immature thymocytes. J Immunol 162, 314-22. Watson, J. D., Morrissey, P. J., Namen, A. E., Conlon, P. J., and Widmer, M. B. (1989). Effect of IL-7 on the grth of fetal thymocytes in culture. J Immunol 143, 1215-22. Webster, J. C., Jewell, C. M., Bodwell, J. E., Munck, A., Sar, M., and Cidlowski, J. A. (1997). Mouse glucocorticoid receptor phosphorylation status influences multiple functions of the receptor protein. J Biol Chem 272, 9287-93. Westerrnann, J ., Aicher, A., Qin, Z., Cayeux, Z., Daemen, K., Blankenstein, T., Dorken, B., and Pezzutto, A. (1998). Retroviral interleukin-7 gene transfer into human dendritic cells enhances T cell activation. Gene Ther 5, 264-71. Westphal, U., and Devenuto, F. (1966). Steroid-protein interactions. XI. Electrophoretic characterization of corticosteroid-binding proteins in serum of rat, man and other species. Biochim Biophys Acta 115, 187-96. Whitlock, C. A., Robertson, D., and Witte, O. N. (1984). Murine B cell lymphopoiesis in long term culture. J Immunol Methods 6 7, 353-69. Whitlock, C. A., and Witte, O. N. (1982). Long-term culture of B lymphocytes and their precursors from murine bone marrow. Proc Natl Acad Sci U S A 79, 3608-12. Wienands, J. (2000). The B-cell antigen receptor: formation of signaling complexes and the function of adaptor proteins. Curr Top Microbiol Immunol 245, 53-76. Witte, O. N. (1990). Steel locus defines new multipotent growth factor [published erratum appears in Cell 1990 Nov 30;63(5):following 1112]. Cell 63, 5-6. 185 Witte, P. L., Frantsve, L. M., Hergott, M., and Rahbe, S. M. (1993). Cytokine production and heterogeneity of primary stromal cells that support B lymphopoiesis. Eur J Immunol 23, 1809-17. Witte, P. L., Robinson, M., Henley, A., Low, M. G., Stiers, D. L., Perkins, S., Fleischman, R. A., and Kincade, P. W. (1987). Relationships between B-lineage lymphocytes and stromal cells in long-term bone marrow cultures. Eur J Immunol 1 7, 1473-84. Wu, Q., Li, L., Cooper, M. D., Pierres, M., and Gorvel, J. P. (1991). Aminopeptidase A activity of the murine B-lymphocyte differentiation antigen BP-l/6C3. Proc Natl Acad Sci U S A 88, 676-80. Wyllie, A. H. (1980). Glucocorticoid-induced thymocyte apoptosis is associated with endogenous endonuclease activation. Nature 284, 555-6. Xu, Y., and Lindquist, S. (1993). Heat-shock protein hsp90 governs the activity of pp60v-src kinase. Proc Natl Acad Sci U S A 90, 7074-8. Yamaguchi, K., Near, R. I., Matulka, R. A., Shneider, A., Toselli, P., Trombino, A. F ., and Sherr, D. H. (1997). Activation of the aryl hydrocarbon receptor/transcription factor and bone marrow stromal cell-dependent preB cell apoptosis. J Immunol 158, 2165-73. Yang, E., Zha, J., Jockel, J., Boise, L. H., Thompson, C. B., and Korsmeyer, S. J. (1995). Bad, a heterodimeric partner for Bcl-XL and Bel-2, displaces Bax and promotes cell death. Cell 80, 285-91. Yuan, J ., Shaham, S., Ledoux, S., Ellis, H. M., and Horvitz, H. R. (1993). The C. elegans cell death gene ced-3 encodes a protein similar to mammalian interleukin-1 beta- converting enzyme. Cell 75, 641-52. Zha, J., Harada, H., Yang, E., Jockel, J., and Korsmeyer, S. J. (1996). Serine phosphorylation of death agonist BAD in response to survival factor results in binding to 14-3-3 not BCL-X(L) [see comments]. Cell 87, 619-28. Zou, H., Henzel, W. J., Liu, X., Lutschg, A., and Wang, X. (1997). Apaf-l, a human protein homologous to C. elegans CED-4, participates in cytochrome c-dependent activation of caspase-3 [see comments]. Cell 90, 405-13. 186 lulivll'lllllflilfill 49 02112 57 11111111111111 3 1293