NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.
Madame Curie Bioscience Database [Internet]. Austin (TX): Landes Bioscience; 2000-2013.
Introductory Remarks
The haematopoietic system produces a wide variety of new functional cells as they are needed, including monocytes, macrophages and myeloid dendritic cells. Macrophages and dendritic cells are spiised for phagocytosis, tissue remodelling and antigen processing and presentation, as we will discuss later. They are derived from primitive precursor cells that can be grown in vitro or in vivo, and which replenish the entire haematopoietic compartment throughout life.
This chapter reviews our current knowledge of the role of cytokines and their receptors in the proliferation and differentiation of pluri-potent stem cells (PPSC) as they develop into cells of the monocyte/macrophage/dendritic cell lineage. Detailed discussion is therefore limited to the roles of CSF-1, SCF, FL, GM-CSF, IL-3, TNF, IFN-gamma, and IL-4, and to a lesser extent IL-6 and TPO. We have concentrated on experiments that use gene manipulation technology to unravel pathways of activation of cellular function and proliferation.
Much of our knowledge of haematopoiesis comes from mouse models. Till and McCulloch1-3 were the first to show that individual bone marrow cells could form colonies in the spleens of irradiated recipient mice, and that these colonies also contained transplantable colony forming cells as well as all the other cellular elements of blood. In spite of this early observation most of what we know about pluri-potent stem cells of blood is inferred from analysis of their progeny, since the PPSC is present in very small numbers and until recently was difficult to maintain in vitro.4
Mouse PPSC are capable of the long-term (greater than six months) repopulation of both myeloid and lymphoid systems in lethally irradiated recipient mice. PPSC represent only 10-5 of the nucleated bone marrow cell pool, and in the steady state they either do not cycle or they cycle very slowly.5,6 During times of demand, and under the control of a number of cytokines, PPSC proliferate and differentiate to give rise to common myeloid7 and common lymphoid8 progenitor cells that become increasingly more committed to a particular lineage until they can only form one mature type of blood cell.9 Progenitor cells can be detected in in vitro assays where they form colonies in a soft agar matrix under the influence of a particular group of cytokines called colony-stimulating factors. Initially the ‘colony assay’ was used to detect macrophage and granulocyte colonies.10,11 It was shown that colony numbers were linearly related to the number of cells cultured and, at constant cell numbers, the number of colonies generated was proportional to the concentration of cytokine present. This permitted the use of in vitro colony assays in the purification of cytokines that stimulated colony formation,12-14 and in the molecular cloning of the genes that encoded them.
Sources of Cytokines and Regulation of Their Production
Haematopoietic cytokines are glycoproteins that are normally present in the circulation (CSF-1, SCF, FL, G-CSF, EPO and TPO) in very low (pico-molar) doses, or are induced to appear in response to infection or inflammation (GM-CSF, IL-3, IL-5, IL-6, and IL-11). Some are induced in response to particular stimuli e.g., IL-3 is induced almost entirely by antigen stimulation of T cells. Cytokines contain carbohydrate moieties that are not required for activity but which seem to affect their half-life in the circulation and their localisation in vivo. The crystal structures of many cytokines have been determined including those of CSF-1,15 GM-CSF,16 G-CSF,17 IL-218,19 IL-320 and IL-6.21 Surprisingly the tertiary structures of these very disparate primary amino acid chains are rather similar. In general, like growth hormone,22 they contain four α-helical bundles and an antiparallel β ribbon. Some are homo-dimers (SCF, CSF-1, FL and IL-5) but others are monomeric. SCF, CSF-1 and FL are expressed as membrane spanning forms that are active at the cell surface and are involved in local regulation.
Signalling
Cytokines exert their action through high-affinity receptors on the cell surface that are linked to pathways of cellular activation, survival, proliferation and differentiation. Cross-linking of receptor subunits on the outside of the cell wall leads to abutting of kinases associated with the intracellular receptor tails, either as intrinsic activities, or because of pre-association of secondary kinase molecules. This intracellular association of signalling molecules results in phosphorylation of tyrosine residues in the receptor tail and binding of further signalling molecules that have phospho-tyrosine-binding domains. Several aspects of the downstream intracellular pathways of cytokines are similar, because different activated receptor cytoplasmic domains often bind a common signalling molecule or family of signalling molecules. Thus stoichiometry and rate of reaction are important regulatory influences that allow discrimination between signalling processes with different outcomes.
Cooperative Interaction of Cytokines in Proliferation and Differentiation
For efficient in vitro proliferation and differentiation, PPSCs and multi-potent progenitor cells require a combination of cytokines. (eg SCF, IL-1, IL-3, IL-6, GM-CSF, and CSF-1).23,24 As might be expected from this observation, immature haematopoietic cells have been shown to co-express a number of different lineage specific receptors at low levels.25 As these immature cells develop, they lose receptors for some cytokines e.g., SCF and IL-3, while retaining receptors for later acting cytokines such as CSF-1. Eventually at least some of the immature cells reach the stage of committed progenitor cell, where their further proliferation and differentiation are along one particular lineage, dictated by the relevant lineage-restricted cytokine. It may be important in lineage commitment for the level of receptors for the lineage-restricted cytokine to increase as differentiation proceeds.26-28 Single lineage cytokines (such as CSF-1) often regulate the survival of their target cells including fully differentiated end cells29,30 where they may prime the cells to receive secondary activation signals from elements such as bacterial lipo-poly-saccharide or immune complexes. Despite the apparent overlap in target cell specificity of several cytokines their functions are largely non-redundant as indicated by the distinct haematopoietic phenotypes of cytokine or cytokine-receptor deficient mice.
Synergy occurs between some late-acting lineage restricted cytokines such as CSF-1, EPO and G-CSF, with early acting cytokines such as SCF and IL-3, in stimulating the proliferation and differentiation of multi-potent cells. This provides a mechanism by which the tightly regulated changes in the level of a late acting cytokine, can be coupled to the channelling of multi-potent progenitor cells into a lineage to satisfy the demand for differentiated cells. The underlying mechanism of synergy may lie at the level of either the receptors or at the level of post receptor signalling pathways. There is little evidence that one cytokine directly influences levels of receptor for a synergistic cytokine, although it has been shown that IL-3 induces increased expression of c-fms mRNA in macrophages and bone marrow cells of A/J mice.27,31 It is not clear whether the apparent increase in mRNA is due to de novo synthesis or to stabilisation of existing messenger RNA. The signalling is unique in that it almost certainly occurs as the result of binding of IL-3 to the mouse unique IL-3 specific beta sub-unit of the IL-3 receptor.
In an interesting experiment to determine the influence of cytokines on haematopoiesis, it was shown that common lymphoid progenitors can be redirected to the myeloid lineage by stimulation through exogenously expressed IL-2 and GM-CSF receptors. Furthermore it was shown that granulocyte and monocyte differentiation signals can be regulated by different domains of the IL-2 receptor.32
The Role of Transcription Factors
Experiments using gene manipulation have shown that nuclear transcription factors are essential for stem cell lineage commitment.GATA-133 reprograms avian myelomonocytic cells from a myoblast fate to eosinophil, thromboblast and erythroblast differentiation. PU.1 on the other hand induces myeloid lineage commitment in multi-potent haematopoietic progenitors.34 Pre B cells in Pax 5 negative mice fail to develop into B cells, instead becoming progenitors of macrophages, osteoclasts, granulocytes, NK cells and T cells.35 This suggests that transcription factors may act as both positive and negative regulators of differentiation
Tyrosine Kinase Receptors
CSF-1, SCF and FL are homodimeric cytokines that share some sequence homology and structurally are similar to one another.36-40 All three cytokines have complex patterns of expression due to alternative mRNA splicing. This allows them to be expressed as either membrane-spanning, cell-surface or secreted glycoproteins. All three growth factors are widely expressed in all tissues and at least two of them, SCF and CSF-1, affect non-haematopoietic as well as haematopoietic cells.
The receptors for the above cytokines are members of the PDGF receptor family.39-42 Each receptor possesses an extra-cellular domain comprising five Ig-like repeats which are heavily glycosylated with N-linked sugars, a trans-membrane domain, and intra-cellular domains containing a juxta-membrane region, a src-related tyrosine kinase domain that is interrupted by a kinase insert domain, and a carboxy-terminal tail. The three amino terminal Ig-like domains incorporate the ligand binding domains of the SCF and CSF-1 receptors.
Binding of this type of dimeric receptor to its cognate ligand stabilises the non-covalent association between the two chains of the receptor at the cell surface and permits the trans-phosphorylation of the intra-cellular domain of one chain by the other.
Tyrosine phosphorylation in response to cytokine binding is not restricted to those proteins that are stably associated with the receptor. Regions containing tyrosines that are phosphorylated as a consequence of receptor activation act as docking sites for src homology region 2 (SH2) domains of signalling and adaptor proteins. These proteins in turn may interact with plasma membrane associated proteins. An example is the association of recruited Grb2/Sos with Ras, which leads to their activation. Or the associated proteins may themselves become tyrosine phosphorylated.
Many of the signalling pathways activated by SCF, CSF-1 and FL receptors, including the Ras/Raf-mitogen activated protein kinase cascade, the Janus kinase (JAK) signal transducers and activation of transcription (STAT) pathway, Src family members and phosphatidylinositol-3-kinase (PI3K), are shared.43 All three receptors are likely to exhibit ligand induced, Cbl-mediated decreases in receptor expression.44
The SCF and CSF-1 receptors are encoded by the proto-oncogenes c-kit and c-fms respectively. 45 The oncogenes derived from these two proto-oncogenes are present in mutated forms in retroviruses that cause sarcoma in cats. The mutations in the receptor genes cause constitutive activation of the receptors in the absence of cytokines.45 thus contributing to unregulated cell proliferation.
Cytokines and Receptors with a Significant Effect on Cells of the Monocyte Macrophage Lineage
FL and FLT3 Receptor
The receptor flt3 was isolated using a cloning approach aimed at identifying new receptor tyrosine kinase genes. The ligand FL was then discovered due to its ability to bind soluble flt3.46-49
There are several isoforms of flt3. In humans the predominant iso-form is a membrane spanning glycoprotein that is biologically active. The extracellular domain of this form may be released by proteolysis. In mice, the predominant iso-form is cell-surface associated and does not span the membrane. A third iso-form found in both mice and humans arises from alternative splicing of exon 6, introducing a stop codon near the end of the extracellular domain. This generates a relatively rare secreted, but biologically active protein. The similarity of FL to CSF-1, especially when comparing the amino terminal 150 amino acids that are essential for biological activity suggests that the tertiary structures of these molecules may be similar. FL exists as a non-covalently bonded homo-dimer.48
Within haematopoietic cells, expression of flt3 is largely restricted to the progenitor cell pool. Bone marrow monocytes and a fraction of lymphocytes express flt3 but the functional significance of this has yet to be established. One iso-form of the mouse flt3, missing the fifth Ig-like motif in the extracellular domain, is still able to bind ligand and to become tyrosine phosphorylated.49 Again the significance of this is not presently understood.
FL is not itself mitogenic for haematopoietic progenitor cells but like SCF it can act in synergy with other cytokines. The effects do not appear to be on erythroid or megakaryocyte progenitor cells since FL receptor-nullizygous mice appear to have no defects in red blood cell, megakaryocyte or platelet production. However FL- and flt3- deficient mice have reduced numbers of pro B cells although they have normal numbers of mature B cells.50,51
FL synergizes with IL-7 to stimulate proliferation of B cell progenitor cells and thymic precursor cells and with IL-15 to promote the expansion of natural killer (NK) cells. FL-deficient mice fail to develop splenic NK cell activity and have reduced dendritic cell (DC) numbers in other tissues.50
In combination with GM-CSF, tumour necrosis factor and IL-4, FL enhances the production of DC from bone marrow progenitor cells. FL dramatically increases the numbers of DC in lympho-haemopoietic and other tissues and dendritic cell numbers are reduced in FL-deficient mice.51
CSF-1
CSF-1 regulates the survival, function, proliferation and differentiation of cells of the mononuclear phagocyte series, as well as the function of cells of the female reproductive tract.40,52-54 The 150 amino terminal amino acids of CSF-1 are sufficient for biological activity. This section contains the four alpha helical bundle-anti-parallel beta sheet structure.15 and is expressed as an end to end disulfide-linked dimer. It is synthesised in the endoplasmic reticulum as a disulphide linked homo-dimeric, membrane-spanning precursor containing 522 amino acids. In secretory vesicles the mature proteoglycan (120 — 160 kD) and glycoprotein (100kD) are cleaved from this precursor and secreted.55 A smaller CSF-1 glycoprotein precursor of 224 amino acids from which the proteolytic cleavage and glycosaminoglycan sites have been due to alternative splicing is expressed on the cell surface when the secretory vesicle fuses with the cell membrane.
The 58 kb gene encoding the CSF-1 R is known as c-fms and maps to the long arm of chromosome 545 near to the genes encoding GM-CSF and IL-3. When ligand binds the CSF-1 receptor dimerization occurs and the cytoplasmic tails of the molecule trans-phosphorylate on tyrosine. This results in a cascade by which several cytoplasmic proteins, some of which have been shown to play a role in signal transduction, are phosphorylated.55,56 After activation and signalling the receptor ligand complexes are internalised and destroyed within lysosomes.44,56-58
CSF-1 is synthesised by many cell types, including fibroblasts, endothelial cells, bone marrow stromal cells, osteoblasts, keratinocytes, astrocytes, myoblasts, and uterine epithelial cells in pregnancy.40 Circulating CSF-1 is probably made by endothelial cells of small blood vessel walls. Ninety five percent of circulating CSF-1 is cleared by sinusoidal macrophages such as Kupffer cells, through CSF-1R-mediated endocytosis and intra-lysosomal destruction. The number of sinusoidal macrophages serves as a simple feed-back control on the circulating level of CSF-1.58,59
The half-life of circulating CSF-1 at physiological concentrations is about ten minutes but at saturating pharmacological concentrations the half-life can extend to 1.6 hours.57 Bacterial endotoxin, viruses, and parasitic infections raise the levels of circulating CSF-1.40,52-54 Repeated injection of CSF-1 into mice can cause a tenfold increase in circulating monocytes and an increase in macrophages in the periphery.60
Osteopetrotic (Csf1op/Csf1op) mice have an inactivating mutation in the CSF-1 gene.61,62 These mice display impaired bone resorption associated with a paucity of osteoclasts, the production of which requires both osteoprotergerin ligand and CSF-1. They have no incisors, poor fertility, a lower body weight, absence of evoked auditory and visual responses, a thinner than normal dermis, a deficiency in blood monocytes and in macrophages in certain tissues and a shorter life expectancy.63,64
Restoration of circulating CSF-1 in Csf-1op/CSF-1op mutant mice cured their osteopetrosis and monocytopenia and caused increases in some but not all tissue macrophage populations. This is consistent with local as well as humoral regulation by CSF-1.63-65 There was complete rescue of these phenotypes by normal tissue specific and developmental expression of a transgene encoding all three forms of CSF-1 but only partial rescue by a transgene encoding the cell surface iso-form.66,67 The two secreted forms of CSF-1 are both found in the circulation. Both the cell surface and secreted proteoglycan CSF-1 variants are thought to play a role in local regulation in part because the proteoglycan form is sequestered by specific extracellular matrices.40,45
Studies with CSF-1op/CSF-1op and Csf1r-/Csf1r- mice clearly indicate that CSF-1 is the primary regulator of the monocyte macrophage series, including osteoclasts. The pleiotropic phenotype of Csf1op /Csf1op mice, and the restricted distribution of CSF-1R to cells of the mononuclear phagocyte system and female reproductive tract, suggest that the major role of CSF-1 is to generate and maintain macrophages that have trophic, as well as scavenger functions that are important in organogenesis and tissue turnover.65 Csf1op/Csf1op mice exhibit normal T- and B-cell-dependent responses to simple and corpuscular antigens.68,69 However, CSF-1 has a well-documented priming role in macrophage activation and it plays an important role in innate immune responses.76 It appears that the development of macrophages involved in inflammatory and immunologic functions depends on other cytokines that regulate the activities of macrophages generated by CSF-1.70,71
The CSF-1 R is expressed at low levels on early multi-potential cells. The level of CSF-1R increases tenfold as these cells give rise to committed macrophage precursors.26,72 There is some evidence that this increase can be mediated by IL-3.27 In vitro experiments have shown that in contrast to committed progenitors the multi-potent cells are unable to respond to CSF-1 alone, but CSF-1 can synergise with IL-1, SCF, IL-3 and IL-6 to stimulate multi-potential cell proliferation and differentiation to committed macrophage progenitor cells.23,24,26,71-73
As well as its role in macrophage development, CSF-1 is involved in bone metabolism through its role in the proliferation and differentiation of osteoclast precursors. CSF-1 is synthesised locally by osteoblasts and synergizes with osteoprotergerin ligand in the maturation of osteoclasts from small mononuclear cell precursors that proliferate in response to CSF-1 alone.74 CSF-1 also plays and important role in both male and female reproductive systems75 and the action of CSF-1 on the trophoblast is critical for embryonic resistance to certain infections. 76 Under control of the female endocrine system, CSF-1 is synthesised in the oviduct and the uterine epithelium during pregnancy and can influence several cell populations including maternal macrophages and decidual cells, as well as early embryonic cells and various trophoblast cell types.40,60 Circulating levels of CSF-1 are elevated in patients with myeloid and lymphoid malignancies or with carcinomas of the ovary, breast and endometrium. In ovarian cancer elevated levels of circulating CSF-1 are associated with a poor prognosis77,78 IL-3 and GM-CSF.
IL-3, GM-CSF and IL-5 are cytokines that signal through receptors that contain a common beta sub-unit.79 Although their amino acid sequences share little homology, it appears that IL-3, IL-5 and GM-CSF share a common ancestral gene. All three genes map to in close proximity on the long arm of chromosome 5 and have similar structure. The tertiary protein structure for all three is similar, with four alpha helical bundles and antiparallel beta ribbons except that IL-5 is a covalently disulphide linked dimer. All three cytokines signal through a receptor that contains an alpha chain that exhibits low affinity for the cytokine and a common beta chain. When cytokine-specific alpha chain and beta chains join, a high affinity receptor complex capable of signalling is formed. In mice there is a second form of the beta chain which appears to be IL-3 specific and which binds IL-3 with low affinity. This variant beta chain may be able to signal in the absence of alpha chain because in some strains of mice the alpha chain is missing and yet gene expression but not proliferation can be induced in bone marrow cells of these mice by IL-3.27 The affinity of binding of the IL-3 specific beta chain is contentious as is the availability of complementary alpha chain in mutant mice.
Ligand-induced disulphide bonding between alpha and beta sub units as well as dimerization of beta subunits are required for signalling (REF). In strains in which the alpha chain is missing, possibly beta chain dimerization is sufficient. The high affinity cytokine-receptor complex resembles the IL-6/IL-6R high affinity complex, consisting of two receptor alpha chains, two beta chains and two ligand molecules.79 Because of the shared beta sub-unit, IL-3, IL-5 and GM-CSF compete with each other for binding to their high affinity receptors,80 though in mice IL-3 can to some extent circumvent this because of the IL-3 unique variant of the beta chain. The biological significance of the competition between IL-3, IL-5 and GM-CSF for the shared beta sub-unit is not understood. However, it is clear that the cytoplasmic domains for both alpha and beta receptor subunits can be involved in signalling.80
The 120kD beta sub-unit contains the conserved cytokine motif found in the alpha sub-unit. It has a much longer cytoplasmic tail which is required for proliferative signalling and it has box I and box 2 motifs that contain docking sites required for the recruitment of signalling molecules by the activated receptor. Activation of the receptor complex results in tyrosine phosphorylation of the receptor and recruited molecules.
Interleukin-3
Interleukin-3 is a secreted monomeric 133 amino acid 25-30kD glycoprotein containing an intramolecular disulphide bridge that stabilizes the loop containing the binding site. The IL-3R comprises a low affinity alpha sub-unit and a beta sub-unit in the human but, as indicated above, in the mouse there is also a low affinity IL-3 specific beta sub-unit which interacts in a productive manner only with the IL-3 specific alpha chain.79-83
IL-3 is synthesised almost exclusively by T cells in response to antigenic stimulation.82 It is a pleiotropic cytokine supporting the proliferation and differentiation of multi-potent progenitor cells and committed progenitors.83-84
IL-3 alone stimulates multi-potent stem cells to form colonies containing neutrophils, basophils, monocytes, eosinophils and megakayocytes. Yet more primitive precursors are stimulated by IL-3 in the presence of CSF-183 in populations enriched for haematopoietic stem cells. IL-3 nullizygous mice have diminished DTH but no obvious steady state haematopoietic phenotype. 82 This suggests that IL-3 may be involved only when there is an unusual demand for increased haematopoiesis such as eventuates in an immune response.
Granulocyte-Macrophage Colony Stimulating Factor
GM-CSF is a glycoprotein of between 18 and 32 kD.85 It has a receptor similar to that of IL-3. The alpha sub-unit of the receptor is a membrane spanning protein of 80 kD containing the extracellular cytokine-binding domain. There are three alternative transcripts, one for the main form just described, one encoding a soluble form that has the region encoding the trans-membrane and cytoplasmic region spliced out, and a third species encoding an alternative membrane spanning alpha chain with a slightly longer c-terminal end.86 All bind ligand but their physiological significance is poorly understood.
GM-CSF is constitutively synthesised by macrophages, endothelial cells and fibroblasts and its expression is inducible in T cells.87 The cytokine acts as survival molecule and as a mitogen on progenitor cells of many lineages.88 including all the cells of the neutrophil, macrophage and eosinophil lineages. At most stages the cells of these lineages also require more lineage specific factors such as G-CSF, CSF-1 and IL-5. GM-CSF synergises with EPO and TPO to generate erythroid and megakaryocyte progeny. In vivo GM-CSF acts to increase the number of circulating neutrophils, monocytes and eosinophils and the number of fixed tissue macrophages. GM-CSF-deficient mice have pulmonary alveolar proteinosis caused by a deficiency of alveolar macrophages.89 However these mice have normal granulocytes and macrophage numbers in the steady state, thus GM-CSF apparently does not play an indispensable role in blood cell production.
Interleukin-4
Is a glycoprotein of 15 to 19 kD. There is a splice variant lacking amino acids 22 to 37 that is expressed more strongly in thymocytes. This IL- 4d2 inhibits IL-4 induced T cell proliferation. The IL-4 R is expressed on a wide range of cell types including B and T lymphocytes, monocytes, granulocytes, fibroblasts, endothelial and epithelial cells and expression is induced by IL-4.90
Two forms of IL-4 R are known, one comprising the IL-4Rβ and γc subunits requires JAK3 for STAT activation is predominantly expressed in haematopoietic cells. The other comprising the IL-4β and IL-13Rα subunits requires JAK2 for STAT 6 activation and is predominantly expressed in non hematopoietic cells.90 A soluble form of IL-4 receptor binds IL-4 with high affinity.91
IL-4 is expressed in TH2 cells. It has complex biological actions many of which may be indirect through the modulation of cytokine production.90-92 IL-4 enhances antigen-presenting activity of B cells to T cells. It enhances antigen receptor and CD40 triggering of B cell proliferation and differentiation, but it antagonises IL-2-induced co-stimulation, possibly because it sequesters the γc chain, shared with the IL-2 receptor. Normal B cell development is seen IL-4 nullizygous mice suggesting that some of its in vitro effects can be subsumed by other factors in vivo.92 IL-4 stimulates the generation of TH2 cells and the proliferation and maturation of thymocytes, while in contrast it inhibits early T cell development in foetal thymus organ culture. IL-4-nullizygous mice have a TH2 cell deficiency and diminished IgG1 and IgE responses consistent with a role for IL-4 in regulation of T helper cell development and in Ig switching.90-94
IL-4 inhibits CSF-1 induced macrophage and megakaryocyte colony formation, probably by accelerating terminal cell differentiation, but enhances G-CSF induced colony formation and, with IL-3, the generation of basophils and eosinophils.
Interferon-γ
A non-covalent homo-dimeric glycoprotein of 50 kD is produced by activated or cytokine-stimulated T lymphocytes and NK cells. Interferon-γ has a receptor composed of two or three subunits. The primary receptor contains a ubiquitously expressed alpha chain, an inducible beta chain, and in some circumstances, a third sub-unit is required for signalling.95 IFN-γ deficient mice have defects in the immune system that become apparent following infection suggesting that these mice fail to deal with infectious agents effectively at the phagocytic level. IFN augments the immune response by activating macrophages toward phagocytosis, and enhancing antigen recognition by increasing the expression of class one and class two antigens on antigen processing DCs. IFN is also a general inhibitor of cell proliferation by a variety of mechanisms.96
Tumour Necrosis Factor
Tumour necrosis factor alpha and lympho-toxin alpha (TNF-beta) are products of activated macrophages and T cells. Receptors for these mediators are ubiquitous, serving important functions in sepsis, granulomatous disease, autoimmunity, graft versus host disease and other inflammatory events particularly involving the activation of macrophages or neutrophils. TNF-α deficient mice have a disorganised splenic architecture and peripheral lymph node deficiencies, and have a reduced capacity to fight infection.97,98 The TNFs induce production of myeloid CSFs, IFN-γ, IL-6 and chemokines by various cell types and they inhibit production of TGF-β. The actions of the TNFs can therefore be quite complex, particularly in the regulation of antigen processing cells (APC's).
Antigen Processing Cells
Recently, renewed effort has been directed towards establishing the role of APC's in immune regulation.99 The best understood APC's are macrophages, Langerhan's cells (LC), myeloid dendritic cells (MDC) and lymphoid dendritic cells (LDC).100
Macrophages are M-CSF (CSF-1) responsive, strongly adherent, highly phagocytic and express cell-surface F4/80,101 Mac-1 and cell surface Fms.102 They are derived from blood monocytes and reside in discrete sites as tissue macrophages.65 As monocytes they are highly mobile and can migrate readily into sites of inflammation. Macrophages are subject to regulation by a wide variety of stimuli that affect their maturation and function. They are important in tissue remodelling and removal of debris including opsonised micro-organisms. The role of the classical macrophage in immune induction is contentious. Macrophages present antigen poorly and in fact may be immuno-suppressive.103 It is generally thought that only cells which express an intact antigen processing and presentation pathway, including an active TAP system and appropriate MHC class II antigens, as well as other required co-stimulatory molecules, are effective in antigen presentation.70-71 These are properties not generally seen in macrophages, though activation of some phagocytes by IFN-γ for example, induces expression of some of these molecules.94 The phagocytes so induced may be more closely related to DC than to classical macrophages as we shall discuss later.
MDC and LC take up antigen in the periphery by pinocytosis or phagocytosis then migrate to T-cell areas of lymphoid organs such as spleen, lymph nodes and Peyer's patches. LC ingest antigens in the dermal region then migrate via afferent lymphatics to local lymph nodes.105 Studies of mouse.106 and human107,108 LC and MDC have shown that both cell types pass through an immature stage where they express receptors that enhance antigen capture and adsorption such as the macrophage mannose receptor107 and Fcγ and Fcε receptors.108 At this stage MDC and LC lack the cell surface molecules CD40, CD54 and CD86,100 which are important for the delivery of activation signals to naive T cells. After DC have taken up antigen, it is transported to a particular type of late endosomal compartment which is rich in MHC Class II.108-111 In this compartment the antigen is digested into peptides and loaded onto MHC class II molecules. When microbial or inflammatory products stimulate DC to mature, the DC migrate to T cell areas of lymphoid tissue and the antigen-loaded MHC class II molecules are passed to the cell surface for presentation to a re-circulating pool of T cells.111,112 T cell clones with particular antigen specificities are then activated and expanded to achieve effector or helper cell status.
LDC are different from MDC in that they secrete much higher levels of IL-12, and they are located in the T cell areas of the spleen whereas MDC are found in the marginal zone.113 In addition, LDC are long lived cells whose MHC molecules tend to carry peptides derived from self antigens.100 It has been suggested that unlike MDC which specialise in triggering T cell immune responses against foreign antigens such as those of pathogenic micro-organisms. LDC may also be required for maintaining self tolerance.114,115
Classification and Sources of Antigen Processing Cells
MDC can be isolated from blood, lymph nodes, liver, bone marrow, cord blood and thymus either as mature DC, monocytes, or CD34+ myeloid precursor cells. DC progenitor cells can then be stimulated to differentiate into DC by the addition of various combinations of cytokines and growth factors,116 Some of these stimuli in mouse systems include GM-CSF,117 GM-CSF and IL-1,118 GM-CSF, SCF and TNF-α,119 and in human systems include IL-4 and IFN-γ,120 IL-4 and GM-CSF,107,120-122 IL-4, GM-CSF and IFN-γ,123-124 TNF-α and GM-CSF,125 TNF-α, GM-CSF and CD40L.126
MDC are derived from the same GM-CSF-responsive progenitor cell pool in bone marrow that gives rise to macrophages and LC's.99,127-132 Human CD34+ blood cells also contain a macrophage, DC and LC progenitor cell.130-133 The point at which the LC and DC diverge during maturation in vitro appears to be after CD34+ progenitor cells are exposed to GM-CSF and TNF-α. This results in two populations of cells, one of which is CD1a+ and the other CD14+.100 The CD1a+ cells differentiate into LC in response to further culture in the presence of GM-CSF and TNF-α.134 The CD14+ cells are bi-potent maturing into DC in response to GM-CSF and TNF-α (or other cytokine combinations described above), or macrophages in response to M-CSF.
Currently MDC are characterized by the expression of CD1a, CD40, CD80, CD86 and MHC class II cell surface antigens, they endocytose FITC dextran, and stimulate T cells in a mixed lymphocyte reaction.70,71 LC are distinguished from MDC by expression of CLA,135 Birbeck granules, a granule-associated antigen called LAG-1136 and E-cadherin.137
LDC progenitors can be isolated from the thymus and express low levels of CD4 whereas the mature LDC expresses CD8 and lack myeloid antigens such as CD11b/CR3 receptor, CD13 and CD33.138 LDC share most properties of MDC, but unlike the MDC which is CD11c+, the LDC is CD11c-, and lacks the myeloid markers CD13/CD33 and CD45RO. LDC are claimed to express high levels of the alpha chain of the IL-3 receptor but not alpha chains for the GM-CSF receptor and mature in response to IL-3 treatment.139 LDC also express high levels of self peptides and fas-ligand capable of inducing CD4+ T-cell death, but they do not endocytose FITC-dextran.139
Origin of APC
Studies performed on the differentiation of both mouse and human DC suggest the existence of a bi-potent myeloid progenitor cell that can develop, via blood monocytes, into either a dendritic cell or a macrophage.134,140-147 The resulting dendritic cells and macrophages are closely related and in certain circumstances seem to inter-convert.132,148-150 En route to their final location committed progenitors probably receive the various combinations of signals that modulate their ability to home and acquire their final set of functional characteristics. This may account for the apparent heterogeneity in sorted DC populations and hence the notion of “Subsets of DC”.
The developmental pathways of MDC's and macrophages are complex. At one extreme, driven by M-CSF, is the classical tissue macrophage, certain trophic functions which we would suggest is involved largely in tissue re-modelling and general phagocytosis and destruction of unwanted debris, including opsonised bacteria, viruses and tumour cells, as well as some aspects of innate immunity.70,71,76 Next are the cells that are regulated by a combination of IL-3 and CSF-1, the so called high proliferative potential cells that give rise to very large numbers of progeny in response to the synergistic mitogenic effect of these two cytokines.73 These cells may give rise to a variety of other types including those with increased ability to respond to GM-CSF. Switching the primary mitogenic signal in some of these cells to GM-CSF seems to select for a population with increased ability to acquire dendritic cell properties following exposure to TNF-α and IFN-γ.
It is likely that all DC and ‘immune’ macrophages arise from the same myeloid progenitor population and that the plurality of ‘end’ cells seen is a reflection of different final maturation pathways induced by the signalling molecules indicated (and probably by others). It may even be that many supposed end cells in this system can undergo a form of trans-differentiation as we have been able to show for TNF-α/IFN-γ induced DC, which in the presence of IL-4 rapidly acquire the characteristic morphology of macrophages (Hapel et al unpublished))
Whether the LDC is derived from this same myeloid pool awaits further investigation. Studies thus far suggest that the LDC is derived from a pool of progenitor cells that can also give rise to T lymphocytes.151-152
In Vitro Model Systems for Studying the Differentiation of Dendritic Cells
There are significant difficulties in precisely defining DC and their progenitors because cell populations isolated ex vivo are somewhat heterogeneous with respect to cell surface phenotype, morphology and tissue origin. By setting parameters carefully, FACS selection can provide a relatively homogeneous population of cells. But results generated by this sort of experimental approach must be treated with caution because FACS sorting provides a selected, not necessarily homogeneous population. The recent notion that there may be subsets of DC that preferentially stimulate different subsets of T cells, e.g., Th1 and Th2, further complicates analysis of FACS sorted populations of DC and their progenitors.
The majority of in vitro studies on the development of DC have relied on FACS sorted populations of CD34+ human peripheral blood monocytes. These cells are cultured in the presence of GM-CSF and IL-4 for between 7 and 11 days after which time dendritic cells appear in the cultures.153,154 Treatment of these cultures with TNF-α, CD40L or LPS is required to stimulate full antigen-presenting function. It is difficult, if not impossible, to precisely analyze the differentiation steps that are required to produce DC from this heterogeneous cell population over the 7-11 day culture period. Neither the starting cell nor the final mature DC have been identified. What one studies is a population effect.
In a more recently developed system155 co-culture of a FACS sorted CD34+ progenitor population with umbilical vein endothelial cells was reported to yield functional DC in 2 days. This suggests that DC can be generated quite rapidly from progenitor cells given a suitable environment and adequate differentiation signals. The DC maturation system described by Randolph155 may resemble more closely the in vivo maturation of DC but it still does not resolve the problem of using an innately heterogeneous FACS sorted progenitor cell population to study the DC lineage. A preferred system would be clonally derived progenitor cell lines that retain the ability to respond to differentiation-inducing cytokines, and other stimuli, in a homogenous manner. In such a system more precise determination of DC-specific cell markers can be made and correlated with different stages of functional maturation.
Conclusion
Macrophages and dendritic cells appear to be overlapping populations of cells that are specialised for tissue remodelling, innate immunity, and removal of cellular debris on one hand (macrophages) and antigen procesing and presentation to the immune response on the other (dendritic cells). Classical macrophages are induced by CSF-1 while the dendritic lineage relies for its maintenance on GM-CSF and other “immune” cytokines. Both cell lineages display phagocytosis but dendritic cells are far more potent at antigen processing and presentation. In some circumstances the two lineages may appear to be indistinguishable. Further experimentation with cytokine and receptor gene knock out mice is required to fully elucidate the precise function of these two lineages in immune regulation.
References
- 1.
- Till JE, McCulloch EA. A direct measurement of the radiation sensitivity of normal mouse bone marrow cells. Radiat Res. 1961;14:213–222. [PubMed: 13776896]
- 2.
- Siminovitch L, Mcculloch EA, Till JE. The distribution of colony forming cells among spleen colonies. J Cell Physiol. 1963;62:327–336. [PubMed: 14086156]
- 3.
- Wu AM, Till JE, Siminovitch L. et al. A cytological study of the capacity for differentiation of normal hematopopietic colony forming cells. J Cell Physiol. 1967;69:177–184. [PubMed: 6033948]
- 4.
- Yagi M, Ritchie KA, Sitnicka W. et al. Sustained ex vivo expansion of of hematopoietic stem cells mediated by thrombopoietin. Proc Natl Acad Sci USA. 1988;85:822–826. [PMC free article: PMC22199] [PubMed: 10393959]
- 5.
- Hodgson GS, Bradley TR. Properties of haematopoietic cells surviving 5-fluorouracil. Nature:1979. 281:381–382. [PubMed: 481601]
- 6.
- Lerner C, Harrison DE. 5-Fluorouracil spares haemopoietic stem cells responsible for long term repopulation. Exp Haematol. 1990;18:114–118. [PubMed: 2303103]
- 7.
- Akashi K, Traver D, Miyamotoo T. et al. A clonogenic common myeloid progenitor that gives rise to all myeloid lineages. Nature. 2000;404:193–197. [PubMed: 10724173]
- 8.
- Kondo M, Weissman I, Akashi K. Identification of clonogenic common lymphoid progenitors in mouse bone marrow. Cell. 1997;91:661–197. [PubMed: 9393859]
- 9.
- Harrison DE. Evaluating functional abilities of primitive haematopoietic stem cell populations. Curr Top Microbiol Immunol. 1992;177:13–30. [PubMed: 1638867]
- 10.
- Bradley TR, Metcalf D. The growth of mouse bone marrow cells in vitro. Aust J Exp Biol Med Sci. 1966;44:287–299. [PubMed: 4164182]
- 11.
- Pluznik DH, Sachs L. The cloning of normal mast cells in tissue culture. J Cell Comp Physiol. 1965;66:319–324. [PubMed: 5884359]
- 12.
- Metcalf D. Control of granulocytes and macrophages: molecular, cellular, and clinical aspects. Science. 1991;254:529–533. [PubMed: 1948028]
- 13.
- Metcalf D. Haematopoietc regulators. Trends Biochem Sci. 1992;17:286–289. [PubMed: 1412701]
- 14.
- Stanley ER, Jubinsky PT. Factors affecting the growth and differentiation of haemopoietic cells in culture. In: McCulloch EA, ed. Clinics in Haematology. London: WB Saunders. 1984:329–336. [PubMed: 6088143]
- 15.
- Pandit J, Bohm A, Jancarik J. et al. Three dimensional structure of dimeric human recombinant macrophage colony stimulating factor. Science. 1992;258:1358–1362. [PubMed: 1455231]
- 16.
- Diederichs K, Boone T, Karplus PA. Novel fold and putative receptor binding site of granulocyte-macrophage colony stimulating factor. Science. 1991;254:1779–1782. [PubMed: 1837174]
- 17.
- Hill CP, Osslund TD, Eisenberg D. The structure of granulocyte colony stimulating factor and its relationship to other growth factors. Proc Natl Acad Sci USA. 1993;90:5167–5171. [PMC free article: PMC46676] [PubMed: 7685117]
- 18.
- Bazan JF. Unravelling the structure of IL-2: Response. Science. 1992;257:410–412. [PubMed: 1631562]
- 19.
- McKay DB. Unravelling the structure of IL-2: Response. Science. 1992;257:412–413. [PubMed: 17832837]
- 20.
- Feng Y, Klein BK, Vu L. et al. 1H, 13C and 15N NMR resonance assignments, secondary structure and backbone of topology of a variant of human interleukin 3. Biochemistry. 1995;34:6540–6551. [PubMed: 7756285]
- 21.
- Somers W, Stahl M, Seehra JS. A crystal structure of interleukin-6. Implications for a novel mode of receptor dimerisation and signalling. EMBO J. 1997;16:989–997. [PMC free article: PMC1169699] [PubMed: 9118960]
- 22.
- Abdel-Meguid SS, Shieh HS, Smith WW. et al. Three dimensional structure of a genetically engineered variant of porcine growth hormone. Proc Natl Acad Sci USA. 1987;84:6434–6437. [PMC free article: PMC299091] [PubMed: 2819877]
- 23.
- Bradley TR, Hodgson GS. Detection of primitive macrophage progenitor cells in mouse bone marrow. Blood. 1979;54:1446–1450. [PubMed: 315805]
- 24.
- Stanley ER, Bartocci A, Patinkin D. et al. Regulation of very primitive multipotent haematopoietic cells by hemopoietin-1. Cell. 1986;45:667–674. [PubMed: 3085956]
- 25.
- Cross MA, Enver T. The lineage commitment of haematopoietic progenitor cells. Curr Opin Genet Devel. 1997;7:609–613. [PubMed: 9388776]
- 26.
- Bartelmez SH, Stanley ER. Synergism between hemopoietic growth factors (HGFs) detected by their effects on cells bearing receptors for a lineage specific HGF: assay of hemopoietin-1. J Cell Physiol. 1985;122:370–378. [PubMed: 2981896]
- 27.
- Morris CF, Salisbury J, Kobayashi M. et al. Absence of interleukin-3 induced proliferation in A/J bone marrow cell culture: a novel system for studying the synergistic activities of IL-3. Brit J Haematology. 1990;74:131–137. [PubMed: 2138495]
- 28.
- Hapel AJ, Fung M-C, Mak N-K. et al. Bone marrow cells from A/J mice do not proliferate in interleukin-3 but contain normal numbers of interleukin-3 receptors. Brit J Haematology. 1992;82:488–93. [PubMed: 1486029]
- 29.
- Tushinski RJ, Oliver IT, Guilbert LJ. et al. Survival of mononuclear phagocytes depends on a lineage specific growth factor that the differentiated cells selectively destroy. Cell. 1982;28:71–81. [PubMed: 6978185]
- 30.
- Koury MJ, Bondurant MC. Erythropoietin retards DNA breakdown and prevents programmed cell death in erythroid progenitor cells. Science. 1990;248:378–381. [PubMed: 2326648]
- 31.
- Morris C, Young IG, Hapel AJ. Molecular and cellular biology of interleukin-3In: Dexter, Garland, Testa, eds.Colony Stimulating Factors: Molecular and Cellular BiologyNew York: Marcel Decker Inc.,1989177–214.
- 32.
- Kondo M, Scherer DC, Miyamoto T. et al. Cell fate conversion of lymphoid committed progenitors by instructive actions of cytokines. Nature. 2000;407:383–386. [PubMed: 11014194]
- 33.
- Kulessa H, Framption J, Graf T. GATA-1 reprograms avian myelomonocytic cell lines into eosinophils thromboblasts and erythroblasts. Genes and Dev. 1995;9:1250–1262. [PubMed: 7758949]
- 34.
- Nerlov C, Graf T. PU.1 induces myeloid lineage commitment in multipotent haematopopietic progenitors. Genes and Dev. 1998;12:2403–1412. [PMC free article: PMC317050] [PubMed: 9694804]
- 35.
- Nutt S, Heavey B, Rolink A. et al. Commitment of the B-lymphoid lineage depends on the transcription actor Pax 5. Nature. 1999;401:556–562. [PubMed: 10524622]
- 36.
- Anderson DM, Williams DE, Tushinski R. et al. Alternate splicing of mRNA encoding human mast cell growth factor and localisation of the gene to chromosome 12q22-24. Cell Growth Differ. 1991;2:373–378. [PubMed: 1724381]
- 37.
- Bazan JF. Genetic and structural homology of stem cell factor and macrophage colony stimulating factor. Cell. 1991;65:9–10. [PubMed: 1707344]
- 38.
- Flanagan JG, Chan DC, Leder P. Transmembrane form of the kit ligand growth factor is determined by alternative splicing and is missing in the Sld mutant. Cell. 1991;64:1025–1035. [PubMed: 1705866]
- 39.
- Lyman SD, Jacobsen SEW. c-kit ligand and flt-3 ligand : stem cell factors with overlapping yet distinct activities. Blood. 1998;91:1101–1134. [PubMed: 9454740]
- 40.
- Stanley ER. CSF-1In: Oppenheim JJ, Feldmann M, eds.Cytokine DatabaseLondon: Academic2000913–934.
- 41.
- Qui F, Ray P, Brown K. et al. Primary structure of c-kit: relationship with the CSF1/PDGF receptor kinase family - oncogenic activation of v-kit involves deletion of extracellular domain and C terminus. EMBO J. 1988;7:1003–1011. [PMC free article: PMC454427] [PubMed: 2456920]
- 42.
- Coussens L, van BeverenC, Smith D. et al. Structural alteration of viral homologue of receptor proto oncogene fms at carboxy-terminus. Nature. 1986;32:277–280. [PubMed: 2421165]
- 43.
- Linnekin D. Early signalling pathways activated by c-Kit in haemopoietic cells. Int J Biochem Cel Biol. 1999;31:1053–1074. [PubMed: 10582339]
- 44.
- Lee PS, Wang Y, Dominguez MG. et al. The Cbl protooncoprotein stimulates CSF-1 receptor multiubiquitination and endocytosis, and attenuates macrophage proliferation. EMBO J. 1999;18:3616–3628. [PMC free article: PMC1171440] [PubMed: 10393178]
- 45.
- Sherr CJ. Colony stimulating factor 1 receptor. Blood. 1990;75:1–1. [PubMed: 2153029]
- 46.
- Rosnet O, Marchetto S, DeLapeyriere O. et al. Murine Flt 3, a gene encoding a novel tyrosine kinase receptor of the PDGFR/CSF-1R family. Oncogene. 1991;6:1642–1650. [PubMed: 1656368]
- 47.
- Matthews W, Jordan CT, Wiegland GW. et al. A receptor tyrosine kinase specific to hematopoietic stem and progenitor cell enriched populations. Cell. 1991;65:1143–1152. [PubMed: 1648448]
- 48.
- Lyman SD, James L, vanden Bos T. et al. Molecular cloning of a ligand for the flt3/flk2 tyrosine kinase receptor: a proliferative factor for primitive hematopoietic cells. Cell. 1993;75:1157–1167. [PubMed: 7505204]
- 49.
- Hannum C, Culpepper J, Campbell D. et al. Ligand for FLT3/FLK2 receptor tyrosine kinase regulates growth of haematopoietic stem cells and is encoded by variant RNAs. Nature. 1994;368:643–648. [PubMed: 8145851]
- 50.
- McKenna HJ, Stocking KL, Miller RE. et al. Mice lacking flt3 ligand have deficient hematopoiesis affecting hematopoietic progenitor cells, dendritic cells, and natural killer cells. Blood. 2000;95:3489–3497. [PubMed: 10828034]
- 51.
- Mackarehtschian K, Hardin JD, Mooore KA. et al. Targeted disruption of the flk2/flt3 gene leads to deficiencies in primitive hematopoietic progenitors. Immunity. 1995;3:147–161. [PubMed: 7621074]
- 52.
- Roth P, Stanley ER. The biology of CSF-1 and its receptor. Cuirr Top Mictrobiol Immunol. 1992;181:141–167. [PubMed: 1424779]
- 53.
- Stanley ER, Berg KL, Einstein DB. et al. Biology and action of CSF-1 Mol. Reprod Dev. 1997;46:4–10. [PubMed: 8981357]
- 54.
- Stanley ER. Colony stimulating factor -1 (macrophage colony stimulating factor). In: Thompson AW, ed. The cytokine Handbook.2nd ed. San Diego: Academic. 1994:387–418.
- 55.
- Yeung Y-G, Wang Y, Einstein DB. et al. Colony stimulating factor 1 stimulates the formation of multimeric cytosolic complexes of signalling proteins and cytoskeletal components in macrophages. J Biol Chem. 1998;273:17128–17137. [PubMed: 9642280]
- 56.
- Yeung Y-G, Stanley ER. Proteomic approaches to analysis of early events in CSF-1 signal transduction. Mole Cell Proteomics. 2003;2:1143–1155. [PubMed: 12966146]
- 57.
- Bartocci A, Mastrogiannis DS, Migliorati G. et al. Macrophages specifically regulate the concentration of their own growth factor I the circulation. Proc Natl Acad Sci USA. 1987;84:6179–6183. [PMC free article: PMC299033] [PubMed: 2819867]
- 58.
- Guilbert LJ, Stanley ER. The interaction of 125I-colony stimulating factor 1 with bone marrow-derived macrophages. J Biol Chem. 1986;261:4024–4032. [PubMed: 3485098]
- 59.
- Dai X-M, Ryan GR, Hapel AJ. et al. Targeted disruption of the mouse colony-stimulating factor 1 receptor gene results in osteopetrosis, mononuclear phagocyte deficiency, increased primitive progenitor cell frequencies and reproductive defects. Blood. 2002;99:111–120. [PubMed: 11756160]
- 60.
- Hume DA, Pavli P, Donahue RE. et al. The effect of human recombinant macrophage colony stimulating factor (CSF-1) on the murine mononuclear phagocyte system in vivo. J Immunol. 1988;141:3405–3409. [PubMed: 3053899]
- 61.
- Wiktor-Jedrzejczak W, Bartocci A, Ferrante Jr. AW. et al. Total absence of colony stimulating factor 1 in the macrophage deficient ostreopetrotic (op/op) mouse. Proc Natl Acad Sci USA. 1990;87:4828–4832. [PMC free article: PMC54211] [PubMed: 2191302]
- 62.
- Yoshida H, Hatashi S-I, Kunisada T. et al. The murine mutation “osteopetrosis” (op) is a mutation in the coding region of the macrophage colony stimulating factor (Csfm) gene. Nature. 1990;345:442–444. [PubMed: 2188141]
- 63.
- Felix R, Hofstetter W, Wetterwald A. et al. Role of colony stimulating factor 1 in bone metabolism. J Cell Biochem. 1994;120:1357–1372. [PubMed: 7962166]
- 64.
- Pollard JW, Stanley ER. Pleiotropic roles for CSF-1 in development defined by the mouse mutation osteopetrotic (op). Adv Dev Biochem. 1996;4:153–193.
- 65.
- Cecchini MG, Dominguez MG, Mocci S. et al. Role of colony stimulating factor 1 in the establishment and regulation of tissue macrophages during postnatal development of the mouse. Development. 1994;120:1357–1372. [PubMed: 8050349]
- 66.
- Ryan GR, Dai X-M, Dominguez MG. et al. Rescue of the colony-stimulating factor 1 (CSF-1)-nullizygous mouse (Csf1op /Csf1op ) phenotype with a CSF-1 transgene and identification of sites of local CSF-1 synthesis. Blood. 2001;98:74–84. [PubMed: 11418465]
- 67.
- Dai X-M, Zong XH, Sylvestre V. et al. Incomplete restoration of colony stimulating factor-1 (CSF-1) function in CSF-1-deficient Csf1op/Csf1op mice by transgenic expression of cell surface CSF-1. Blood. 2004;103:1114–1123. [PubMed: 14525772]
- 68.
- Chang M-DY, Stanley ER, Khalili H. et al. Osteopetrotic (op/op) mice deficient in macrophages have the ability to mount a normal T-cell-dependent immune response. Cellular Immunol. 1995;162:146–152. [PubMed: 7704903]
- 69.
- Wiktor-Jedrzejczak W, Ansari AA, Szperl M. et al. Distinct in vivo functions of two macrophage subpopulations as evidenced by studies using macrophage-deficient op/op mouse. Eur J Immunol. 1992;22:1951–1954. [PubMed: 1378025]
- 70.
- Watts C. Immunology Inside the gearbox of the dendritic cell. Nature. 1997;388:724–725. [PubMed: 9285579]
- 71.
- Banyer J, Hapel AJ. Myb transformed myeloid cell lines as a model for monocyte differentiation intomdendritic cells and macrophages. J Leucocyte Biology. 1999;66:1–7. [PubMed: 10449157]
- 72.
- Bartelmez SH, Sacca R, Stanley ER. Lineage specific receptors used to identify a growth factor for developmentally early hemopoietic cells: assay for hemopoietin 2. J Cell Physiol. 1985;122:362–369. [PubMed: 2981895]
- 73.
- Bartelmez SH, Bradley TR, Bertoncello I. et al. Interleukin 1 plus interleuki- 3 plus colony stimulating factor 1 are essential for clonal proliferation of primitive myeloid bone marrow stem cells. Exp Hematol. 1989;17:240–245. [PubMed: 2783913]
- 74.
- Lacey DL, Timms E, Tan HL. et al. Osteoprotegerin is a cytokine that regulates osteoclast differentiation and activation. Cell. 1998;93:165–176. [PubMed: 9568710]
- 75.
- Cohen PE, Nishimura K, Zhu L. et al. Macrophages: important accessory cells for reproductive function. J Leucocyte Biol. 1999;66:765–772. [PubMed: 10577508]
- 76.
- Guleria I, Pollard JW. The trophoblast is a component of the innate immune system during pregnancy. Nature Med. 2000;6:589–593. [PubMed: 10802718]
- 77.
- Kacinski BM. CSF-1 and its receptor in ovarian, endometrial and breast cancer. Ann Med. 1995;27:79–85. [PubMed: 7742005]
- 78.
- Kacinski BM. CSF-1 and its receptor in breast carcinomas and neoplasms of the female reproductive tract. Mol Reprod Dev. 1997;46:71–74. [PubMed: 8981366]
- 79.
- Miyajima A, Mui AL-F, Ogorochi T. et al. Receptors for granulocyte macrophage colony stimulating factor, interleukin 3 and interleukin 5. Blood. 1993;82:1960–1974. [PubMed: 8400249]
- 80.
- Bagley CJ, Woodcock JM, Stomski FC. et al. The structural and functional basis of cytokine receptor activation: lessons from α common a subunit of the granulocyte-macrophage colony stimulating factor, interleukin-3 (IL-3), and IL-5 receptors. Blood. 1997;89:1471–1482. [PubMed: 9057626]
- 81.
- Fung M-C, Hapel AJ, Ymer S. et al. Molecular cloning of cDNA for murine interleukin-3. Nature. 1984;307:233–237. [PubMed: 6420702]
- 82.
- Mach N, Lantz CS, Galli SJ. et al. Involvement of interleukin-3 in delayed-type hypersensitivity. Blood. 1998;91:778–783. [PubMed: 9446636]
- 83.
- De VriesP, Brasel KA, Eisenman JR. et al. The effect of recombinant mast cell growth factor on purified murine haematopoietic stem cells. J Exp Med. 1991;173:1205–1211. [PMC free article: PMC2118854] [PubMed: 1708810]
- 84.
- Hapel AJ, Fung M-C, Young IG. et al. Biological properties of molecularly cloned and expressed murine interleukin 3. Blood. 1985;65:1453–1459. [PubMed: 3922456]
- 85.
- Gearing DP, King JA, Gough NM. et al. Expression cloning of a receptor for human granulocyte macrophage colony stimulating factor receptor. EMBO J. 1989;8:3667–3676. [PMC free article: PMC402049] [PubMed: 2555171]
- 86.
- Raines MA, Liu L, Quan SG. et al. Identification and molecular cloning of a soluble human granulocyte macrophage colony stimulating factor receptor. Proc Natl Acad Sci USA. 1991;88:8203–8207. [PMC free article: PMC52475] [PubMed: 1832774]
- 87.
- Burgess AW. Granulocyte macrophage colony stimulating factorIn: Sporn MB, Roberts AB, eds.A Handbook of Experimental Pharmacology. Vol 95. Peptide Growth Factors and Their ReceptorsBerlin: Springer,1990723–745.
- 88.
- Kannourakis G, Johnson GR. Fractionation of subsets of BFU-E from normal human bone marrow: responsiveness to erythropoietin, human placental-conditioned medium, or granulocyte macrophage colony stimulating factor. Blood. 1988;71:758–765. [PubMed: 3278756]
- 89.
- Dranoff G, Crawford D, Sadelain M. et al. Involvement of granulocyte macrophage colony stimulating factor in pulmonary homeostasis. Science. 1994;264:713–715. [PubMed: 8171324]
- 90.
- Chomarat P, Rybak ME, Bancereau J. Interleukin-4. In: Thomson AW, ed. The Cytokine Handbook. 3rd ed. San Diego: Academic. 1998:133–174.
- 91.
- Murata T, Taguchi J, Puri RK. et al. Sharing of receptor subunits and signal transduction pathway between the IL-4 and IL-13 receptor system. Int J Haematol. 1999;69:13–20. [PubMed: 10641437]
- 92.
- Kuhn R, Rajewsky K, Muller W. Generation and analysis of interleukin-4 deficient mice. Science. 1991;254:707–710. [PubMed: 1948049]
- 93.
- Kopf M, Legros G, Coyles AJ. et al. Immune responses of IL-4, IL-5, IL-6 deficient mice. Immunol Rev. 1995;148:45–69. [PubMed: 8825282]
- 94.
- Hapel AJ, McColl SR. Cytokines in Immunology. In: Principles of Medical Biology. Madison: JAI Press. 1996
- 95.
- Shankaran V, Schrieber RD. IFNγ receptor In: Oppenheim JJ, Feldman M, eds.Cytokine databaseLondon: Academic20001819–1836.
- 96.
- Billiau A, Vandenbroeck K. IFNγ In: Oppenheim JJ, Feldman M, eds.Cytokine DatabaseLondon: Academic,2000641–688.
- 97.
- Aggarwal BB, Samanta A, Feldmann M. TNF-α In: Oppenheim JJ, Feldmann M, eds.Cytokine DatabaseLondon: Academic,2000413–434.
- 98.
- Ruddle N, Ware CF. Lymphotoxin α and β In: Oppenheim JJ, Feldmann M, eds.Cytokine Database. London: Academic 2000435–446.
- 99.
- Banchereau J, Steinman RM. Dendritic cells and the control of immunity. Nature. 1998;392:245–392. [PubMed: 9521319]
- 100.
- Cella M, Sallusto F, Lanzavecchia A. Origin, maturation and antigen presenting function of dendritic cells. Curr Opin Immunol. 1997;9:10–16. [PubMed: 9039784]
- 101.
- Austyn JM, Gordon S. F4/80, a monoclonal antibody directed specifically against the mouse macrophage. Eur J Immunol. 1981;11:805–815. [PubMed: 7308288]
- 102.
- Roth P, Stanley ER. The biology of CSF-1 and its receptor. Curr Top Microbiol Immunol. 1992;181:141–167. [PubMed: 1424779]
- 103.
- Holt PG, Oliver J, Bilyk N. et al. Downregulation of the antigen presenting cell function(s) of pulmonary dendritic cells in vivo by resident alveolar macrophages. J Exp Med. 1993;177:397–407. [PMC free article: PMC2190916] [PubMed: 8426110]
- 104.
- Chegini N. The role of growth factors in peritoneal healing: transforming growth factor beta (TGF-beta). Eur J Surg Suppl. 1997;577:17–23. [PubMed: 9076448]
- 105.
- Lukas M, Stossel H, Hefel L. et al. Human cutaneous dendritic cells migrate through dermal lymphatic vessels in a skin organ culture model. J Invest Dermatol. 1996;106:1293–1299. [PubMed: 8752673]
- 106.
- De SmedtT, Pajak B, Muraille E. et al. Regulation of dendritic cell numbers and maturation by lipopolysaccharide in vivo. J Exp Med. 1996;184:1413–1424. [PMC free article: PMC2192842] [PubMed: 8879213]
- 107.
- Sallusto F, Cella M, Danieli C. et al. Dendritic cells use macropinocytosis and the mannose receptor to concentrate macromolecules in the major histocompatibility complex class II compartment: down-regulation by cytokines and bacterial products. J Exp Med. 1995;182:389–400. [PMC free article: PMC2192110] [PubMed: 7629501]
- 108.
- Sallusto F, Lanzavecchia A. Efficient presentation of soluble antigen by cultured human dendritic cells is maintained by granulocyte/macrophage colony-stimulating factor plus interleukin 4 and downregulated by tumor necrosis factor alpha. J Exp Med. 1994;179:1109–1118. [PMC free article: PMC2191432] [PubMed: 8145033]
- 109.
- Winzler C, Rovere P, Rescigno M. et al. Maturation stages of mouse dendritic cells in growth factor-dependent long-term cultures. J Exp Med. 1997;185:317–328. [PMC free article: PMC2196118] [PubMed: 9016880]
- 110.
- Nijman HW, Kleijmeer MJ, Ossevoort MA. et al. Antigen capture and major histocompatibility class II compartments of freshly isolated and cultured human blood dendritic cells. J Exp Med. 1995;182:163–174. [PMC free article: PMC2192095] [PubMed: 7790816]
- 111.
- Pierre P, Turley SJ, Gatti E. et al. Developmental regulation of MHC class II transport in mouse dendritic cells. Nature. 1997;388:787–792. [PubMed: 9285592]
- 112.
- Cella M, Engering A, Pinet V. et al. Inflammatory stimuli induce accumulation of MHC class II complexes on dendritic cells. Nature. 1997;388:782–787. [PubMed: 9285591]
- 113.
- Maraskovsky E, Pulendran B, Shortman. KLymphoid-related dendritic cellsIn: Lotze MT, Thompson, AW, eds.Dendritic CellsAcademic Press.199993–107.
- 114.
- Kronin V, Vremec D, Winkel K. et al. Are CD8+ dendritic cells (DC) veto cells? The role of CD8 on DC in DC development and in the regulation of CD4 and CD8 T cell responses. Int Immunol. 1997;9:1061–1064. [PubMed: 9237115]
- 115.
- Suss G, Shortman K. A subclass of dendritic cells kills CD4 T cells via Fas/Fas-ligand-induced apoptosis. J Exp Med. 1996;183:1789–1796. [PMC free article: PMC2192509] [PubMed: 8666935]
- 116.
- Peters JH, Gieseler R, Thiele B. et al. Dendritic cells: from ontogenetic orphans to myelomonocytic descendants. Immunol Today. 1996;17:273–276. [PubMed: 8962630]
- 117.
- Inaba K, Inaba M, Romani N. et al. Generation of large numbers of dendritic cells from mouse bone marrow cultures supplemented with granulocyte/macrophage colony-stimulating factor. J Exp Med. 1992;176:1693–1702. [PMC free article: PMC2119469] [PubMed: 1460426]
- 118.
- Heufler C, Koch F, Schuler G. Granulocyte/macrophage colony-stimulating factor and interleukin 1 mediate the maturation of murine epidermal Langerhans cells into potent immunostimulatory dendritic cells. J Exp Med. 1988;167:700–705. [PMC free article: PMC2188828] [PubMed: 3279156]
- 119.
- Zhang Y, Mukaida N, Wang J. et al. Induction of dendritic cell differentiation by granulocyte-macrophage colony-stimulating factor, stem cell factor, and tumor necrosis factor alpha in vitro from lineage phenotypes-negative c-kit+ murine hematopoietic progenitor cells. Blood. 1997;90:4842–4853. [PubMed: 9389701]
- 120.
- Peters JH, Xu H, Ruppert J. et al. Signals required for differentiating dendritic cells from human monocytes in vitro. Adv Exp Med Biol. 1993;329:275–280. [PubMed: 8379382]
- 121.
- Ruppert J, Schutt C, Ostermeier D. et al. Down-regulation and release of CD14 on human monocytes by IL-4 depends on the presence of serum or GM-CSF. Adv Exp Med Biol. 1993;329:281–286. [PubMed: 7691031]
- 122.
- Steinbach F, Krause B, Thiele B. Monocyte derived dendritic cells (MODC) present phenotype and functional activities of Langerhans cells/dendritic cells. Adv Exp Med Biol. 1995;378:151–153. [PubMed: 8526042]
- 123.
- Romani N, Gruner S, Brang D. et al. Steinman PO, Schuler RM, Proliferating dendritic cell progenitors in human blood. J Exp Med. 1994;180:83–93. [PMC free article: PMC2191538] [PubMed: 8006603]
- 124.
- Xu H, Kramer M, Spengler HP. et al. Dendritic cells differentiated from human monocytes through a combination of IL-4, GM-CSF and IFN-gamma exhibit phenotype and function of blood dendritic cells. Adv Exp Med Biol. 1995;378:75–78. [PubMed: 8526149]
- 125.
- Caux C, Dezutter-Dambuyant C, Schmitt D. et al. GM-CSF and TNF-alpha cooperate in the generation of dendritic Langerhans cells. Nature. 1992;360:258–261. [PubMed: 1279441]
- 126.
- Caux C, Massacrier C, Vanbervliet B. et al. Activation of human dendritic cells through CD40 cross-linking. J Exp Med. 1994;180:1263–1272. [PMC free article: PMC2191669] [PubMed: 7523569]
- 127.
- Witsell AL, Schook LB. Macrophage heterogeneity occurs through a developmental mechanism. Proc Natl Acad Sci USA. 1991;88:1963–1967. [PMC free article: PMC51146] [PubMed: 1705715]
- 128.
- Rutherford MS, Witsell A, Schook LB. Mechanisms generating functionally heterogeneous macrophages: chaos revisited. J Leukoc Biol. 1993;53:602–618. [PubMed: 8501399]
- 129.
- Rutherford MS, Schook LB. Differential immunocompetence of macrophages derived using macrophage or granulocyte-macrophage colony-stimulating factor. J Leukoc Biol. 1992;51:69–76. [PubMed: 1740646]
- 130.
- Reid CD, Stackpoole A, Meager A. et al. Interactions of tumor necrosis factor with granulocytemacrophage colony-stimulating factor and other cytokines in the regulation of dendritic cell growth in vitro from early bipotent CD34+ progenitors in human bone marrow. J Immunol. 1992;149:2681–2688. [PubMed: 1383322]
- 131.
- Inaba K, Inaba M, Deguchi M. et al. Granulocytes, macrophages, and dendritic cells arise from a common major histocompatibility complex class II-negative progenitor in mouse bone marrow. Proc Natl Acad Sci USA. 1993;90:3038–3042. [PMC free article: PMC46232] [PubMed: 8464920]
- 132.
- Boehmelt G, Madruga J, Dorfler P. et al. Dendritic cell progenitor is transformed by a conditional v-Rel estrogen receptor fusion protein v-RelER. Cell. 1995;80:341–352. [PubMed: 7834754]
- 133.
- Santiago-Schwarz F, Belilos E, Diamond B. et al. TNF in combination with GM-CSF enhances the differentiation of neonatal cord blood stem cells into dendritic cells and macrophages. J Leukocyte Biol. 1992;52:274–281. [PubMed: 1387891]
- 134.
- Caux C, Vanbervliet B, Massacrier C. et al. CD34+ hematopoietic progenitors from human cord blood differentiate along two independent dendritic cell pathways in response to GM-CSF+TNF alpha. J Exp Med. 1996;184:695–706. [PMC free article: PMC2192705] [PubMed: 8760823]
- 135.
- Strunk D, Egger C, Leitner G. et al. A skin homing molecule defines the Langerhans cell progenitor in human peripheral blood. J Exp Med. 1997;185:1131–1136. [PMC free article: PMC2196235] [PubMed: 9091586]
- 136.
- Kashihara M, Ueda M, Horiguchi Y. et al. A monoclonal antibody specifically reactive to human Langerhans cells. J Invest Dermatol. 1986;87:602–607. [PubMed: 3534103]
- 137.
- Tang A, Amagai M, Granger LG. et al. Adhesion of epidermal Langerhans cells to kertinocytes mediated by E-cadherin. Nature. 1993;361:82–85. [PubMed: 8421498]
- 138.
- Steinman RM, Pack M, Inaba K. Dendritic cells in the T-cell areas of lymphoid organs. Immunol Rev. 1997;156:25–37. [PubMed: 9176697]
- 139.
- Ito T, Inaba M, Inaba K. et al. Distinctive subsets of dendritic cells in human peripheral blood. J Leuk Biol. 1998;Suppl 2:A1019.
- 140.
- Goordyal P, Isaacson PG. Immunocytochemical characterization of monocyte colonies of human bone marrow: a clue to the origin of Langerhans cells and interdigitating reticulum cells. J Pathol. 1985;146:189–195. [PubMed: 3897494]
- 141.
- de FraissinetteA, Schmitt D, Dezutter-Dambuyant C. et al. Culture of putative Langerhans cell bone marrow precursors: characterization of their phenotype. Exp Hematol. 1988;16:764–768. [PubMed: 3169159]
- 142.
- Burkly L, Hession C, Ogata L. et al. Expression of relB is required for the development of thymic medulla and dendritic cells. Nature. 1995;373:531–536. [PubMed: 7845467]
- 143.
- Szabolcs P, Avigan D, Gezelter S. et al. Dendritic cells and macrophages can mature independently from a human bone marrow-derived, post-colony-forming unit intermediate. Blood. 1996;87:4520–4530. [PubMed: 8639819]
- 144.
- Metcalf D. Murine hematopoietic stem cells committed to macrophage/dendritic cell formation: Stimulation by Flk2-ligand with enhancement by regulators using the gp130 receptor chain. Proc Natl Acad Sci USA. 1997;94:11552–1556. [PMC free article: PMC23534] [PubMed: 9326647]
- 145.
- Ryncarz RE, Anasetti C. Expression of CD86 on human marrow CD34(+) cells immunocompetent committed precursors of macrophages and dendritic cells. Blood. 1998;91:3892–3900. [PubMed: 9573027]
- 146.
- Hausser G, Ludewig B, Gelderblom HR. et al. Monocyte-derived dendritic cells represent a transient stage of differentiation in the myeloid lineage. Immunobiol. 1997;197:534–542. [PubMed: 9413752]
- 147.
- Goerdt S, Kodelja V, Schmuth M. et al. The mononuclear phagocyte-dendritic cell dichotomy: myths, facts, and a revised concept. Clin Exp Immunol. 1996;105:1–9. [PMC free article: PMC2200469] [PubMed: 8697614]
- 148.
- Rossi G, Heveker N, Thiele B. et al. Development of a Langerhans cell phenotype from peripheral blood monocytes. Immunol Lett. 1992;31:189–197. [PubMed: 1371267]
- 149.
- Murphy GF, Messadi D, Fonferko E. et al. Phenotypic transformation of macrophages to Langerhans cells in the skin. Am J Pathol. 1986;123:401–406. [PMC free article: PMC1888260] [PubMed: 3521301]
- 150.
- Peters JH, Ruhl S, Friedrichs D. Veiled accessory cells deduced from monocytes. Immunobiol. 1987;176:154–166. [PubMed: 3502337]
- 151.
- Ardavin C, Wu L, Li CL. et al. Thymic dendritic cells and T cells develop simultaneously in the thymus from a common precursor population. Nature. 1993;362:761–763. [PubMed: 8469288]
- 152.
- Wu L, Li CL, Shortman K. Thymic dendritic cell precursors: relationship to the T lymphocyte lineage and phenotype of the dendritic cell progeny. J Exp Med. 1996;184:903–911. [PMC free article: PMC2192802] [PubMed: 9064350]
- 153.
- Bender A, Sapp M, Schuler G. et al. Improved methods for the generation of dendritic cells from non-proliferating progenitors in human blood. J Immunol Methods. 1996;196:121–135. [PubMed: 8841451]
- 154.
- Romani N, Reider D, Heuer M. et al. Generation of mature dendritic cells from human blood. An improved method with special regard to clinical applicability. J Immunol Methods. 1996;196:137–151. [PubMed: 8841452]
- 155.
- Randolph GJ, Beaulieu S, Lebecque S. et al. Differentiation of monocytes into dendritic cells in a model of transendothelial trafficking. Science. 1998;282:480–483. [PubMed: 9774276]
- Introductory Remarks
- Sources of Cytokines and Regulation of Their Production
- Signalling
- Cooperative Interaction of Cytokines in Proliferation and Differentiation
- The Role of Transcription Factors
- Tyrosine Kinase Receptors
- Cytokines and Receptors with a Significant Effect on Cells of the Monocyte Macrophage Lineage
- CSF-1
- Interleukin-3
- Granulocyte-Macrophage Colony Stimulating Factor
- Interleukin-4
- Interferon-γ
- Tumour Necrosis Factor
- Antigen Processing Cells
- Classification and Sources of Antigen Processing Cells
- Origin of APC
- In Vitro Model Systems for Studying the Differentiation of Dendritic Cells
- Conclusion
- References
- Cytokines, Receptors and Signalling Pathways Involved in Macrophage and Dendriti...Cytokines, Receptors and Signalling Pathways Involved in Macrophage and Dendritic Cell Development - Madame Curie Bioscience Database
- Physiological Roles and Mechanisms of Signaling by TRAF2 and TRAF5 - Madame Curi...Physiological Roles and Mechanisms of Signaling by TRAF2 and TRAF5 - Madame Curie Bioscience Database
- Sympathetic Nervous System Regulation of Metastasis - Madame Curie Bioscience Da...Sympathetic Nervous System Regulation of Metastasis - Madame Curie Bioscience Database
- VEGF and Endothelial Guidance in Angiogenic Sprouting - Madame Curie Bioscience ...VEGF and Endothelial Guidance in Angiogenic Sprouting - Madame Curie Bioscience Database
- Physical Background and Technical Realizations of Hyperthermia - Madame Curie Bi...Physical Background and Technical Realizations of Hyperthermia - Madame Curie Bioscience Database
Your browsing activity is empty.
Activity recording is turned off.
See more...