The description of skin dendritic cells (DC) by Langerhans in 1868 was followed by prolonged speculation as to their function. Steinman and Cohn identified mouse spleen DC in 19731 and initiated a series of experiments that established lymphoid tissue–derived DC as potent stimulators of primary immune responses.2-5 The observation that similar cells were present in the nonlymphoid tissues of both rodents6-8 and humans,9,10 combined with early evidence that they played an important role in heart and kidney transplant rejection,11-14 generated further interest in DC. However, a paucity of markers for DC, the difficulty distinguishing DC from monocytes/macrophages (Mo/MØ), and the problems involved in purifying DC made for slow progress. Nonetheless, several laboratories persisted with their investigations, leading to the current acceptance that DC represent discrete leukocyte population(s) of specialist or “professional” antigen presenting cells (APC), with an extraordinary capacity for initiating primary (and secondary) T-lymphocyte responses.15 16 

It is now possible to grow DC-like cells in culture17-21 and the first monoclonal antibodies (MoAbs)22,23 that react with human DC are available. DC migration24 is probably directed by a range of mediators and recent data suggest that reciprocal T-lymphocyte feedback provides antigen (Ag)-specific control of DC Ag presentation.16,25 26 This new understanding of DC function may provide opportunities for therapeutic intervention in bone marrow transplantation (BMT), solid organ transplantation, and autoimmune disease. Protocols for clinical immunotherapy programs, targeted on malignant cell Ag or infectious agents, are being designed to exploit DC as “nature's adjuvant” for optimal therapeutic vaccination.

A practical issue facing hematologists is the lack of a clear definition of a DC. This, combined with uncertainty as to the ontogeny of DC, limited data as to the interrelationship of different DC populations, and the fact that activation/differentiation events may dramatically change the morphology, molecular expression (phenotype), and function of DC, has created difficulties. This review will concentrate mainly on the myeloid lineage–derived DC, which associate with T lymphocytes within lymphoid tissue, to provide critical APC activity for initiating specific T-lymphocyte activation and proliferation.15,27,28 Follicular dendritic cells (FDC), which are present within the B-lymphoid follicle, do not originate from the BM, have a different phenotype, and retain Ag for prolonged periods, thereby restimulating B lymphocytes and perhaps T lymphocytes, will not be reviewed. Nor will dendritic epithelial cells, a specialized subset of T lymphocytes with extensive membrane processes, which are found in epithelial surfaces in mice but are rare in humans. Myeloid DC should be further distinguished from thymic DC, which appear to derive from a lymphoid stem cell.29 Thymic DC (and perhaps other lymphoid lineage–derived DC) have an entirely different role30,31; they delete maturing T lymphocytes and the prediction that they may have different properties from the myeloid-derived DC, although controversial, has received experimental support.32 

The cardinal properties of the myeloid lineage–derived DC (hereafter DC) include (1) the ability to take up, process, and present Ag; (2) the ability to migrate selectively through tissues; and (3) the ability to interact with, stimulate, and direct T-lymphocyte responses. DC may be the only cell capable of stimulating a naive T lymphocyte but other “nonprofessional” APC (and DC) can stimulate experienced (activated or memory) T lymphocytes. As such, DC have unique cell interaction capabilities, some of which relate to their extensive cell membrane processes that are acquired late during DC differentiation/activation (Fig 1). DC have been defined, as of necessity, by these specialized functional characteristics15,27 33 to distinguish them from Mo and MØ.

Fig. 1.

The morphology of human (see text for description) DC. (A) Fresh “lineage negative” blood DC, isolated without a period of tissue culture, using immunoselection.57 (May-Grünwald-Giemsa [MGG], original magnification [OM] × 1,433). (B) CMRF-44 sorted, Nycodenz gradient purified cultured blood DC77 MGG, OM × 1,433). (C) Tonsil low-density cultured DC stained with anti–HLA-DR. The veils and dendritic processes are more obvious in these preparations (OM × 1,433). (D) Mo-derived DC preparation stained with CMRF-44 using an immunoenzyme (brown, Peroxidase-DAB) technique (in preparation) (OM × 1,433). (E and F ) Fresh “lineage negative” blood DC clustered with CD4+ purified autologous T lymphocytes in the presence of staphylococcal entertoxin A (SEA) and MGG stained. The DC is stained in another cluster (F ) for the costimulator molecule CD86 using an immunoenzyme (Alkaline Phosphatase-Fast Blue) technique (OM × 1,197). (G) EM appearances of a CMRF-44–positive cultured blood DC. Note the mitochondria endosomes and lysosomal vacuoles (OM × 17,000). (H) DC in the interstitial tissues of rat heart identified by anti-MHC class II staining (immunofluorescence) (OM × 479). (I) Dermal CMRF-44+ DC (red, Peroxidase-AEC) and T lymphocytes (blue, Alkaline Phosphatase-Fast Blue) within a section of normal skin adjacent to a hair follicle (OM × 479). (J) Lymph node interfollicular (T lymphocyte) area containing CMRF-44+ IDC (brown, Peroxidase-DAB) compared with CD14 Mo and CD20 B lymphocytes (blue, Alkaline phosphatase-Fast Blue) (OM × 143). (K) Lymph node interfollicular region with CMRF-44+ IDC (blue, Alkaline phosphatase-Fast Blue) showing nuclear labeling for the transcription factor Rel B (brown, Peroxidase-DAB) (OM × 1,197). Bar = 10 μm.

Fig. 1.

The morphology of human (see text for description) DC. (A) Fresh “lineage negative” blood DC, isolated without a period of tissue culture, using immunoselection.57 (May-Grünwald-Giemsa [MGG], original magnification [OM] × 1,433). (B) CMRF-44 sorted, Nycodenz gradient purified cultured blood DC77 MGG, OM × 1,433). (C) Tonsil low-density cultured DC stained with anti–HLA-DR. The veils and dendritic processes are more obvious in these preparations (OM × 1,433). (D) Mo-derived DC preparation stained with CMRF-44 using an immunoenzyme (brown, Peroxidase-DAB) technique (in preparation) (OM × 1,433). (E and F ) Fresh “lineage negative” blood DC clustered with CD4+ purified autologous T lymphocytes in the presence of staphylococcal entertoxin A (SEA) and MGG stained. The DC is stained in another cluster (F ) for the costimulator molecule CD86 using an immunoenzyme (Alkaline Phosphatase-Fast Blue) technique (OM × 1,197). (G) EM appearances of a CMRF-44–positive cultured blood DC. Note the mitochondria endosomes and lysosomal vacuoles (OM × 17,000). (H) DC in the interstitial tissues of rat heart identified by anti-MHC class II staining (immunofluorescence) (OM × 479). (I) Dermal CMRF-44+ DC (red, Peroxidase-AEC) and T lymphocytes (blue, Alkaline Phosphatase-Fast Blue) within a section of normal skin adjacent to a hair follicle (OM × 479). (J) Lymph node interfollicular (T lymphocyte) area containing CMRF-44+ IDC (brown, Peroxidase-DAB) compared with CD14 Mo and CD20 B lymphocytes (blue, Alkaline phosphatase-Fast Blue) (OM × 143). (K) Lymph node interfollicular region with CMRF-44+ IDC (blue, Alkaline phosphatase-Fast Blue) showing nuclear labeling for the transcription factor Rel B (brown, Peroxidase-DAB) (OM × 1,197). Bar = 10 μm.

Close modal

The Existence of a Lineage?

The original description of mouse spleen DC drew attention to their irregular shape with numerous cell membrane processes, including spiny dendrites, bulbous pseudopods, and lamellipodiae or veils.1 Distinctive electron microscopic features with a paucity of intracellular organelles and prominent mitochondria were also described. The presence of endosomes and lysosomes essential for Ag processing have been emphasized more recently.34,35 Mouse spleen DC adhere to plastic initially but become nonadherent after overnight tissue culture1 and this critical property, plus their low density, allowed them to be separated from MØ. Spleen DC were further distinguished from MØ on the basis of cytochemical reactions, a lack of phagocytic activity, and an apparent lack of Fc and complement (C) receptors, features which may be present in less mature DC. Similar cells were identified in lymph node (LN) preparations.2 Subsequent studies (reviewed by Steinman15 ) showed that DC express major histocompatibility complex (MHC) molecules, particularly class II loci products, in high density (mouse MØ have low-density MHC class II) and are exceptionally effective stimulators of primary T-lymphocyte responses, both in allogeneic mixed leukocyte reactions (allo-MLR)3 and Ag-specific systems.36 

Leukocytes with a high density of MHC class II molecules and irregular cell membrane processes were noted in the interstitium of most rat tissues, except the immuno-privileged sites of brain and testes.6-8,37 These interstitial DC were shown to be BM-derived, relatively poorly phagocytic, and to have similar characteristics to the lymphoid tissue–derived DC.7 Interstitial DC stimulate strong allograft responses13,14 and isolated interstitial DC stimulate T lymphocytes.24 Human interstitial DC were described in kidney initially9 and subsequently in other tissues.10 

Initial emphasis on their Fc and C receptors as MØ-like features delayed appreciation that Langerhans cells (LC) represent a stage of DC differentiation.33 Electron microscopy distinguished the Birbeck granule as a distinctive membrane inclusion within LC33 but this was difficult to follow technically and is rarely present in other DC types. When techniques were evolved to isolate LC from both mouse38 and human39 skin, cultured LC were shown to develop the morphological and phenotypic features of lymphoid tissue–derived DC. Functional studies confirmed that LC could stimulate a strong primary allo-MLR, but in some studies only after a significant period of culture.40-42 Similar allostimulatory properties were attributed to veiled cells and soon it was appreciated that these migrating DC gave rise to the interdigitating dendritic cell (IDC) within draining nodes.

Thus, the concept grew that DC have a growth and differentiation pathway (see A DC Hematopoietic Lineage and Migration Pathway), which interrelates with their migration24 and functional activation.15,16 BM precursors2,7 give rise to maturing DC, which can be distinguished from other myeloid cell types within human BM.44 Small numbers of DC precursors (and mature or recirculating forms) circulate in the blood and exit from the bloodstream into the tissues. These interstitial DC (LC form a subset in the skin33 and other epithelia) have extended cell processes and act as sentinel cells. After exposure to Ag or inflammatory stimuli, interstitial DC migrate out of the tissues into afferent lymph, where they are identified as veiled cells.45 Upon entering draining LN the DC are found as IDC in the T-lymphoid areas of LN. IDC are also found in tonsil and the white pulp of spleen.15,16,46 The spleen contains a second population of (red pulp, ie, blood-derived) marginal-zone DC.47 

Functional Activity of DC

Although DC lack substantial phagocytic activity, this relative deficiency has been over-emphasized.16 The lack of phagocytic activity is a feature of more differentiated or mature DC and less differentiated or immature DC have selective phagocytic activity48-50 (see Antigenic Material). In vitro studies show DC to have marked cell motility and the ability to extend/retract their cell-membrane processes.51 

The allo-MLR has been used widely to test DC Ag presenting activity.3,52 Although certain features (notably the high-responder T-lymphocyte precursor frequency) are not typical of a true naive primary immune response,53,54 the ability of DC to stimulate a potent allo-MLR distinguished them from other leukocytes. Activated B lymphocytes also have allostimulatory activity in vitro55,56 but are poor allostimulators compared with DC.57 58 

DC stimulate an autologous MLR,59-61 which may equate to a primary response to exogenous Ags (see Autologous Ag).

The extraordinary ability of DC to take up and present Ag for primary specific T-lymphocyte responses, in vitro62 and in vivo,63-65 is their main distinguishing feature.

The unique properties of DC become most evident when pure populations of DC, Mo, and B lymphocytes are compared.15,27 B lymphocytes probably contribute little to primary immune responses in vivo.66 67 DC are also active APC for secondary responses as are MØ, endothelial cells, and activated B lymphocytes.

Isolation of DC

A comprehensive review of methodology27 is available. DC purifications generally use one of two main approaches.

The first exploits the fact that DC lack certain lineage-specific markers expressed on other defined leukocyte populations (Fig 1A), eg, T-lymphocyte (CD2 [at least in high density], CD3), B lymphocyte (CD19, CD20, CD24), natural killer (NK) cell (CD16, CD56, CD57), Mo (high-density CD14), and granulocyte (CD15) markers. Various MoAb cocktails have been used with a variety of immunoselection systems to isolate “lineage negative” populations of DC from cell suspensions of hematopoietic tissue or after enzymatic digestion of other tissues such as skin.

The second approach has been to capitalize on the low density of DC, which results after a period of in vitro culture.68 Culture of peripheral blood mononuclear cells (PBMC) (often sheep red blood cell [SRBC] rosette negative), particularly in fetal calf serum at 37°C, activates DC and changes their size as well as their density, aiding purification over density gradients.17,69,70 Various gradient media have been used and the original bovine serum albumin (BSA) gradient71 has been replaced by metrizamide,51,72,73 Nycodenz (Nycomed Pharma, Oslo, Norway),68 or Percoll (Amrad, Pharmacia Biotech, Aukland, New Zealand).17,69,70 The Nycodenz gradient has proved very effective (Fig 1B).68 The gradient media all perform a similar function but these compounds and their osmotic effects may induce phenotypic or functional changes.25,74 For example, Kabel et al75 have suggested that metrizamide may alter Mo cells so that they resemble DC.

The major selective procedure in mouse DC preparation, transient adherence of DC to plastic,1,15 is ineffective in humans.16 Human DC, Mo, and other cell types appear to segregate variably after adherence to both plastic and teflon surfaces. Adherence enriches human Mo but this does not lead to reliable increases in stimulatory cells.16 In addition, adherence to different surfaces may influence the phenotype of Mo76 and perhaps DC.

Elutriation has been used27,75 but its application is limited by the availability of equipment. Methods that used other physical characteristics which may discriminate DC from other cells, eg, phagocytosis and clustering,27 have fallen into abeyance.

Sophisticated flow cytometric and immunomagnetic bead selection have now become the norm and a new era of positive selection with the CD83 and CMRF-44 reagents has made high-purity preparations of low-density DC feasible.77,78 An effective method to prepare DC for immunostimulation involves gradient separation over a standard density gradient to obtain PBMC, T-lymphocyte depletion (optional), overnight incubation for 16 hours in autologous serum/medium (±cytokines), and Nycodenz gradient separation. A significant lineage-negative, CMRF-44+ cell population (Figs 1B and 2) can then be isolated by positive selection.77 A similar approach using CD83 to select DC from a metrizamide gradient is also effective.78 

Fig. 2.

The costimulatory capacity of blood derived CMRF-44 purified DC compared with CD14+ Mo and CD19+ B lymphocytes. (a) The flow cytometry labeling of Nycodenz low-density cells is shown before sorting CMRF-44+ CD14 DC and CMRF-44 CD14+ Mo. (b) Allogeneic T-lymphocyte proliferative responses measured with 3H thymidine, are shown using each FACS purified cell populations as stimulators.77 The allostimulatory capacity of sorted DC, Mo, and B lymphocytes (CD19+ sorted) are compared at a ratio of 1 APC to 50 allogeneic T lymphocytes. (c) DC and Mo are compared in their ability to process and present the soluble protein antigens to KLH and tetanus toxoid to autologous T lymphocytes at an APC:T lymphocyte ratio of 1:20.

Fig. 2.

The costimulatory capacity of blood derived CMRF-44 purified DC compared with CD14+ Mo and CD19+ B lymphocytes. (a) The flow cytometry labeling of Nycodenz low-density cells is shown before sorting CMRF-44+ CD14 DC and CMRF-44 CD14+ Mo. (b) Allogeneic T-lymphocyte proliferative responses measured with 3H thymidine, are shown using each FACS purified cell populations as stimulators.77 The allostimulatory capacity of sorted DC, Mo, and B lymphocytes (CD19+ sorted) are compared at a ratio of 1 APC to 50 allogeneic T lymphocytes. (c) DC and Mo are compared in their ability to process and present the soluble protein antigens to KLH and tetanus toxoid to autologous T lymphocytes at an APC:T lymphocyte ratio of 1:20.

Close modal

Cytochemical Characteristics

DC lack many of the enzymes associated with classic Mo/MØ. Cytochemical reactions show that DC lack myeloperoxidase1,7,44,57,79 and have low levels of 5′ nucleotidase, dipeptidyl peptidase (DPP1), and cathepsin B activity.33,73,80 Other intracellular enzymes (eg, nonspecific esterase [NSE], acid phosphatase) or lysosomal Ags (eg, CD68)46 may be present in DC but their intensity is less or their intracellular distribution differs from Mo and MØ. NSE staining decreases with DC activation.81 

Problems Defining the DC Lineage

There are notable species differences. In mice and rats, BM and blood DC precursors express few MHC Ags and are poor stimulators in the allo-MLR.20,82,83 However, in humans, BM is allostimulatory44,84 and at least one or more fresh DC populations in blood are potent allostimulatory cells.52,57 85 Other surface molecules on DC have different patterns of expression between the species, eg, CD11c86 and CD187 (see Cell Surface and Other DC-Associated Molecules).

Inbred mouse and rat strains7,14 have different DC counts and their number may decrease with age.88 89 Informal reports (routine blood DC counts are just available) suggest considerable individual variations in DC numbers circulating in humans.

The different methods of purifying DC have considerable and variable effects on DC activation, which may modify DC phenotype and function. It remains a problem as to how to select for “resting” DC in human blood and thereby remove the contaminating small lymphoid population, stem cells, basophils, endothelial precursors, etc. Other cell populations, notably adherent cells, may be contaminated by numbers of highly activated DC.16 Thus, the original adherent preparations of Mo and MØ for APC studies were performed, not surprisingly, without regard for the existence of DC. Only recently, when CD14+ Mo were selected to exclude DC, did it become clear that Mo were poor APC compared with DC.57 80 

A Working Definition of DC

The putative DC lineage encompasses DC populations with very different properties, characteristic of their differentiation/activation state. Although DC were named because of their distinctive morphology,1 7 this feature is insufficient to define a DC. A definition of a DC must emphasize its ability to migrate and to stimulate a primary T-lymphocyte response. A consensus working definition might be:

  1. The ability to stimulate a primary T-lymphocyte response (this may require differentiation/activation of the DC population).

  2. Marked cell motility and the ability to extend and retract cell membrane processes freely at 37°C in vitro. The ability to migrate through tissues and track to the T-lymphocyte–dependent areas of lymph node.

  3. Relatively specialized phagocytic activity (in vitro) — DC uptake of extracellular material is probably greater than hitherto realized in vivo (see Antigen Uptake, Processing, and Presentation by DC). Active fluid-phase endocytosis is a feature of DC.

  4. Spontaneous initial and rapid clustering with T lymphocytes at 37°C in vitro.

  5. A cell-surface Ag phenotype distinguishing it from other leukocytes, notably Mo/MØ and B lymphocytes (see Cell Surface and Other DC-Associated Molecules).

  6. Expression of certain DC-associated Ags according to their differentiation/activation state (see Cell Surface and Other DC-Associated Molecules).

  7. Cytochemical reactions, which differ from Mo/MØ (see earlier).

Two additional features, “dendritic” morphology and high-density membrane MHC class II Ag, are confirmatory only (particularly in humans), as other cells, notably B lymphocytes46,90 and fibroblasts, can adopt a very similar appearance. DC morphology is also influenced by temperature.21 Note that FDC, dendritic epithelial cells, and lymphoid DC (for the moment) are excluded by these criteria.

DC express a repertoire of molecules common to other leukocytes, eg, MHC molecules, CD45 (leukocyte common) isoforms, adhesion molecules etc. These have dominated attempts to produce DC lineage specific MoAbs but persistence has resulted in some useful new reagents (Table 1).

Table 1.

Some Discriminating Leukocyte Differentiation Antigens Expressed by Human DC

Antigen/MoleculeBlood DC (uncultured)DC activated/Differentiated*Blood MoActivated B Lymphocyte
DC lineage 
CMRF-44 (+) +++ ± (+) ++ 
CMRF-56 − ++ − (+) (+) 
CD83 − − − (+) 
DEC-205 NAρ ++ NA ± ± 
S100 NA ++ ± NA − 
Mo/MØ lineage 
CD14 − − ++ ± 
CD68 +++ ++ − 
CD115 (M-CSFR) − − ++ − 
Myeloid/lymphoid    ++ 
CD1 − (++)1-155 − (+) − 
CD4 ± − 
CD5 ++ (+) 
CD13 − 
CD33 ± − 
Fc receptors 
CD64 ± − ++ ++ − 
CD32 ± ++ ++ 
CD16 − − (+) − 
Adhesion molecules 
CD11a (LFA-1) ++ ++ ++ ++ 
CD11b − (+) ++ − 
CD11c ± ++ (+) 
ICAM-1 (CD54) +++ +++ 
ICAM-2 (CD50) 
ICAM-3 (CD102) +++ +++ − 
LFA-3 (CD58) ++ ++ ++ ++ ++ 
Costimulator molecules 
CD40 ± ++ ++ 
CD80 − ++ − ++ 
CD86 − ++ ++ 
Leukocyte common 
CD45RA ± − − 
CD45RO − ++ − ++ 
MHC 
HLA-ABC ++ +++ ++ ++ 
HLA-DP +++ ++ ++ 
HLA-DQ ++ +++ ++ ++ 
HLA-DR ++ +++ ++ ++ 
Antigen/MoleculeBlood DC (uncultured)DC activated/Differentiated*Blood MoActivated B Lymphocyte
DC lineage 
CMRF-44 (+) +++ ± (+) ++ 
CMRF-56 − ++ − (+) (+) 
CD83 − − − (+) 
DEC-205 NAρ ++ NA ± ± 
S100 NA ++ ± NA − 
Mo/MØ lineage 
CD14 − − ++ ± 
CD68 +++ ++ − 
CD115 (M-CSFR) − − ++ − 
Myeloid/lymphoid    ++ 
CD1 − (++)1-155 − (+) − 
CD4 ± − 
CD5 ++ (+) 
CD13 − 
CD33 ± − 
Fc receptors 
CD64 ± − ++ ++ − 
CD32 ± ++ ++ 
CD16 − − (+) − 
Adhesion molecules 
CD11a (LFA-1) ++ ++ ++ ++ 
CD11b − (+) ++ − 
CD11c ± ++ (+) 
ICAM-1 (CD54) +++ +++ 
ICAM-2 (CD50) 
ICAM-3 (CD102) +++ +++ − 
LFA-3 (CD58) ++ ++ ++ ++ ++ 
Costimulator molecules 
CD40 ± ++ ++ 
CD80 − ++ − ++ 
CD86 − ++ ++ 
Leukocyte common 
CD45RA ± − − 
CD45RO − ++ − ++ 
MHC 
HLA-ABC ++ +++ ++ ++ 
HLA-DP +++ ++ ++ 
HLA-DQ ++ +++ ++ ++ 
HLA-DR ++ +++ ++ ++ 

Based on summarized data referenced in text.

*

Cultured overnight or longer in the presence of serum. IDC downregulate CD11c, CD13, and CD33.

Key: −, negative; ±, variable; +, positive; ++, strong positive; +++, extremely strong; ( ), subset only positive.

Mouse data only at present

ρ NA, not available.

F1-155

Langerhans cells.

Relatively DC-Specific Ags

The difficulties in isolating pure DC for immunization and screening of MoAb undoubtedly limited endeavors and hence the reagents available. It seems unlikely that DC lack unique membrane features and cell-surface molecules. Indeed, when sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) protein analysis of pure human tonsil DC proteins was compared with other leukocytes,91 DC appeared to have a very different protein composition, even if this level of resolution did not identify DC-specific molecules.

Mouse DC

The 33D1 rat MoAb92 identifies a low-density Ag on mouse (marginal zone) spleen DC. The antibody does not stain DC in cryostat sections and does not react with LC. No biochemical data on the Ag are available. Nonetheless, this antibody has proved extremely useful for C lysis of mouse spleen DC.

The hamster MoAb N41886 reacts with an epitope of the β2 integrin family heterodimer p150/95 (CD11c), which is expressed in high density on mouse DC and other leukocytes including MØ.

The rat NLDC-145 MoAb was raised by immunizing with mouse lymphoid tissue stroma and screening for reactivity on DC in tissue sections.93 It stains DC in T lymphocyte areas, including spleen white pulp but not spleen marginal-zone DC or LC. The Ag is also expressed on activated MØ and thymic cortical epithelium.93 Subsequent isolation of the Ag recognized by NLDC-145 and protein sequencing enabled a full-length mouse cDNA to be isolated.94 Now renamed DEC-205 (as a result of a revised molecular weight), the molecule, which has some similarities with mannose receptors,95 is postulated to act as a receptor for Ag uptake.94 

Other Animal Species

Another potential member of the integrin family that is preferentially expressed on rat DC is defined by the MoAb MRC OX62.96 The Ag is found on other cells but has considerable potential for defining DC.49 

Although DC have been studied in pigs, sheep, and monkeys, no reagents specific to DC in these species have been produced. Sheep and rabbit, but not mouse, DC express CD1.

Humans

The CD1 gene family includes at least three relatively well-characterized gene products: CD1a, CD1b, and CD1c, which have similarity with MHC class I molecules and are expressed by cortical thymocytes. LC express CD1a and variable amounts of CD1c87,97 while CD1b has been reported on dermal and migrating LC.98 

IDC in LN draining the skin may express CD1a, CD1b, and CD1c.99 Blood57,70 and tonsil46 DC do not express CD1a, although the former have been reported to both express CD1c78 and to lack CD1c.51 CD1 expression may be influenced by exogenous β2 -microglobulin levels in media.87 The CD1 Ags are inducible on Mo100 and probably on other DC populations in certain circumstances.18,101 102 

CD11c (p150/95) reagents have been investigated in humans as potential DC markers. This integrin has been described on a wide range of human leukocytes including Mo and MØ.103 CD11c reagents probably differ in the epitopes recognized,104 and have not proved particularly useful for identifying human DC. Nonetheless, CD11c expression on a subset of activated freshly isolated blood DC has been reported,78,85,105 although it is absent from tonsil DC.46 

Other reagents, including an MoAb, X-11, which detects a neo-epitope of C9106 and the “HLA-DQ–associated Ag” defined by RFD-1,107 also found on MØ, have received less attention. These Ags and the molecule present within LC identified by the Lag MoAb108 have yet to be characterized biochemically. An antibody, IRAC, which stains IDC in LN sections,109 has a wider tissue distribution by flow cytometry. Factor XIIIa, which identifies a dermal cell with dendritic morphology,110 has been suggested to be a marker for a tissue MØ population.104 

Two new reagents, HB15 (CD83) and CMRF-44, which react with DC membrane Ags are contributing toward better identification of DC in humans.22 23 These MoAbs and a third reagent, CMRF-56 (submitted), react with separate differentiation/activation Ags, which appear in higher density on cultured blood DC, as follows:

  1. The CD83 antibody HB15 has limited reactivity with activated B lymphocytes but stains cultured human blood DC,78 LC74, and some interdigitating cells in LN.23,74 The Ag is encoded by a cDNA that defines a member of the Ig gene superfamily.23 The mouse homologue maps to the MHC region.111 As yet it has no assigned function, although preliminary data suggest soluble CD83 may influence T-lymphocyte proliferation.

  2. The MoAb CMRF-44 reacts strongly with cultured human blood DC22 and identifies a small population of freshly isolated DC in human blood.77 The fact that these DC subsequently express CD83 upon activation77 suggests that these are partially differentiated/activated cells rather than recirculating cells. The CMRF-44 Ag is expressed in moderate density on isolated LC, consistent with the postulated DC differentiation/activation pathway.74 Although inducible on Mo/MØ by high doses of interferon-γ (IFN-γ),22 the CMRF-44 Ag expression on Mo/MØ is more limited in physiologic circumstances. Thus, the substantially higher density of CMRF-44 Ag expressed on DC allows positive selection of DC.77 The Ag is also weakly expressed on B lymphocytes. Biochemical studies have yet to fully characterize the Ag.

  3. The third reagent CMRF-56 identifies a DC differentiation/activation Ag of unknown function, which is not present on Mo/MØ (submitted).

Finally, the sequence for human DEC-205 has been obtained (submitted) and reagents to this molecule are eagerly anticipated.

Non–Lineage-Restricted Membrane Ags on DC

Broadly Expressed Molecules

DC express the CD45 Ag60 including the CD45RA, CD45RO, and CD45RC isoforms.51,78,112,113 The CD45RO is probably induced by activation.105 Signaling via the phosphatase active CD45 cytoplasmic portion may contribute to DC function.60 

The CD43 Ag (leukosialin) is less evident on human tonsil DC than on most Mo/MØ.114 Human blood DC preparations57 78 and perhaps mouse DC115 also show variable low-level expression.

DC also express CD48, CD148 (a molecule with phosphatase activity), and several other non–lineage-restricted Ags.78 113 

Myeloid Ags

In humans, most blood DC express CD33, an early marker of myeloid differentiation.44,80 LC and dermal DC express some CD33.116,117 Expression of CD33 appears to be reduced on tonsil DC,46,118 suggesting that it downregulates as DC differentiate. CD33 has been postulated to be a sialoadhesin119 but its function on DC is unknown.

CD13 is expressed on blood DC precursors but not on more differentiated DC.46,85,120 In contrast, tissue MØ are CD13+.103 

The CD14 molecule is expressed in high density on Mo and acts as a lipopolysaccharide (LPS) receptor.121,122 A soluble form is released by direct secretion and by cleavage of the phosphatidylinositol (PI)-linked membrane form which may bind to other cells. Low and variable levels of CD14 have been reported on blood DC80,85 while other flow cytometry studies suggest DC are CD14.57,77,78,123 The CD14 gene structure makes reverse transcriptase polymerase chain reaction (RT-PCR) analysis of CD14 mRNA difficult124 but the presence or absence of CD14 transcripts in DC needs clarification. Given that CD14 binds LPS,121 which triggers cellular responses, including DC migration,24 this is a pertinent question. The expression of CD14 can be modified by cytokines, notably IL-4125 and IFN-γ.126 CD14 bright cells have little allostimulatory activity57,58 but there may be a CD14+ stage of DC differentiation.84 127 

Human blood DC are CD36+ (thrombospondin receptor)78 and CD36 identifies a perivascular subset of probable DC in skin.104 CD36 stains subsets of Mo128 and MØ104 as well.

Human blood DC are weakly CD15s+78 but, like tonsil DC,46 CD15. DC do not stain with certain MØ-specific reagents, eg, CD115 (M-CSFR) and AM-3K.113 129 

Lymphoid Ags

Freshly isolated human blood DC probably express low levels of CD278,130 as do rat DC. However, cultured human blood DC,70,130 LC,117 and tonsil DC131 lack CD2 Ag. The CD5 Ag is found on blood DC.78,85,120 The CD4 Ag was identified on tonsil DC,46 LC,97 and recently on blood DC.105,120,132 Lymphoid-derived mouse DC express CD8.133 

Adhesion Molecules

Clearly, DC are involved in a number of adhesive interactions during their migration and subsequent interaction with T lymphocytes. The initial tethering and rolling of DC precursors on endothelium may prove to be mediated by selectins. Firmer DC adhesion and migration through vessel walls is likely to involve integrins and intercellular adhesion molecules (ICAMs) or other Ig superfamily members. Integrins, the CD44 Ag or the syndecans may mediate interstitial tissue interactions. The more stable DC interactions with epithelia probably involve the cadherins.

The Ig superfamily members ICAM-1(CD54), ICAM-2 (CD50), and ICAM-3 (CD102), which are all ligands for LFA-1(CD11a), are expressed on DC.134-136ICAM-1 is present in low density initially on blood DC but its expression is readily upregulated.135 ICAM-2 is present and its density seemingly changes little with activation.136 In contrast, ICAM-3 is present in highest apparent density as judged by flow cytometry staining and appears to be the most important LFA-1 ligand for initial DC-T lymphocyte adhesion.135 ICAM-1, which upregulates,137 and ICAM-3, which is not upregulated by culture in vitro, are also expressed by LC.138-140 

LFA-3 is a PI-linked protein that is expressed on DC and LC.134 This molecule probably has an adhesive function as well as signaling via T lymphocyte CD2.

Another family member, V-CAM (CD106), although not present on resting DC, is inducible at least on gut DC,141 possibly by IL-4 exposure.

PECAM-1 (CD31) is found on blood78 and tonsil DC.131 

Several integrin family members are expressed on DC. Certain β1 integrins (VLA subfamily α1-α9, αV ) are expressed selectively on LC.142 VLA-5 (fibronectin receptor) and VLA-6 (laminin receptor) may be involved in DC interstitial reactions. The β2 integrin, LFA-1 (CD11a), is expressed on mouse and human DC in high density.134,143 The expression of CD11b is limited to, at most, a subset of DC with low-density expression.57,85 CD11c appears to be present in higher density (N418 staining) on mouse lymphoid tissue–derived DC but it is also on a subset of human blood DC.85,105 CD11b and CD11c are also reported on human LC.104 144 The β3 subfamily (α11b) and the vitronectin receptor (αv) are probably absent from DC. Data on the β4-β8 integrins are awaited.

The cadherins are transmembrane glycoproteins that localize to adhering cell junctions. E-cadherin is present on both mouse145 and human146 LC. This molecule is expressed on blood DC and downregulates as LC migrate, suggesting an isologous adhesion interaction with skin epithelial cells.

The selectins are likely to be expressed on DC and may account for some of their trafficking. It is reported that human blood DC lack L selectin (CD62L),78 which is widely expressed on other leukocytes.147 However, it is shed after cell activation and thus perhaps during DC preparation. LC express the E-selectin ligand, cutaneous lymphocyte–associated Ag (CLA).148 The carbohydrate (CHO) moiety, sialyl Lewis X (CD15s), on DC is related to the ligands for E and P selectins.

The CD44 molecule is present in very high density on DC134 and the V3, V6, V9 isoforms are preferentially expressed on Mo-derived DC (Mo-DC).149A comprehensive analysis of CD44 isoforms in DC is justified, as CD44 acts as a ligand for hyaluronate, a connective tissue component likely to stabilize DC interactions with the tissue interstitium. CD44 may also have costimulatory functions.

The syndecans (CD138) bind growth factors and matrix molecules (fibronectin, collagen, laminin, and vitronectin) extracellularly and associate with the intracellular cytoskeleton. Thus, their presence on DC is of great interest.

Low levels of surface CD68 on DC may act as a ligand for E selectin.150 DC stain strongly for endoglin (CD105),78,113 a molecule found on endothelium, which may also have adhesion properties and act as part of the TGFβ receptor complex.151 They also express neurothelin (CD147), another endothelial expressed molecule with a suggested adhesion function.113 The fact that DC express the urokinase plasminogen activation receptor (CD87) implies this may be involved in tissue extravasation.113 Finally, CD36 may encourage DC binding to wound surfaces.152 

Potential Ag Uptake (C, Fc, pattern recognition) Receptors

Intact pathogens or antigenic components must bind to DC via opsonization (C and antibody) or other nonspecific recognition receptors.

Trace amounts of C receptors are detected on DC, particularly mouse24,40 and human153 LC. The CR1 (CD35) receptor is absent, as is CR2 (CD21),46 but trace levels of CD11b (CR3, see above) are documented. CD11c on DC may have potential as another C receptor. C5a receptors (CD88) are present on dermal DC and a subpopulation of LC.154 Blood DC are CD55+ and CD59+, providing some protection against C-mediated damage.

Early in their differentiation, DC probably express some CD32 (FcγRII),80,97,155 and CD64 (FcγRI)155 but no CD16 (FcγRIII).80 These may be downregulated with activation. A recent publication emphasized that the preparative method may affect FcR expression by DC and suggested CD64 and CD32 were both present in high density and functional (to a lesser degree than Mo).155 Human LC express CD32 and the high-affinity receptor for IgE (FcεRI)156 as well as the low-affinity FcεRII (CD23).157 The FcεRI on DC is a multimeric structure containing FcεRI α and γ chains but lacks the β chain present in basophils.158 A molecule with considerable homology to the poly Ig receptor, the CMRF-35 antigen, has been described159 and similar molecules are expressed on human blood DC, LC, and tonsil DC (submitted).

Nonspecific “pattern recognition” receptors exhibit binding specificities for structural patterns typically displayed by cell surface molecules of many microorganisms (eg, LPS and glycans) but not normally found on the surface of host cells.150Membrane-bound proteins, including the mannose receptor,95,160 DEC-205,94 or other lectin-like molecules,161 may act as non–self-recognition receptors on DC. Functional assays have recently identified the MØ mannose receptor on Mo-DC,162,163 although previous studies on rat DC were negative.164 The DEC-205 molecule on DC is predicted to have glycan binding activity associated with its CHO recognition domains.94 CD11c165 may be more relevant than CD14 as a DC LPS receptor. The scavenger receptors include the two class A MØ receptors SR-AI and SR-AII,150 belonging to a cysteine-rich (SRCR) family of proteins, which may be expressed on some DC.166 A third class A scavenger receptor described only on MØ to date, MARCO,167 remains to be studied on DC. The class B scavenger receptors CD36 and SR-BI are primarily lipid binding proteins150; the former is present on DC.78 The collectins are soluble serum proteins with lectin like specificity168 that may also interact with DC.

Costimulatory/Signaling Molecules (see DC Functional Properties section for functional data)

The CD40 molecule was first identified on human tonsil DC46 and then LC39 after activation in vitro. Blood DC express low levels of CD40 and upregulate its expression in vitro and after cytokine exposure.26 

CD80/86 Ags are not found on resting blood DC,25,169 may be on LC,170 and are upregulated rapidly upon DC activation (see DC Functional Properties section).

The heat stable antigen (HSA) identified in mice as a heavily glycosylated molecule with a small PI-linked protein core has been described on a subset of mouse spleen DC, LC,171,172 liver173 and thymic DC.174 The human homologue, CD24,175 was described initially as an Ag recognized by several MoAbs, many reacting with CHO epitopes. CD24a (IgM) antibodies react with a CD24 CHO epitope (contained within mucin) and CD24b antibodies (predominantly IgG) react with the CD24 protein core.176 Blood DC were found to be negative for CD24 using both CD24b MoAb and RT-PCR but specific staining with CD24a MoAb recognized CHO epitopes on a potentially novel molecule(s) (DC24) on DC.177 This reactivity was associated with a more immature DC phenotype.

Activation Markers

Several inducible (activation) surface Ags are found on DC. CD25 (IL-2R39,81 ) is induced on DC (cytokine receptors are discussed in Molecular Events Involved in DC Clustering and Signaling to T Lymphocytes section). Other markers, eg, CD98 (4F2)113 and CMRF-37178 are found on DC. The Reed Sternberg/Hodgkin cell (RSC/HC) associated CD30 Ag179 has not proved inducible on DC in limited studies to date. CD38 is expressed on tonsil DC (in preparation).

Inhibitory Molecules

Migrating LC undergo apoptosis in vitro.180 DC have also been shown to undergo steroid and UV light–induced apoptosis.181 They are predicted to express the Fas Ag (CD95)16,182 but confirmation of a functional effect on DC is awaited. Human DC express Fas-ligand. A subset of mouse CD8+ spleen DC express Fas-ligand.183 

Cytoplasmic Molecules

The hamster MoAb M342 stains an intracellular granule-associated Ag within mouse IDC and some B lymphocytes.47 Absent from freshly isolated spleen DC, the Ag is induced by culturing. Another MoAb, MIDC-8, stains similarly, but this Ag is not present in B lymphocytes.47 

The lysosomal-associated CD68 Ag (FAII in mice,184 ED1 in rats49 ) has been a moderately useful marker.185 CD68 staining has a limited (dot) perinuclear distribution in DC compared with a widespread cytoplasmic distribution in MØ.185-188 The S100 MoAb detects a series of intracellular isomers that were first considered specific to nerve tissues. Technical factors limit its use as an intracellular marker of activated DC for histological studies on formalin fixed section.189 190 

A 55-kD actin bundling protein that is induced in B lymphocytes by Epstein-Barr virus (EBV) infection has been shown to be expressed in the cytoplasm of the majority of human blood DC but not other leukocytes.191 Restin is a novel 160-kD cytoskeletal protein found in the RSC/HC and is associated with the intermediate filament cytoskeletal network.192 RT-PCR detected its expression by human DC (unpublished).

Novel DC-associated proteases that might be involved in processing Ag or other DC-secreted products (apart from chemokines or cytokines), which might contribute to DC migration via the extracellular matrix, may be anticipated.193 

Transcription Factors

It is logical that certain transcription factors (or combinations) will be associated selectively with the DC lineage. The early data reinforce the concept that DC form a unique lineage, with distinct transcriptional control of several genes.

Rel-B is expressed in murine thymus, spleen, and LN DC.194 Immunoblotting techniques have shown expression of c-rel, rel-B, NF-κBp65, and NF-κBp50 in DC grown from peripheral blood with granulocyte-macrophage colony-stimulating factor (GM-CSF ).195 Interestingly, these DC lacked the widely expressed transcription factor SP-1. The lack of CD14 on DC may relate to the fact that SP-1 is critical for the expression of CD14 on monocytes (Mo).196 Rel-B has also been identified in the Mo-DC197 and human tonsil DC (Fig 1K).198 

Fascinating studies with v-rel transformed chicken BM precursors suggested that when v-rel was induced, B-lymphoid cells resulted: suppression of v-rel caused the transformed cells to generate DC-like characteristics.199 The absence of rel B in gene-targeted mice is associated with a lack of functional mature DC (discussed later).200 201 

Novel DC Molecules

The search for novel molecules that play a role in the special functions of DC as an APC continues apace using both MoAbs and the new differential display technology. Several novel molecules expressed relatively selectively in DC have been identified, and further information on these will soon be available.

Immunophenotypic and functional analyses have defined the different DC populations (Fig 3). The interrelationship of nonlymphoid and lymphoid tissue DC has been inferred from the behavior of murine LC both in vitro and in vivo, the homing of DC injected into recipients and Ag tracking studies. The ability of DC to migrate into the tissues and from there to the LN is critical to their overall function as APC.

Fig. 3.

A putative hematopoietic differentiation pathway for myeloid and lymphoid DC. Lymphoid DC have different properties (?tolerizing) from the myeloid DC, which is immunostimulatory in most circumstances. The surveillance tissue-based DC:LC in the skin, DC in the respiratory tract, gut, or other nonlymphoid tissues, migrate after exposure to infection, tissue damage/inflammation, or antigen (ie, danger signals) via the afferent lymphatics to the T-lymphocyte–dependent areas of the draining LN. It is possible that epithelial based CD1+ DC have an independent derivation from the stem cell. The ability of Mo and Mø to convert to DC in vivo has yet to be established.

Fig. 3.

A putative hematopoietic differentiation pathway for myeloid and lymphoid DC. Lymphoid DC have different properties (?tolerizing) from the myeloid DC, which is immunostimulatory in most circumstances. The surveillance tissue-based DC:LC in the skin, DC in the respiratory tract, gut, or other nonlymphoid tissues, migrate after exposure to infection, tissue damage/inflammation, or antigen (ie, danger signals) via the afferent lymphatics to the T-lymphocyte–dependent areas of the draining LN. It is possible that epithelial based CD1+ DC have an independent derivation from the stem cell. The ability of Mo and Mø to convert to DC in vivo has yet to be established.

Close modal

Bone Marrow

BM isolated from mice20 and rats83 does not have constitutive allostimulatory activity. Mouse BM cultured in low21 or high concentrations82 of GM-CSF, with further mechanical (decanting) selection against granulocytic development, enables a DC-like population to emerge. Curiously, these cell preparations lacked significant expression of the mouse DC markers, NLDC-145 and N418,20 but were allostimulatory and homed to LN. Subsequent reports by others describe the generation of DC from mouse BM using GM-CSF and interleukin-4 (IL-4).202,203 OP/OP M-CSF–deficient mice204 have deficiencies in MØ populations but have been reported to have normal LC and DC networks,205 although LC were said to be reduced by 40% in number and abnormal in morphology.206 

In rats, conditioned medium, GM-CSF,83 and a novel cytokine combination including M-CSF with IL-3, linoleic acid, α tocopherol, and cholecalciferol,207 generated DC-like cells from BM.

Human BM contains allostimulatory DC-like cells.44 Further fractionation of the morphologically heterogenous allostimulatory “lineage negative” population of BM mononuclear cells established that some of the CD34+ progenitor cells were also allostimulatory.84 Putative precursors of LC-type DC have been identified in human BM18,102,208,209 using CD1a labeling. CD1a is inducible on Mo by GM-CSF100 and is not present on the uncultured BM allostimulatory population44 or fresh-blood DC.57 More recently the CLA Ag was shown to define a subpopulation of blood CD34+ cells, which appear to give rise to a CD1a+ Lag+ LC.210 

Analysis of human BM precursor growth in semisolid media has enabled DC-like cells to be defined morphologically,18,127,211 which when isolated had functional properties very akin to DC. The cells isolated by Young et al212 have virtually all the characteristics of DC and it is noteworthy that a significant proportion are CD1a.

Human BM cultured in liquid suspension84,213 with a range of cytokines has produced DC-like populations (see In Vitro Cultivation of DC and DC Lines). A CD14 stage of DC differentiation may emerge in vitro,84,127,213 although no CD14+ precursor was evident in directly sorted BM.44,84 The purity of the resulting DC-like populations has been hard to assess but these cell populations certainly include potent allostimulatory cells. Curiously, few CMRF-44+ or CD83+ cells have been identified in these first culture studies and prolonged culture in specific conditions may be required.84 127 

It is possible that a second DC differentiation pathway commits independently from a CD34+ precursor to provide Lag+ CD1+ epithelial-associated LC.210 

Galy et al214 suggested a human CD10+ lymphoid precursor gave rise to T, B, NK cells, and DC. In addition, a CD34+ CD38 (dim) human thymic progenitor has been shown to give rise to T lymphocytes, NK cells, and DC.215 These intriguing results will encourage more functional studies on putative lymphoid lineage DC.

Not surprisingly, SCF (c-kit ligand) improves yields of DC generated in vitro from human BM.211,212,216 It also increases DC (and MØ) colony yields in semisolid media.211 The alternative growth factor flt-3 ligand has been suggested to drive both myeloid and lymphoid DC growth in mice217 and facilitates human DC growth twofold.218 GM-CSF, either exogenous or endogenous, seems important in vitro for DC growth but lower doses than generally used may be sufficient.211 Curiously, both types of DC are found in mice in the absence of GM-CSF,219 suggesting other cytokines control DC production in vivo. IL-6216 and IL-4 almost certainly contribute to DC hematopoiesis. The latter downregulates CD14125 and suppresses MØ generation.220 Interestingly, IL-13 may reduplicate this IL-4 effect,221 as may IL-7. Tumor necrosis factor (TNF ) may contribute more to late DC differentiation/activation.

The high cytokine concentrations in vitro may drive aberrant costimulatory function and skew differentiation of committed Mo/MØ lineage cells. It will be reassuring to see culture conditions evolved that allow significant upregulation of “DC lineage” markers. Studies using CD34+ cord blood generated DC are discussed in the In Vitro Cultivation of DC and DC Lines section.

Blood DC

Mouse peripheral blood mononuclear cells generate DC-like cells when cultured in GM-CSF alone,19 suggesting a circulating DC precursor. These are destined to provide interstitial DC in nonlymphoid tissues, including the skin and liver.49 

Human peripheral blood DC were first isolated using methods similar to those used for the isolation of murine splenic DC.71 DC circulate in low numbers in blood and were difficult to isolate in quantity or high purity at this time.61 71 Most workers estimate DC (and their precursors) to represent only 0.1% of PBMC (a routine 300 to 400 mL blood donation provides 105 to 106 DC).

Highly purified human blood Mo, including a CD16 subpopulation,222 are poor stimulators in the allo-MLR compared with “lineage-negative” DC-enriched fractions.57,58 These “freshly isolated” blood DC57,58,105,155 or “immature” DC are smaller and lack dendritic morphology but have the nuclear features of DC (Fig 1A). A small population of larger cells are CMRF-44+ may be activated. These immature DC develop classic DC-like features after culture. It is noteworthy that these “lineage-negative” fresh DC preparations are not homogeneous and may include five or more cellular subsets.77 

Gradient separation of cultured T-lymphocyte–depleted PBMC removes contaminating small “lymphoid cells” (perhaps even some DC precursors) and with further cell sorting, relatively homogeneous DC preparations result, perhaps by coordinating the differentiation/activation of the DC precursor and CMRF-44+ DC populations. These have the following (MGG stain) cytological features: an indented nucleus with open chromatin pattern and rare nucleoli, substantial by a large pale blue cytoplasm with a prominent light Golgi (perinuclear) zone, minimal granules but some minor vacuolation, and more extensive membrane protrusions, veils, or dendrites (Fig 1B).

In the presence of human serum (ie, perhaps normal physiological conditions) human Mo differentiate into MØ,223,224 apparently distinct from DC. These results and the separation of circulating DC precursors (±an activated circulating DC population) from Mo suggest two separate lineages evolve from a common precursor (Fig 3). Recent data127 again support this concept but an independent CD1a+ intermediate precursor of LC remains a definite but elusive possibility.

As illustrated in Fig 3, a third differentiation pathway may also occur at least in vitro whereby, in certain conditions,223,225 nonproliferating Mo differentiate into Mo-DC (Fig 1D). This involves high concentrations (? nonphysiological) of cytokines in combinations, which may not occur in vivo. The critical differentiation event in this process appears to be cytokine downregulation of the M-CSF receptor, which allows a DC phenotype to emerge197 (see In Vitro Cultivation of DC and DC Lines).

Nonlymphoid Tissue-Derived Interstitial DC

There is an extensive network of interstitial DC encompassing virtually all organs except the brain, parts of the eye, and the testes.7,10,37,226 These cells develop from precursors in the blood7 and provide a sentinel system of APC. LPS,115 GM-CSF,227 IL-6, and probably other stimuli recruit precursors to the tissues. Mo-DC migrate in response to classical chemoattractants (formyl peptides and C5a) and some chemokines (MCP-3, MIP-1α, and RANTES) but not IL-8.228 A basal rate of tissue entry is boosted (5- to 10-fold) by tissue damage/inflammation, ie, “danger” signals.229 The kinetics of the DC response in the lung are rapid, similar to the neutrophil response. Freshly recruited mouse LC and other interstitial DC are N418, NLDC-145 but become NLDC-145 + in time and may express low levels of the macrophage-associated F4/80 Ag.

“Danger” signals, ie, tissue damage/inflammation, also mobilize DC from the tissues. After sensing inflammatory change230 or antigenic exposure,229 tissue DC migrate into afferent lymph.231 The properties of fresh LC and cultured LC28 (and fresh and cultured interstitial DC from other sites24 173 ) define two separate DC immune functions: (1) Ag acquisition and processing, and (2) Ag presentation to naive T lymphocytes in association with costimulatory signals.

LC

Epidermal LC form an extensive suprabasal network with branched cytoplasmic processes extended in physical contact with multiple adjacent LC.33 CD1a+ LC containing Birbeck granules are found in the dermis97 and may represent cells moving into the afferent lymphatic system. The number of LC (≈1,000/mm2 ) vary according to site.232 LC in the skin are subject to neuroendocrine control and are intimately associated with nerve endings. A low level of in vivo LC proliferation (? recently entered blood precursors) has been described.233 A population of CD1a dermal DC (immunohistology97 ) or CD1adim (immunofluorescence/isolated cells) is also present.74 Factor XIIIa positive dermal DC may be another potential subset110 among a substantial dermal MØ population.104 

The isolation of LC from skin is achieved by separating epidermal sheets from the dermis by short (<1 hour) incubations in trypsin39,42,138,234,235 or dispase.74 Epidermal cell suspensions are then produced by mechanical disruption and keratinocytes depleted. LC may be isolated at this stage, or after a period of culture, by density gradient centrifugation. Positive selection using CD1 or MHC class II MoAbs allows high purity LC preparations.235,236 Dermal DC are isolated by collecting the cells that migrate from dermal fragments.98 137 

As with blood DC, there are clear phenotypic and functional differences between fresh236,237 and cultured LC.39,42,234 Periods of culture, particularly in the presence of keratinocytes which produce copious cytokines, notably GM-CSF and IL-1, stimulate upregulation of MHC class II Ags, CD40, and allostimulatory activity. Low-level expression of functional Fc and C receptors on LC40 are lost on activation.231 Culture causes LC veiled processes to increase in number and length, whereas acidic organelles and Birbeck granules decrease or disappear. The characteristic CD1a marker also disappears. Normal skin contains only a small proportion of activated CD1a+ CMRF-44+ CD83+ DC (Fig 1I) mainly in association with T lymphocytes.74 CD1a+ DC have recently been identified in other epithelial surfaces such as the urothelium238 and gut epithelium.232 

Interstitial DC

First noted as strongly MHC class II positive cells with irregular membrane processes in rat tissue sections,6-8,37 these DC are distinguished from MØ by phenotype and radiosensitivity.239 Similar DC have been observed in mouse heart and kidney,43 although the level of MHC class II expression in the mouse appears reduced, hindering detection. In humans, DC have been documented in most organs, including liver,185 kidney,9 heart, and other connective tissue and tend to be associated with vascular structures.10 

Liver DC are preferentially associated with portal triads whereas Kupffer cells are located in the sinusoids. Their histological location suggests that DC enter lymphatics in portal triads or capsular lymphatics.185 DC in liver were poorly phagocytic cells compared with Kupffer cells,7 but recently divided BM-derived DC precursors enter the liver with phagocytic capacity for colloidal carbon.49 Rat liver DC have now been isolated and shown to have a relatively immature DC phenotype and functional capability,173 differentiating when exposed to type I collagen.

Interstitial DC in the heart (Fig 1H) show close association with small capillaries.7,239 They are not found in heart valves. Isolated mouse heart DC (N418 and NLDC-145) have limited phagocytic capacity and upregulate allostimulatory activity upon in vitro culture — characteristics reminiscent of the LC.43 

Interstitial DC in kidney are prominent in the cortex and are noted between renal tubules close to renal capillaries.6,9 They also occur in the medulla. Again they show limited phagocytic capability and upregulate allostimulatory activity in vitro.43 DC are found throughout the rest of the urogenital system, particularly associated with vascular structures below the urothelium. The presence of DC in the sclera and limbus of the eye, with a decreasing number of DC more centrally in the cornea, has been documented.226,240 Recently, DC have been isolated from the rat iris — a technical tour de force, and these too have similar properties to LC.241 

Interstitial DC are found in endocrine organs, notably thyroid, adrenal, and islet cells.8 These cells have no unique characteristics as yet but it is curious to note that thyroxine, vitamin D, and other medications influence DC function. DC were not obvious in nervous tissue7 but DC-like cells are reported in rat pineal.242 

Skeletal muscle and other supporting tissues all contain interstitial DC.10 Indeed, DC are found in major blood vessels, below the endothelium, and have been suggested to contribute to the atheromatous reaction.150 

DC are associated with the supporting structures of normal and abnormal joints. Arthritic synovial fluid is a source of human tissue-derived DC (see Autoimmune Disease).

Interstitial DC, like the epithelial associated LC, are believed to form a sentinel network of Ag-receptive cells, which subsequently move Ag centrally to provide activated APC and generate T-lymphocyte responses. Although both subpopulations are considered to parallel each other in terms of putative (? independent) DC differentiation pathways (Fig 3), the differential CD1a and CLA expression is notable and probably functionally important.

Mucosal Surface-Associated DC

An interdigitating sentinel epithelial network of DC has been described in the mucosa of the oral cavity, intestinal tract, and the respiratory tract, of mice, rat, and humans.243-246 These mucosal DC mature after weaning and their phenotype is responsive to external environmental influences.247 Inflammatory stimuli also recruit DC to the mucous membranes.248 Mucosal DC are also likely to have properties peculiar to the epithelial (external environment) surface that they serve.

Isolated oral mucosal DC express CD1a and contain Birbeck granules.243 CD1a+ cells are found in the oral and rectal mucosa.232 MHC class II positive DC have been identified in the epithelium of stomach, small and large bowel, lamina propria,244,249 and in the vagina and cervix.232 The draining mesenteric lymphatic DC population contains strongly MHC class II positive allostimulatory cells, with inclusion bodies indicative of a postphagocytic history.164 

The network of rat respiratory tract epithelial DC extends from the trachea, into the bronchi, to the terminal bronchioles,32,245,246 with their density decreasing toward the peripheries. It has been possible to prepare respiratory epithelium and isolate the DC that have similar characteristics to LC. DC have also been purified from human lung parenchyma on the basis of weak adherence, lack of FcR, and strong MHC class II staining.250,251 These cells are distinguished from alveolar MØ by surface staining and intracellular structures. Lung DC cannot be obtained by bronchiolar lavage, unlike MØ: their isolation requires tissue digestion,250 consistent with their interstitial location. Pulmonary epithelial-derived DC present Ag to T-cell hybridomas in both rats251 and humans.250 Lung interstitial DC stimulate strong allo-MLR responses whereas alveolar MØ do not; indeed, MØ may inhibit the MLR.251 

Afferent Lymphatic DC (veiled cells)

Various stimuli, including contact sensitizing agents, micro-organisms, and inflammatory mediators stimulate migration of DC from the skin, other nonlymphoid interstitial sites, and the mucosal surfaces into afferent lymph, where they are recognized as veiled cells.64,252-254 The agents, which mediate these effects, have all proved to be relatively nonspecific, eg, LPS, IL-1, and TNF-α.230,255 Thus, the activation of DC and the subsequent trafficking appears to be initiated by the Ag independent “danger signals.”229 Active changes in DC, dependent on tyrosine kinase activity, are involved in DC emigrating from tissues.171 Large mononuclear cells with veiled morphology were first isolated from pig45 and human256 lymph and constituted up to 20% of lymph cells. Peripheral DC migrate toward afferent lymphatics and traffic in the draining lymph to the LN where, after entering via the cortical sinus, the DC migrate to the T-lymphocyte–rich paracortical areas.257 This changes the DC homing characteristcs such that mouse spleen DC are unable to enter skin or heart transplants.258 

Skin-Derived Lymphatic DC

Microscopic studies on skin have identified LC migrating from the epidermis into dermal lymphatic channels.231 Veiled cells have been identified in skin lymphatics.256,259 In addition to their dramatic morphological appearances some cells remain CD1a+259 or, in the mouse, retain the NLDC-145 marker. Some appear to retain Birbeck granules.260 LC induced to migrate on contact with Ag present the Ag as the draining lymphatic DC.65,261 LC migrate out of mouse skin grafts forming cords of exiting NLDC-145+ cells in the lymphatics231 and migrate from human skin fragments placed in vitro.98 137 LC when reinfused home to the skin.

Nonlymphoid Organ DC

Visceral organs are presumed to respond to similar stimuli with an efflux of interstitial DC into draining lymphatics. The close association of DC with vascular structures make this likely and migration from rat liver to coeliac nodes has now been observed. Tissue DC migrate out of transplanted heart and kidneys into the circulation and the spleen.262 In rats peritoneal inflammation induces DC traffic into lymphatics.

Mucosal Surface-Derived Lymphatic DC

Similar cells are found in the afferent lymph from rat gut.164,263 Increased numbers of these cells are obtained by removing mesenteric LN and direct thoracic duct cannulation. They are also mobilized by LPS.264 The ultimate fate of these cells in the intact host is not yet clear.

Purification of lymphatic DC is readily achieved.27 Veiled cells show high-density MHC class II expression and can also take up Ag,265 form spontaneous clusters with T lymphocytes in vitro252 and stimulate resting T-lymphocyte proliferation after in vivo priming.265 They increase their allostimulatory activity and change surface phenotype upon culture.81 

Lymph Node and Tonsil DC

Practical issues have dictated that, apart from immunohistological studies in humans, functional data on LN-derived DC have been confined virtually entirely to other animals. Once again, it is prudent to consider LN draining skin, visceral organs, and mucosal (gut and respiratory) tissues as somewhat different. It has been assumed (there have been few direct comparisons) that the DC isolated from intact lymphoid tissue2 represent the IDC observed by light and electron microscopy266 and immunophenotyping.267 

Subsequent studies have isolated fluorescein- or Ag-tagged DC which have migrated from the site of Ag exposure.63 65 

The possibility of an alternative CD8+ DC subset in mouse LN DC was mentioned earlier.

Tonsils from routine surgery provide the most readily available source of human lymphoid tissue–derived DC from otherwise healthy donors. Cultured low-density tonsil DC were obtained with high purity (Fig 1C) and used for initial functional studies relating to the mechanisms of T-lymphocyte activation.268 Mo/MØ and “null” lymphoid population contamination is low27 and the omission of enzymatic tissue digestion precludes contamination with FDC. Cultured tonsil DC express high-density MHC products, low-level CD4, and certain CD14 epitopes detected by some MoAbs.46,269 They are potent APC in allo-MLR46,60,134 and oxidative mitogenesis assays.269 

Recent work has identified two DC populations within uncultured tonsil DC isolated using an immunoselection technique (in preparation). The majority of these CD13CD33 DC were not activated, ie, CMRF-44 and CD83 and the small proportion of CD80 and CD86 activated DC probably reflect the CMRF-44+ and CD83+ IDC noted in LN by direct in situ staining (Fig 1J). These CMRF-44+ cells express CD80 and CD86 in situ.169,270 The major population of “unactivated” DC could be identified within the T-lymphocyte areas as DC-24+, 2.7+ and CD38+ DC, but were CMRF-44, CD83, CD80, and CD86. A similar population of CD4+ CD11c, CD13, CD33 DC have also been described as corresponding to the “plasmacytoid T lymphocyte.”118 These populations may represent newly arrived IDC which have yet to upregulate these DC activation/differentiation Ags and costimulator molecules. Alternatively, they may represent IDC that have downregulated these molecules after interacting with T lymphocytes. This and other data74 localize the main events of DC costimulation to the lymphoid anatomical compartment. Some of these IDC may be relatively long-lived APC for stimulating T-lymphocyte memory, but the majority probably undergo apoptosis.118 A third population of CD4+ CD11c+, CD13+, CD33+ DC has also been identified within tonsil germinal centers.271 They had little capacity to phagocytose or take up soluble Ag but were allostimulatory. These DC are postulated to promote the T-B lymphocyte interaction within germinal centers and may regulate B lymphocytes.

Splenic DC

Given the spleen's different functions and dependence on a blood and not a lymphatic supply for cellular entry and exit, it seemed likely that splenic DC populations might have different characteristics.16 At least two subpopulations of DC have been defined in mouse spleen.47 The first, situated in the T-lymphocyte area in a periarteriolar distribution (ie white pulp), does not express the DC marker 33D1 or HSA but are NLDC-145+/DEC-205+ and can be induced to express M342 by culture.47 This population equates to the IDC described above. Direct visualization identifies a second population of DC in mouse spleen at the peripheries of the white pulp — the marginal zone. These cells have the inverse phenotype, ie, they are 33D1+ HSA+ NLDC-145 M342.47 LPS appears to induce the marginal zone DC to mature and migrate into and then out of the T-lymphocyte area.272 These marginal zone, 33D1, and HSA-positive DC may represent a circulating blood-derived DC more equivalent to human blood DC and comparative studies with other forms of lymphoid tissue–derived DC are merited. Indeed, GM-CSF generates DC from mouse splenocytes that resemble LC in phenotype and function.273 In contrast, different techniques isolate a (? lymphoid-derived) CD8+ DC subset from mouse spleen with different properties183 — their anatomical location is yet to be established.

Human splenic DC have been phenotyped in situ79 and significant regional variation occurs as in the mouse. Human periarteriolar (white pulp) DC lack lysozyme, NSE, FcR, and CD11b but are CD4+ and CD14±. The marginal zone cells appear to have a more MØ-like phenotype. Isolating DC from human spleen is difficult but possible,274 and the resulting cells are indistinguishable from cultured tonsil DC.

Gross phagocytic and proteolytic activity are more features of professional scavengers such as MØ.15 In contrast, relatively limited phagocytic/pinocytic capabilities provide DC with enough Ag to be processed into peptide for T-lymphocyte activation. After Ag uptake DC migrate, meet, and initiate membrane interactions with T lymphocytes and costimulate specific T-lymphocyte responses.

DC form distinct clusters275 with T lymphocytes at 37°C (Mo do not form stable clusters at this temperature) and these (Fig 1E and F ) are further stabilized as the interaction proceeds.36,59 Activated B lymphocytes have been reported to have different clustering characteristics with T lymphocytes.143 Analysis of DC-T lymphocyte clusters shows a high frequency of specific responding CD4+ T lymphocytes.

DC prime CD8+ T lymphocytes within clusters for cytotoxic responses and may do this via CD4+ help or independently of helper T-lymphocyte function.276-279 B lymphocytes are also incorporated into DC-T lymphocyte clusters and antibody secretion occurs in vitro.5 

NK cells were reported to regulate the mouse MLR280 but inhibition of allogeneic DC generated MLRs in humans by NK cells was inconsistent (unpublished results).

Stimulation of T-Lymphocyte Responses

Alloantigen

DC are considerably more effective than purified blood B lymphocytes and Mo, as stimulators in an allo-MLR.46,57,58 Tonsil B lymphocytes are often poor stimulators of an MLR.46 On the other hand, activated B (memory) lymphocytes, either LPS stimulated or EBV-derived human B-lymphoid lines, have significant allo-MLR activity55,56,143,281 and likewise, cytokine-activated and differentiated Mo and endothelial cells may acquire greater MLR activity. Indirect responder APC contributions are minimal.57 The allogeneic MHC- peptide complex may involve self-antigens, including processed self MHC molecules or acquired exogenous or endogenous56 viral Ags. Extensive DC differentiation and activation may occur during the time course of an MLR and it does not measure Ag uptake and processing.

In vivo data are discussed later (see DC and Transplantation).

Autologous Ag

Purified mature DC stimulate significant autologous T-lymphocyte proliferative responses — the autologous MLR.46,52,60 61 These “background” results have been downplayed but may be significant, eg, in peptide-based cancer immunotherapy (see later). It is assumed that the preparation of DC leads to activation and expression of endogenous autoantigens or exogenous Ags. Some reduction in autologous MLR occurs when foreign Ags are removed from preparative media. In mice, increased specific responses upon rechallenge with BSA support the notion that fetal calf serum is a significant sensitizing Ag.

Exogenous Ags

Purified Ags have been used to prime mouse DC to generate primary immune responses. In vivo administration of splenic DC after in vitro incubation with ovalbumin,62 sperm whale myoglobin,63 or hen egg lysozyme63,67 have been shown to stimulate strong Ag-specific responses when T lymphocytes from these mice are rechallenged with Ag in vitro. Similar results with fluoroscein isothiocynate65 and other Ags282,283 have been obtained using mouse LC. Pulsing DC with PPD or an immunodominant 19-kD protein derived from M tuberculosis primes for strong specific proliferative responses in vivo.284,285 A variety of tumor-derived peptide Ags have been used to prime DC in vivo (see DC Vaccination and Immunotherapy) for T-lymphocyte helper and cytotoxic responses. In vitro priming has also been studied, mainly in mouse models, and primary in vitro responses using peptide pulsed DC have been reported.286 

Strong cytotoxic responses to viral Ags have been generated in DC-immunized mice.287-289 Viral-specific cytotoxic T lymphocytes have also been demonstrated after in vitro priming with DC. In addition to influenza virus,290 mouse DC are effective APC for Herpes simplex virus,288 Moloney leukemia virus,289 and Sendai virus.289 

Strong humoral responses to keyhole-limpet hemocyanin (KLH) and SRBC have been shown in mice after priming with Ag and DC.291 

A variety of Ags have been tested in humans for their ability to generate DC-primed T-lymphocyte responses in vitro. Human DC exposed to KLH77,292 generate primary T-lymphocyte proliferative responses in vitro (Fig 2). Other studies have mainly used potential recall Ags: PPD,77,292-295 tetanus toxoid,52,77,292 Herpes simplex virus, C trachomatis,296 and C albicans,52 amongst others, to produce strong proliferative responses. Presentation of Leishmania48 and mycobacterial Ags295 by DC obtained from nonexposed individuals generates strong, almost certainly primary, responses.

DC elicit vigorous human CTL responses to influenza A virus from primed donors276 but their ability to generate primary CTL responses in vitro may be Ag dependent. Thus, primary CD8+ CTL responses to HIV in autologous DC-pulsed cultures297 may be variable but, in vitro, DC-generated cytotoxic responses to HIV proteins are possible.123,189 298 

Other Systems

DC are accessory cells for CD3 mitogenesis — a paradoxical and variable result, which presumably reflects either contaminating FcR-positive cells or validates reports of the expression of CD64 and CD32 on less differentiated DC. DC also stimulate in oxidative mitogenesis assays for APC function.71,73 269 

Finally, DC act as potent APC for bacterial superantigens that crosslink surface MHC molecules with responding T-lymphocyte T-cell receptor (TCR) complexes.299 

Inhibition of T-Lymphocyte Responses

There are at least three different scenarios in which DC may mediate alternative APC functions by deleting inhibiting or anergizing T lymphocytes.

Thymic DC

The majority of thymic DC are located at the cortico-medullary junction.30,300,301 Extensive mouse studies using congenic strains, transgenic animals, and artificial thymus cultures have established that a high proportion of early thymocytes receive a negative, apoptotic signal if the evolving T-lymphocyte TCR binds with high affinity to “self antigenic peptide plus MHC” on APC within the thymus.302,303 Thymic DC, which were until recently presumed to be of myeloid origin, are thought to be the APC responsible for the induction of tolerance. It is suggested that low concentrations of Ags from nonthymic sources are available to thymic DC to ensure systemic tolerance but there are problems with this concept.229 The outcome of thymic DC Ag presentation was thought to depend on the stage of T-lymphocyte differentiation (mature or immature) and not any intrinsic property of thymic DC. Both isolated splenic and thymic DC can initiate primary in vitro responses when they are cocultured with mature T lymphocytes. Conversely, splenic DC can induce T-lymphocyte tolerance after introduction into thymic lobes.31 

Lymphoid Precursor-Derived DC

Shortman et al have promoted the alternative concept that thymic DC are derived from a common lymphoid committed progenitor. The progenitors seed the thymus coincidentally in proportion with thymocyte progenitors29 and the resulting CD8+ thymic DC deliver tolerogenic signals.32,133,304 Mouse thymic DC appear to deliver a FasL dependent negative signal to CD4+ T lymphocytes183 and an alternative inhibitory signal to CD8+ T lymphocytes.32 Similar CD8+ DC have been isolated from peripheral lymphoid tissues, including mouse spleen.133 These DC, described as a subpopulation in lymphoid tissue preparations, may play a role in maintaining peripheral tolerance.

Human thymic DC have been isolated,305,306 but curiously they are CD4+ and CD8. An early thymic, apparently lymphoid DC progenitor has been described in humans.215 The CD10+ human lymphoid progenitor-derived DC observed in vitro214 may predict for a peripheral equivalent, possibly with immunosuppressive capabilities.

Peripheral DC Lacking a Costimulator Phenotype

The postulated differentiation and activation of myeloid DC into fully competent potent APC is a regulated event16 and there is considerable control of the expression of several crucial DC costimulator molecules.25,26 In a normal inflammatory/immune response, DC are thought to upregulate these molecules (Fig 4, top). However, DC costimulator activity is likely to remain low in the absence of inflammation229 and these cells may migrate into LN. As a result DC could tolerize or anergize T lymphocytes they encounter in the peripheries.16 Furthermore, DC may downregulate costimulator activity in certain circumstances, thereby rendering them ineffective as APC. Neonatal respiratory tract DC are also hyporesponsive.247 Further evidence that myeloid DC may be tolerogenic comes from recent data suggesting DC targeted with the 33D1 antibody induced specific T- and B-lymphocyte tolerance to rat IgG2b antibody.307 Immature DC, which lack costimulator molecules, may tolerize allograft recipients.308 309 

Fig. 4.

Molecular interactions between DC and T lymphocytes. DC undergo three or more stages to become fully active APC. Firstly, DC are activated by environmental danger signals. Secondly, initial adhesion and antigen-specific recognition takes place. Thirdly, full activation of DC costimulator activity follows, including cytokine production. This may be followed by programmed DC death.

Fig. 4.

Molecular interactions between DC and T lymphocytes. DC undergo three or more stages to become fully active APC. Firstly, DC are activated by environmental danger signals. Secondly, initial adhesion and antigen-specific recognition takes place. Thirdly, full activation of DC costimulator activity follows, including cytokine production. This may be followed by programmed DC death.

Close modal

Terminally Differentiated DC

Few DC are found in the efferent lymph. It is probable that central migration of Ag loaded DC and subsequent T-lymphocyte activation triggers DC terminal differentiation resulting in DC death.16 No direct observational data of DC death within LN has been reported but in vitro analysis suggests that DC are susceptible to apoptotic signals,181 after induction of Fas or alternative death signaling molecules. It is feasible that DC, which have interacted productively with T lymphocytes, receive late negative signals which render them ineffective APC.16 This would reduce multiple Ag processing and the possibility of inappropriate autoreactive APC function.

DC Depletion Studies

No natural model of DC deficiency exists. Immunodeficient nude mice15 and rats7 have normal or increased numbers of DC. Likewise, RAG1/RAG2 deficient mice, which lack T- and B-lymphoid cells, have normal DC numbers, including thymic DC.

Two transgenic mouse phenotypes, which include at least a partial deficiency of DC, have been described. The first310 takes advantage of the fact that the LTR promoter of HIV appears to be relatively selectively activated in DC. Mice transgenic for an LTR- thymidine kinase (Herpes-derived) construct delete 90% of their DC in the presence of the drug ganciclovir.310 These animals have reduced APC function and other immune deficiencies. A second model resulted from targeted200 and accidental201 gene insertion mediated disruptions of the rel-B gene in mice. This disrupts normal hematopoiesis and also the production of mature DC in the homozygous knockout animals. They have impaired cellular immunity in contact sensitivity experiments. Curiously, LC were not affected in either mouse model providing further evidence to distinguish this “immature” tissue subset from the “mature” lymphoid tissue–derived DC. Lack of functional rel-B affects T-lymphocyte production of IL-3 and GM-CSF311 and it is possible the absence of DC in rel-B–deficient mice is an indirect rather than a direct effect. Thus, LC might persist in these mice as a result of epithelial driven cytokine production.

These models, if they are sufficiently robust — the potency of the residual DC may cause difficulties — will provide important experimental opportunities.

The ability of DC to take up Ag and process it is highly dependent on the stage of DC differentiation. In vivo this relates to DC location. In vitro it relates to the time in culture, degree of activation, and antigen/T-lymphocyte exposure.

Antigenic Material

MØ may be more effective than DC as scavengers as a result of their portfolio of different nonspecific Ag (scavenger) receptors.150 In early studies, intravenous (IV) administered material such as colloidal carbon, peroxidase, and other Ags were found predominantly in MØ, although lesser quantities were noted in nonlymphoid tissue DC.2,7 Immunogenic fragments of foreign proteins are found in mouse spleen DC after IV administration.312 Phagocytosis of whole cells by DC in vivo has been reported,313 and DC draining the gut appear to contain phagocytic remnants.81,164 Liver DC precursors have phagocytic capability and migrate quickly (retaining immature cytological appearances) into the afferent lymph where they are no longer phagocytic.49 Several different investigators have now documented that administration of sensitizing Ag to the skin surface can be traced to LC and then to afferent lymph and IDC.65,312 Intradermal or mucosal administration of Ag primes afferent lymph DC in mice,67,314 rats,245,263 sheep,252 and humans50 to stimulate T-lymphocyte responses.

Several investigators have noted limited phagocytosis of large particles, ie, intact cells, bacteria, and synthetic beads of various types opsonized with a variety of compounds in vitro.155,315,316 BM-derived DC in mice are capable of active phagocytosis,284,315 but this has yet to be studied in humans. LC have been reported to take up bacteria including staphylococci and Leishmania,48,317 although this may vary between mouse strains. Moll et al318 described LC transporting Leishmania from infected skin to the draining LN. Viral particles can be endocytosed by sheep lymph-derived DC,50 a result which contradicts the findings reported with rat liver lymphatic DC.49 

MØ may contribute by lysing large particles, eg, bacteria, viruses or tumor cells before transferring the degraded products to DC.63,268 MØ-DC cooperation in presenting mycobacterial Ags has been reported in some295 but not other situations.319 

The size cut off for fluid phase Ag uptake (endocytosis, pinocytosis) is conventionally taken to be less than 0.5μm.49 Freshly isolated DC, ie, blood DC or LC, are highly active in this regard. Mouse and human LC bind surface Ag and internalize at least small portions of their plasma membrane.320 Direct visualization of fluoresceinated Ag has been observed within mouse LC and DC.62,65 Tonsil DC show considerable BSA uptake, whereas B-cell lines take up little by comparison.91 This active fluid phase endocytosis almost certainly explains why DC acquire a low density in vitro. The Mo-DC generated by GM-CSF/IL-4 treatment of PBMC take up extracellular fluid actively processing approximately 100× their volume per hour.162 

Nonspecific inflammatory triggers may be required to initiate DC phagocytosis, pinocytosis and even expression of “antigen receptors,” eg, DEC-205. Equally, the downregulation of DC phagocytosis/pinocytosis upon exiting the tissues46 makes good sense as this reduces the risk of extraneous Ag presentation by DC. It may also switch the cellular machinery of the DC into Ag processing and presentation phases in preparation for T-lymphocyte costimulation phases.

Ag Uptake

Preliminary data suggest DC use a variety of membrane receptors for taking up Ag. The resulting receptor-antigen complexes may traffic between cell types321 and perhaps even from MØ to DC. “Cross-priming,” ie, Ag transfer from one cell to APC, may complicate interpretation of in vivo experiments.

FcR

The FcR repertoire on DC was described in the Potential Ag Uptake (C, Fc, Pattern Recognition) Receptors section. Specific antibody presumably acting via these receptors facilitates APC function of sheep lymphatic DC.252,322 Antibodies that bind to isolated LC are internalized.323 Fanger et al155 have used bispecific antibodies to confirm that CD64 and CD32 stimulate human blood DC to phagocytose ox red blood cells but at less than half the rate of Mo. A secondary tetanus toxoid T-cell line response was increased when specific antibody was added to Mo-DC and Ag.149 Culture of blood DC and LC downregulates their FcR and this presumably prevents further Fc Ag loading.

C Receptors

The trace levels of receptors were described in the Potential Ag Uptake (C, Fc, pattern Recognition) Receptors section. C components on the surface of organisms act as opsonins for the uptake of material in LC.33 C5a receptor (CD88) on DC have not been tested for Ag uptake.154 No data on the function of CR3 or indeed CR1 and CR2 on fresh blood DC are available but these receptors downregulate upon culture and activation of LC.324 

Pattern Recognition Receptors

Bacteria express a large number of complex CHO molecules on their surface which may bind to pattern recognition receptors,150 many of which have a lectin-like specificity. Mannose containing residues bind to a mannose receptor protein present in high density on MØ,95,160 but there are subtleties as to the exact sugar specificities of binding, indicating several receptors may be involved.325 Related mannose binding proteins circulate in plasma.168 The mannose receptor binds bacteria and stimulates phagocytosis by MØ.160 Studies on mouse DC suggest mannose-like receptors are present317 and recent experiments in humans show that rapid uptake of mannose-conjugated BSA by GM-CSF/IL-4–differentiated Mo-DC is inhibited by mannose.162 

A rabbit antiserum to mouse DEC-205 stimulated an antibody response to rabbit Ig94 and was interpreted to suggest active receptor uptake by DEC-205. The relative efficacy of this DEC-205 and CHO specificity for loading into DC compared with other surface receptors remains to be determined. The function of the other scavenger receptors,95 326 described earlier, on DC has yet to be established.

MHC and Other Membrane Cycled Molecules

Cycling of surface MHC class I molecules (the class I groove is closed) as an Ag uptake mechanism is thought to be very limited, although some peptide exchange may happen at the surface. The CD1 molecule may be associated with Ag uptake, particularly nonprotein Ag.320,323,327 328 

A proportion of DC surface MHC class II molecules may be incorporated into a second and probably minor endocytotic class II compartment.35,329,330 Some Ags enter and bind in the open-ended surface MHC class II groove as peptides or even intact molecules. Thus, certain rare Ag, eg, fibrinogen, may be presented without processing by fixed DC.331 Bacterial superantigens binding alternative sites on MHC class II molecules may also be incorporated into the DC.

Ag Processing

Endogenous Ag is degraded by intracellular proteases and peptidases to provide peptides in the cytosol. These peptides, bind to transporter (TAP) gene products and are delivered to the MHC class I compartment to be incorporated in the nascent peptide binding groove generated by folding of the class I molecule before moving to the cell surface. This does not occur in the presence of proteasome or Golgi transport inhibitors and defines a conventional pathway. In DC, several additional mechanisms also operate allowing exogenous Ags to be presented on MHC class I molecules.332 Firstly, it is now clear that DC take up Ags by macropinocytosis and release Ag into the cytosol for classical TAP-dependent MHC class I presentation.333 Secondly, peptides produced in high concentration externally may exchange with peptides on mature MHC class I molecules. A third mechanism may involve carrier molecules or chaperones such as hsp96, capturing peptides and delivering then via the DC surface to the endoplasmic reticulum, where they are released, processed, and incorporated into MHC class I molecules.

For influenza Ags at least, material synthesized endogenously is presented more effectively via DC MHC class I molecules, although exogenous material can also be presented.276,290 Soluble protein, eg, albumin, is presented by DC to class I–restricted T lymphocytes.333 Peptides are readily presented by DC, providing good evidence that these alternative pathways operate in DC at least to some extent.123 

Exogenous Ag, eg, intact organisms, particulate Ag or soluble Ag entering DC via a phagocytic or endocytic pathway, must undergo degradation and proteolysis before interacting with the first and major, MHC class II compartment, where peptide loading of class II molecules occurs. Class II α and β chains are synthesized and combine to be transported in association with HLA-DM to a class II compartment where DM dissociates. The class II molecules are membrane-bound with their peptide binding site protected by class II invariant protein (CLIP). Further maturation of these vesicles occurs after fusion with the endocytic lysosomal pathway (CD68 and acid phosphatase staining), prominent in DC.50 This lowers the pH and invariant chain (peptide?) dissociates, allowing appropriate peptides to bind with the MHC. This pathway has been studied recently in both mouse and human DC in a series of elegant immuno-electron microscopic studies.34 35 

Levine and Chain334 have used a flow cytometric assay to show that fluid-phase endocytosis, especially traffic through late endosomes, is as active in DC as in other APC. Previous studies of endocytosis on DC used long pulses of Ag and therefore emphasized the accumulation of Ag in a terminal degradative organelle rather than traffic through late endosomes and the related processing organelle.334 Enzymatic degradation of proteins into peptides and their association with class II molecules probably takes place in acidic lysosomes as chloroquine blocks DC pulsing with certain protein Ags.234 Fresh LC contain many lysosomal and Ag containing membrane bound vesicles, which disappear with culture,329 coincidental with loss of Ag processing capacity. In contrast, differentiated murine spleen DC have a limited lysosomal system.1 The association of novel Ags with mature DC intracellular vesicles (and B cells)47 indicates specialized function for these APC organelles.

Tonsil DC degrade ingested BSA.91 Chain et al335 have likewise demonstrated proteolysis of exogenous protein by DC. Several soluble Ags have been processed by purified DC to be presented effectively to peptide specific T-lymphocyte clones336-339 or to generate primary or secondary in vitro responses.77,123,292,340,341 Preparations of mouse DC have been shown to degrade antigenic material from the surface of beads without ingestion and process it effectively (possibly externally) for a T-lymphocyte response.337 DC may express cell-surface peptidases, eg, CD13, but these have yet to be studied. Serum proteases may contribute by digesting proteins to peptides342 for DC binding.

Finally, other leukocytes may contribute to the proteolytic degradation of Ag prior to recovery by DC.268 

MHC-Peptide Presentation

DC actively synthesize MHC class I molecules and mature DC show high-density surface expression.15 Again, MHC class II molecule synthesis is active in fresh blood DC35 and LC62 and surface MHC Ag expression increases markedly on culture, eg, LC in GM-CSF.38,343 It is noteworthy that low-dose GM-CSF decreases synthesis of class I in mouse MØ.21 IL-4 also upregulates MHC class II production125 as of course does IFN-γ. Isolated mouse liver DC are MHC class II and require interaction with collagen to induce its expression.173 After in vitro culture DC synthesis of MHC Ags is reduced,344 however, cultured LC can passively acquire Ag fragments that do not need processing.234 Indicative of their specialized function, DC retain class II antigenic peptide complexes for prolonged periods (1 to 2 days) in culture,62 whereas MØ have a turnover measured in hours.345 The biochemical explanation for a lesser glycosylation of DC MHC molecules346 and any functional consequence remain elusive.

An early report suggested CD1b presented glycolipid to the T-lymphocyte TCRαβ328; other data suggested a special role for peptide presentation by CD1b to T lymphocytes with TCRαβ.347 CD1a and CD1c has also been postulated as a target recognition structure for TCRαβ348 and TCRγδ T lymphocytes.348 349 

Once DC have phagocytosed/pinocytosed Ag, processed it, and delivered the antigenic peptide to surface MHC molecules, they are unable to process and present new peptides.324,343 Only a few hundred MHC-antigenic peptide complexes amongst the several million molecules on the DC surface may be required for a specific response324 and occupancy of only 0.1% of DC surface class II molecules with superantigen is sufficient to induce T-cell proliferation.299 Having achieved this, DC mobilized by “danger” signals (from Ag receptors) prepare for the next phase of molecular events: migration into LN and clustering with T lymphocytes.

The formation of DC-T lymphocyte clusters in vitro only occurs after multiple transient DC-T lymphocyte interactions.60,275,350 It is an active process and may start in the afferent lymph.187 Cluster formation requires an intact DC cytoskeleton and protein kinase C activation350 to initiate adhesion interactions.60 This is initially an Ag-independent process275 but, once cognate-specific T-lymphocyte recognition occurs, additional events stabilize the cluster (Fig 4).

The molecular interactions from this point are complex and their kinetics may also contribute to the outcome. DC appear to upregulate costimulator activity as a result of Ag independent stimuli in vitro, but Ag-dependent control of DC antigen-presenting function may be more relevant in vivo.

DC must be viable to initiate T-lymphocyte activation. Thus, DC “fixed” with paraformaldehyde early in differentiation lack the ability to change their cell membrane molecular constitution and cannot costimulate T lymphocytes.331 However, activated or cultured DC expressing a full repertoire of costimulator molecules (and surface peptide) can stimulate after fixation.331 Soluble factors derived from DC alone are clearly insufficient to drive T-lymphocyte responses.59 

The term “accessory” molecules describes the additional signal(s) (signal 2) apart from Ag (signal 1) required to activate a T lymphocyte. This includes essential adherence interactions but also other critical costimulator molecules. Several other signals may contribute to regulating this process (see Fig 4). Specific TCR binding to MHC-antigenic peptide might provide a permissive signal to DC. The DC also receives reciprocal signals from other T-lymphocyte membrane ligands. Cytokine signals either from the T lymphocyte or from elsewhere in the environment will also influence DC function. Data26,74 135 now suggest that T-lymphocyte recognition of specific Ag effectively converts DC into relatively Ag-specific APC (Fig 4) — somewhat of a paradigm shift for immunologists.

The character of the T-lymphocyte response may be influenced directly by the DC or by the local environment (eg, other cells and/or cytokines).

Adherence

The DC–T-lymphocyte interaction is stabilized by adhesion molecules.59,70,134,135,269 ICAM-3, which is constitutively expressed on DC in high concentrations, provides the predominant initial adhesion interaction with LFA-1 on T lymphocytes.135 ICAM-2 on DC does not appear to contribute much to this particular adherence interaction.136 ICAM-1 is expressed in relatively lesser amounts on resting DC but is induced rapidly on DC as a result of retrograde signaling from the T lymphocyte,135 thereby stabilizing the cluster.

LFA-3 on DC may also increase cluster stability by binding the CD2 ligand on T lymphocytes.134,269 LFA-1 on DC participates in reciprocal interactions. The relative paucity of CD43,114 a major contributor to cell-surface change, may facilitate DC membrane contact with T lymphocytes. Other adhesion molecules present or induced (eg, VCAM) on DC may play a part.

Costimulator Molecules

The term “costimulation” is used to describe the additional signals required to initiate T-lymphocyte activation apart from an adhesion interaction.351 Without “costimulation,” specific Ag triggering of T lymphocytes may lead to anergy or cell death.351-354 

CD80 (B7.1) and CD86 (B7.2)

An excellent review in Blood describes the properties of these two members of the Ig superfamily.354 They bind different sites on the CD28 molecule (present on 80% of human T lymphocytes), thereby delivering a significant cyclosporin A–independent costimulatory signal. A second CD80/86 ligand, CTLA-4, is expressed on activated T lymphocytes and provides an alternative negative signal.355 Differential T-lymphocyte responses are attributed to CD80 stimulation (TH1 ) or to CD86 stimulation (TH 2 ) in vitro and in vivo.356 

Resting human blood DC lack CD80 mRNA and surface Ag.169 They also lack CD86 surface Ag but have preformed CD86 mRNA present.25 This probably accounts for the rapid surface upregulation of CD86, followed by CD80 mRNA upregulation and subsequent surface expression of both Ags after culture.25 Isolated human LC express slightly increased amounts of CD80/CD86 expression compared to blood DC. This may relate to increased differentiation/activation,74,170,357 or simply be a consequence of the isolation procedure, as sensitive in situ staining suggests LC lack CD80/CD86.74 Similarly, LC isolated from mice lack CD80/CD86.358 Culture of human blood DC,169,292,359,360 human LC,170,357,361 and mouse LC358,362 leads to relatively rapid induction of CD80/CD86. Finally, interstitial DC isolated from rat liver173 likewise appear to lack CD80/86.

The stimuli that control relative CD80/CD86 expression on DC are being clarified. GM-CSF induced both CD80 and CD86 on mouse LC, whereas IFN-γ only upregulated CD86.358 IL-4 induced both molecules on mouse LC.363 Curiously, in vitro generation of CD34+-derived DC-like cells induced CD80 before CD86360 but, in all other contexts, CD86 expression precedes that of CD80. Crosslinking of MHC class II molecules on DC may provide a retrograde signal upregulating CD80/CD86 expression. In contrast, IFN-γ inhibits CD80 expression on mouse LC363 and IL-10 has been reported to prevent CD80/86 expression on LC363 and spleen DC.287 IL-10 suppresses the LC-induced MLR.331 Similar inhibition of DC but not Mo induction of CD86 occurs with IL-10.364 

Although both CD80 and CD86 deliver a strong costimulatory signal via CD28,354 their effect via CTLA-4 (after activation it moves from the cytoplasm to T-lymphocyte surface) is unclear. Blocking CD86 inhibits the MLR to a greater extent than blocking CD8025 and CD86 blockade reduces DC primary Ag specific responses.292 

CD40

The CD40 molecule is a member of the TNF receptor (TNF-R) family. Its ligand CD40L, a TNF-α–like molecule, is expressed on T lymphocytes within a few hours of their activation.365 The CD40L on T lymphocytes provides an essential costimulatory signal for CD40+ B lymphocytes. The CD40 molecule was described on human tonsil DC,46 and subsequently noted to be present in low density on resting human blood DC26,51 and LC.39 It is rapidly upregulated after a brief period of tissue culture. In humans the cytokines IL-1, IL-3, GM-CSF, and TNF-α increase CD40 expression, whereas IL-2, IL-4, IL-6, IL-12, and IFN-γ26 do not. CD40 is upregulated on CD34+-generated DC.366 An analysis of human CD40:CD40L interactions suggests two functions for the CD40 molecule on DC.26 Firstly, retrograde signaling via CD40 was the most potent stimulus identified to date for upregulating CD80/CD86. Secondly, despite this effect, soluble CD40L and CD40-Ig inhibit DC stimulation of allogeneic T lymphocytes,26 suggesting that the DC CD40-CD40L T-lymphocyte interaction provided an additional costimulatory signal. Significant additional inhibitory activity occurred when both CTLA-4Ig (blocking CD80/CD86) and CD40-Ig were used to block a DC stimulated MLR.

HSA/CD24

An MoAb able to block costimulator function was shown to recognize HSA367 and costimulatory activity was attributed to HSA on mouse LC.172 Mouse HSA transfectants stimulated T lymphocytes and double-transfection studies inferred HSA and CD80 had synergistic costimulatory activity.368 The human homologue CD24 is not present on DC, and similar experiments produced no evidence to support the view that human CD24 has costimulator function.369 

OX40L

OX40, another member of the TNF-R family,370,371 is present on activated peripheral CD4+ and a subset of activated CD8+ T lymphocytes. The OX40L is found on EBV-transformed B lymphocytes372 and CD40L-stimulated B lymphocytes.373 It costimulates OKT3, phytohemagglutinin (PHA), or phorbol 12-myristate 13-acetate (PMA) activated T lymphocytes372 and it seems likely that the OX40L:OX40 binding contributes to DC:T-lymphocyte interactions.

4-1BBL

4-1BB is expressed on activated thymocytes and T lymphocytes.374-376 PHA activation of T lymphocytes induces 4-1BB within a few hours.377 The 4-1BBL has been identified on LPS-activated BM MØ and on anti-Ig activated B lymphocytes.375,378 Transfectants expressing 4-1BBL costimulate CD3 triggered T lymphocytes and mimic the effect of antibody crosslinking 4-1BB on T lymphocytes.376,379 4-1BBL is inducible on B-cell lymphomas and acts as an effective costimulatory molecule.380 Again, it seems likely that DC express 4-1BBL, if not 4-1BB, as well.

CD6 Ligand

Crosslinking studies suggests that CD6 can function as an accessory protein in T-lymphocyte activation.381,382 A ligand for CD6, ALCAM (activated leukocyte adhesion molecule), has been identified.382 ALCAM has 5 Ig-like external domains and mediates cell adhesion to CD6. It is not yet known whether this ligand pairing provides a costimulatory interaction. However, as ALCAM is induced on activated Mo, whether it is expressed or not on DC is a relevant question, particularly as CD6 antibodies can influence the autologous MLR.383 384 

SLAM Ligand

A novel 70-kD glycoprotein designated SLAM (CDw150) has recently been described as yet another member of the Ig superfamily, which has limited similarity with CD48 and LFA-3.385 SLAM is constitutively expressed on peripheral blood, memory T lymphocytes, and induced on naive T lymphocytes after activation. It has recently been identified on blood DC and upregulates with activation.113 Monovalent antibody bound to SLAM enhanced Ag-specific T-lymphocyte proliferation and cytokine production.

MHC Signals

Complete or partial activation of the responding T lymphocyte has the potential to provide molecular feedback by T-lymphocyte membrane molecules or secreted products, ie, cytokines (Fig 4). The DC MHC/peptide:TCR/CD4 or CD8 T-lymphocyte interaction has been suggested to provide a retrograde signal to DC after cognate recognition of Ag by the TCR.16 Data have accumulated to suggest the cytoplasmic portion of MHC molecules, notably class II, mediate intracellular signaling events in other cells281 386 and preliminary data suggest that class II signaling upregulates CD80/CD86 on Mo. The effect of crosslinking MHC class II molecules on DC costimulator molecule expression is under investigation.

Cytokines and Chemokines

The issue as to which of the many immunologically important cytokines (reviewed elsewhere) are expressed by DC remains a somewhat vexed one. Conflicting data have arisen because the purity of some DC preparations may have allowed significant cytokine release from contaminating populations. Even more relevant, different states of differentiation/activation, particularly DC phagocytosis/pinocytosis, may profoundly influence DC cytokine production.

IL-1α and IL-1β

For a long time, IL-1 was credited (inappropriately) with providing a second signal from Mo or MØ as APC for primary immune responses.

Findings in the mouse suggested that DC did not produce IL-1387 and raised the alternative possibility that DC themselves (notably in the thymus) were responsive to IL-1 and there was an indirect effect of IL-1 on DC-T lymphocyte priming. Neither IL-1α nor IL-1β is produced in significant amounts by human tonsil or blood DC as assessed by immunolabeling, bioassays, or enzyme-linked immunosorbent assay (ELISA) analysis.16,268,388 RT-PCR analysis shows limited IL-1α and β transcripts in both resting blood16,389 and tonsil16 DC, with the latter being more obvious. It should be emphasized that, as yet, these more sensitive studies (which have the same DC purity caveats) have not demonstrated IL-1 secretion or function. Despite this, it is possible that DC within tissues produce IL-1. Murine LC are thought to produce IL-1β but show differential stimuli dependent release in culture.390 Some data in humans implies that in vitro–generated DC and stimulated blood DC and LC may produce IL-1.366,389,391 Stimuli such as phagocytosis315 or phorbol ester/calcium ionophore366 may be required to induce IL-1 secretion by DC. It is also clear DC produce large quantities of IL-1R antagonist mRNA, but the functional outcome of this is unknown.16 

TNF-α

TNF-α secretion by mouse DC has not been studied in detail, although human studies suggest there are low levels of TNF-α mRNA in blood389,392 and tonsil DC.16 No biologically detectable TNF-α392 was secreted, but this may vary with the stimuli used. DC-like cells generated from CD34+ cells produce TNF-α when stimulated via CD40.366 TNF-α antiserum inhibited a DC stimulated allo-MLR,392,393 but this neutralized the TNF-α produced by activated T lymphocytes, which acts as a significant autocrine factor.392 Lymphotoxin (TNF-β) has limited expression in DC at the mRNA level389 but the significance of this is unclear.

GM-CSF

Small amounts of GM-CSF mRNA have been reported in human blood16,389 and tonsil16 DC, but evidence that this results in secretion of active GM-CSF is awaited.

IL-6

Mouse LC produce IL-6394 and DC are thought to be the major source of IL-6 in the mouse LN. Northern blot analysis failed to detect IL-6 mRNA388 but RT-PCR detects small amounts in human blood DC16,389 and larger amounts in tonsil DC.16 Although IL-6 protein was not obviously produced by blood DC,388 viral exposure induces IL-6 secretion.395 High numbers of DC used as APC are associated with IL-6 production and allogeneic cytotoxic T-lymphocyte responses are blocked by IL-6 antibody.396 

IL-7

IL-7 is a pleiotropic cytokine397 that has an important role in the differentiation of B lymphocytes, the maturation of thymocytes,398,399 and as a growth factor for mature T lymphocytes.400,401 Highly purified CD4+ as well as CD8+ T lymphocytes proliferate in response to IL-7 and a comitogen.402 The expression of IL-7 in human CD83+ blood DC was reported to be variable at the mRNA level.389 Sorg et al403 have shown that activated CMRF-44+ DC but not freshly isolated DC express IL-7 mRNA. Moreover, purified CD14+ Mo failed to express IL-7. An IL-7–neutralizing MoAb was shown to inhibit DC stimulation of an allo-MLR.

IL-10

Intriguing RT-PCR data suggest that DC may produce IL-10, which may further regulate T-lymphocyte responses.389 

IL-12

IL-12 consists of two chains, a p40 protein and a p35 protein, which are both required for activity. Mouse DC coordinate secretion of the p40 and p35 components (p40 alone may inhibit) to produce functional IL-12339,404 and IL-12 production by human DC has also been described.405 Microparticle adsorbed versus soluble Ag may activate IL-12 mRNA synthesis.315 At least in vitro, the production of IL-12 by DC may drive the bias toward a TH1 response so often observed experimentally. Intriguing data suggest that human TH2 clones can induce DC IL-12 and that their cytokine profile was influenced by the APC.406 

IL-15

IL-15 has no homology with IL-2, but the two cytokines share a number of biological activities, presumably as a result of their shared receptor γ chain.407 IL-15 is produced by several cell types, whereas IL-2 is derived primarily from T lymphocytes. This, and the wide distribution of the IL-15Rα chain, suggests IL-15 has a different physiological role to IL-2. IL-15 is produced by Mo, MØ,408 and Mo-DC. Blood-derived DC have been shown to induce IL-15 in RNA and protein.409 

IFN

No IFN-α or -γ is produced by unstimulated DC,16,389 whereas Mo produce large amounts of both. HIV-1132 and HSV395 infection induced high levels of DC IFN-α production.

Other Cytokines

No biologically active IL-2 is produced by DC.16 Low-level IL-3 transcripts have been found by some investigators389 and not others (unpublished). Neither IL-4 nor IL-5 mRNA was detected by RT-PCR in CD83+ cells, a feature which distinguishes DC from Mo.389 Equivocal IL-11 mRNA levels were reported in CD83 + blood DC.389 IL-13 was not found by RT-PCR.389 

The RT-PCR analysis by Zhou and Tedder389 described mRNA for TGFβ1, a result which should stimulate functional studies. TGFβ2 was not been detected.389 

Chemokines

DC may well produce several of these low molecular weight compounds, which facilitate the migration of leukocytes. MIP1α and MIP1β mRNA have been reported in CD83+ blood DC.389 Their production by LC has been reported391 and MIP1α is produced by CD34+-generated DC.366 At least some mRNA for RANTES and MCP-1 may also be present in blood DC.389 

Recent RT-PCR analysis of CD83+ DC indicated that they produce significant amounts of IL-8 mRNA,389 but IL-8 secretion has yet to be established. DC-like cells generated from CD34+ progenitors secrete IL-8 after CD40 ligation.366 

Novel chemokines produced by DC are being investigated.

Cytokine Receptors

DC growth, differentiation, migration, and functional Ag processing and presentation are all regulated by cytokines. The expression of cytokine receptors on DC is likely to determine many of these events16,410 and again the data suggest their density varies according to the state of DC differentiation/activation.411 

IL-1R

The type I and the type II IL-1R both bind IL-1α and β as well as the IL-1R antagonist, but the two receptors use different signal transduction pathways. Direct binding of IL-1 to mouse LC (∼500 sites/cell) with nearly an order of magnitude less binding to spleen DC has been shown.410 Purified human blood and tonsil DC express mRNA for type I but not type II IL-1R.16 Direct labeling of IL-1R with Cdw121α and β MoAb has recently confirmed their expression on human LC.411 IL-1α and Il-1β induce CD40 on human blood DC.26 

TNFR

The type I TNFR (55 kD) and the type II TNFR (75 kD) differ in their intracellular portions, suggesting different signaling pathways. Studies on human blood DC have confirmed mRNA and surface expression of both types.392 Ryffel et al412 demonstrated type II TNFR on thymic DC using histological techniques.

Functionally, it is clear that TNF-α influences DC growth and differentiation,209 prolongs the viability of isolated DC,413 induces CD40,26 and probably upregulates APC function.392 Furthermore, TNF has been implicated as providing a stimulus for LC migration in mice230 255 and as this effect was achieved using human TNF-α, it suggests a type II receptor interaction.

GM-CSF Receptor (GM-CSFR)

The presence of GM-CSFR on DC was anticipated given the effects of this cytokine on DC phenotype, migration and growth (In Vitro Cultivation of DC and DC Lines).17,38 Mouse DC have been shown to have GM-CSFR by direct binding studies (∼3,000 sites/cell)410 and RT-PCR analysis has shown GM-CSFRα mRNA in human blood and tonsil DC.16 Direct staining of blood DC with CDw116 MoAb also identified GM-CSFR.389 The majority of LC express the GM-CSFRα chain but only 15% of CD1a cells expressed the GM-CSFRβ chain, although the latter upregulated in culture.411 

IL-2R

The IL-2Rα (CD25) was not detected on human tonsil DC46 but has been detected on blood DC,51 although the IL-2Rβ chain was not detected.389 Again these differences may reflect the stage of cell differentiation. IL-2R were induced by culture of murine LC414 and by GM-CSF on rat lymphatic DC.81 IL-2Rα was induced on cultured DC by CD40L.366 Despite these findings no functional effects of IL-2 on DC have yet been described.

TGF-R

Endoglin (CD105) is present on DC389 and activated Mo and acts to bind TGFβ1 and TGFβ3 .151 These cytokines may well help direct DC motility.151 

Other Cytokine Receptors

DC do not express the M-CSFR410,413 or respond to M-CSF or G-CSF,413 although a neonatal-skin–derived LC line may do so.415 An RT-PCR analysis found no evidence of IL-4R on human DC,389 which is surprising given the effects of IL-4 on DC precursor growth and differentiation. IL-6R and gp130 were detected on a subset of human LC.411 IFNR have been described on a subset of LC.411 A number of studies suggest DC respond to IFN-γ by inducing MHC class II Ags, ICAM-1,416 CMRF-44,22 CD80, and CD86.26 

Chemokine Receptors

The C-C chemokines MCP-3, MIP1α, and RANTES elicit chemotactic migration and an increase in intracellular Ca2+ in DC.228 DC did not respond to C-X-C chemokines IL-8, IP-10, or the C-C chemokines MCP-1 and MCP-2.228 

T-Lymphocyte Responses

The circumstantial evidence suggests DC are capable of costimulating a wide range of T-lymphocyte responses. Antigen type, route of administration, the overall cytokine milieu, and the genetic background (this has a profound influence on TH1v TH2 responses) will influence that response.

The Ag source and the relevant peptide binding to DC MHC class I or II molecules selects CD4+ or CD8+ and T-lymphocyte responders. A second contribution may come from the DC itself, via the type and balance of costimulator molecules expressed,417 eg, CD80 v CD86418 or DC secreted cytokines (IL-7, IL-12, etc), which drive either a CD4 (TH1 /TH2 ) or CD8 (TC1 /TC2 ) lymphocyte response.406 419 Thirdly, the external environment, ie, other cell types and the cytokine milieu, may influence the resulting phenotype of a DC-stimulated T-lymphocyte response, either directly or indirectly via the plethora of DC-surface receptors. Local microenvironmental influences appropriate to gut, respiratory tract, or skin may well differ.

In vitro data emphasize that DC drive strong CD4+, predominantly TH1 , responses. This may reflect in vitro bias, particularly of the MLR and perhaps other test Ag systems. The use of DC exposed to PPD to immunize mice results in a predominant TH1 response, ie, IL-2 and IFN-γ without IL-4, when the T lymphocytes primed in vivo are retested in vitro.284,285 An interesting difference was noted when the 19-kD immunodominant protein of M tuberculosis was used as an alternative: the T-lymphocyte response induced limited IL-2 but no IFN-γ production.285 Tetanus toxoid induces a TH1 response in vitro,149 yet this is an Ag that in vivo produces a strong antibody response and IL-4 secretion, ie, a TH2 response from previously sensitized lymphocytes. The production of IL-12 by DC339,404 or by other cells, eg, MØ, may be an important factor in this bias toward DC-directed TH1 responses. Other cytokines apart from IL-4, eg, IL-1, may be produced by DC and direct TH2 responses. Human DC are clearly capable of stimulating both TH1 and TH2 clones in vitro420 and the ultimate response may well reflect a balance between T-lymphocyte IL-4 and DC IL-12 production.406 

The strong bias toward TH1 responses is less evident in vivo and the route of DC Ag administration is relevant. The IV administration of Ag-pulsed DC induced an antibody response (TH2 -like), whereas SC administration induced a DTH (TH1 ) response.285,421 This may relate to the production of IL-10 from the spleen.422 Both IL-4 and IFN-γ, indicative of a TH0 response, were produced when mice were injected IV with DC pulsed with KLH.423 However, myoglobulin or human gammaglobulin pulsed DC administered IV stimulated IgG2a production, the antibody isotype associated with the TH1 phenotype.291 Recent in vivo data obtained using DC pulsed with tumor Ags suggest that these DC preparations result in strong cellular (including cytotoxic responses) as well as humoral responses (see later). It has been suggested that mucosal-derived DC generate a preferential TH2 response.424 

Production of DC in vitro has provided cells for basic scientific studies and for therapy. The divergent approaches adopted for growing DC and the problems inherent in identifying DC require some scientific caution.

Normal DC and DC-Like Cells

Several cellular sources have been used to generate DC.

BM

GM-CSF–generated mouse BM cells resemble DC morphologically and have most but not all the phenotypic characteristics of DC.20,21,284 The ability of these cells when injected in vivo to migrate and function as DC is most encouraging.82 Several groups have added IL-4 to GM-CSF in order to produce mouse DC,202,203,425-428 but this may enhance the nonspecific stimulation of T lymphocytes.425 In vitro flt-3 ligand contributes to mouse lymphoid (and it is anticipated myeloid) DC production.304 

The allostimulatory activity of human BM84 makes the functional assessment of the progeny of cultured human BM using the MLR problematic. It appears that committed CD14+ Mo/MØ within human BM are not good primary APC.44,84 It is a significant question as to whether the in vitro conditions, including cytokines, induce such activity in differentiating Mo/MØ. Indeed, there is evidence that CD14+ cells differentiated in vitro from BM precursors acquire APC activity not seen in fresh human BM.84 213 

Young et al127,212,213 have suggested that SCF, GM-CSF, and TNF-α provided optimal conditions to allow at least a significant percentage of DC to grow in liquid culture. Assessment of the cells produced suggests approximately 10% to 15% are differentiated human DC. Overall yields are of the order of 106 cells from 104 CD34+ BM cells after 3 to 4 weeks. Flt3 ligand may increase the number of DC from human BM progenitors.217 The presence of MØ may suppress DC development in culture.

Cord Blood CD34+-Selected Blood Progenitors

Caux et al101 described the culture of CD34+ cord blood cells in GM-CSF and TNF-α to produce a population of CD1a cells (5% to 15% of the progeny) with marked allostimulatory activity. Yields were approximately 107 CD1a cells from 106 CD34+ cells at 2 weeks. Santiago-Schwarz et al429 have described similar results. It appears that IL-3 can substitute for GM-CSF in this system.163 A further report366 explored the use of additional CD40L stimulation and provided a more detailed phenotype, indicating some heterogeneity in the resulting cell population. This complex cell population, which includes DC, can process soluble Ag and prime naive T lymphocytes.430 Other groups have followed suit using similar or modified conditions.218,431-433 The inclusion of SCF for the first 5 days increased cell yields approximately fivefold (day 12 yields were approximately 10-fold starting CD34+ cells).433 The resulting populations were heterogenous but late addition of IL-4 (day 12) induced CD1a expression and increased the allostimulatory activity of the cells. Strobl et al434 used SCF, GM-CSF, TNF-α, and TGFβ1 to generate CD1a+ progeny from CD34+ cord blood cells and showed that TGFβ1 would substitute for serum in this system.

Mobilized Peripheral Blood CD34+ Progenitors

Mobilized CD34+ cells have been used as another source of progenitors for attempts to grow DC. Bernhard et al431 used GM-CSF to culture CD34+ progenitors producing 30% to 60% CD1a+ cells within a 20- to 40-fold expanded total cell population at day 15. Mackensen et al432 added GM-CSF and IL-4 to a cocktail of SCF, EPO, IL-1β, IL-3, and IL-6, thereby generating a high proportion (45%) of CD1a+ cells. Siena et al218 used SCF and flt-3 ligand to supplement GM-CSF and TNF-α culture of CD34+ mobilized cells. Yields were of a similar order (4 × 107 cells from 106 CD34 cells) with 33% to 55% CD1a+ cells in the progeny. A recent study210 raised the possibility that CD34+ cells showed commitment to epithelial-based CD1a+ or CD1a nonepithelial DC populations: expression of the skin homing receptor defined a CD1a+ CLA+ LC precursor.

Blood DC and Their Precursors

Markowicz and Engleman17 established that human blood DC (and their precursors) obtained by Percoll gradient separation and phenotypic selection could be maintained in GM-CSF for 4 to 6 weeks. The cells formed clumps in fluid phase culture and after several days developed allostimulatory cells with a DC morphology.

PBMC (blood Mo)

Sallusto and Lanzavecchia149 cultured adherent human PBMC in GM-CSF and IL-4 or GM-CSF and TNF-α and found potent APC populations emerged. Elegant studies established that the GM-CSF and IL-4–generated APC could process and present soluble Ag to tetanus toxoid-specific clones. Conversely, GM-CSF– and TNF-α–generated APC were less effective at processing specific Ag but had equivalent allostimulatory capacity. Romani et al435 also used GM-CSF and IL-4 to generate DC-like progeny from adherent PBMC. The resulting cells contained a significant proportion of DC-like cells (3 to 8 × 106 cells from 40 mL blood at 5 to 7 days). The experiments did not clarify from which component of PBMC the APC were derived.

Recent experiments suggest that peripheral blood Mo prepared by a variety of methods may differentiate into potent allostimulatory cells.75,223,347,436,437 Particular cytokine combinations, most including IL-4, which suppresses Mo development220 and downregulates CD14 expression,125 may drive Mo to DC differentiation. IL-10438 and TNF-α149 may control the balance between Ag uptake capacity and the upregulation of costimulator activity. Although these Mo-DC have functional attributes not unlike DC, they require testing in Ag-specific assays.

Induction of CD83 has been reported on Mo cultured in similar conditions.225 The CMRF-44 MoAb has also proved useful for analyzing the progeny of cytokine cultured PBMC (Fig 1D). Optimal concentrations of initial GM-CSF, IL-4, and subsequently TNF were evolved and the activity of the CMRF-44 phenotype Mo DC were shown to be superior Ag specific APC (in preparation).

A concern in many of these experiments is that the starting PBMC preparations include both Mo and DC or their precursors and the progeny may be heterogeneous. Two significant phenotypic differences apply to in vitro cytokine-generated DC. Cord blood–derived CD1a+ cells are CD16+366 and CD11b is strongly expressed on CD1a+ cells generated from mobilized CD34+ cells.432 These observations, plus the paucity of CMRF-44 and CD83 staining in certain circumstances (in preparation), suggests that many of these cells retain certain Mo/MØ features.

LC and Tissue DC

The technical requirements for their preparation and the low yield make it unlikely that epidermal LC will be a future source of DC for therapy. Although dermal DC migrate readily from skin explants, this source is unlikely for practical reasons.

To summarize, a unifying approach might use a cytokine mixture to encourage initial hematopoietic precursor proliferation before adding a mixture of DC differentiating agents 4 days later. Antigen loading might then be initiated and further factors used to optimize DC costimulator activity before therapeutic administration. The resulting cells are heterogeneous, both morphologically and phenotypically. Nonetheless, they include significant numbers of functional APC, some of which are likely to be DC.431 Studies to verify their capacity to migrate are needed.

DC-Derived (transformed) Cell Lines

The availability of well-characterized DC lines is essential to enhance progress.

Transformation of Normal DC

Some putative mouse DC lines have been described,439,440 but their DC origin has not been widely accepted, nor has further data with these accrued. Significant progress has been made since using retroviral transformation of murine DC preparations. O'Neill and Ni441 transformed mouse spleen cells and obtained cell lines, which had several characteristics of DC. Transformation of murine progenitor MØ and LC has generated cell lines with antigen-presenting characteristics of DC.442 Superinfection of GM-CSF–transduced mouse BM cells with c-myc and c-raf oncogenes produced 33D1 and DEC-205+ DC clones, which have been cultured for 6 months.443 

LC-like lines have been generated from neonatal mouse skin. These have a block in differentiation preventing MHC class II molecule upregulation and significant costimulation,415 which could be modulated by changing the cytokine environment. Attempts to generate LC lines in humans have also focused on neonatal skin.

DC Malignancies

It would be most curious if DC did not undergo spontaneous malignant change.

Some attempt has been made to ascribe a DC phenotype to a few examples of acute myeloid leukemia444 445 but this is difficult, whilst the phenotype and antigen-presenting activity of cytokine-treated Mo/MØ remains unclear. Even though the French-American-British (FAB) M4/M5 and possibly FAB MO AML subgroups may include some leukemias of DC origin, their definitive recognition will depend on the use of the new DC markers.

The bcr-abl translocation is found in blood DC of CML patients (in preparation) and in Mo-DC derived from CML patients.446 These cells have normal APC function.

Histiocytosis is a rare condition predominantly affecting children.447 Class I histiocytosis, defined by CD1 staining, is probably a malignancy of LC origin and a case labeled with both the CD83 and CMRF-44 MoAb further emphasized the association (unpublished). Some lines derived from Histiocytosis X material are undergoing further characterization. An association between LC histiocytosis, Hodgkin's disease (HD), and non-Hodgkin's lymphoma is well documented.448 The related “indeterminate cell proliferations” present with either solitary or multiple cutaneous nodules. Although morphologically and phenotypically similar to LC histiocytosis, they lack the characteristic Birbeck granules but are S100+ and CD1+.449 

IDC tumors have also been described but are extremely rare with less than 20 reported cases.450,451 These resemble proliferations of LC but lack both Birbeck granules and CDI.450 

The malignant cells in HD, RSC/HC, have several intriguing similarities to DC, despite the fact that they do not resemble DC morphologically. The possibility that these cells might be derived from IDC was raised some time ago and reinforced recently by phenotypic comparisons between RSC/HC and human DC.452,453 The origins of RSC/HC remains uncertain despite evidence that a proportion of selected cases have a clear B-lymphoid origin.454 However, quite a strong argument can be made that at least a proportion of RSC/HC are DC related, which would fit with most biological features of this disease. Recent analyses suggesting that the CD83454 and p55191 Ag are expressed on most HC adds weight to the hypothesis. HC also have potent APC activity, as do certain HD cell lines.455 Although the origin of some of the latter are hotly disputed, one cell line, L428, has been shown to have a unique molecular constitution,456 potent allostimulatory activity,455 and a cell-surface phenotype which includes the CMRF-44, CD83, CD80/CD86, and CD40 Ags.25,169 454-456 

The difficulty in identifying DC has been limiting but the CD83, CMRF-44, and CMRF-56 reagents to track DC have encouraged new clinical studies.

Infectious Disease

HIV

Patients infected with HIV lose the ability to mount primary immune responses and this may, in part, be due to reduced DC function.457,458 MØ secondary antigen presenting function is unaffected in HIV+ subjects. An original report documenting HIV within LC by in situ hybridization has been confirmed by observing direct LC infection using MØ-trophic HIV strains.459 Approximately 1% of LC in acquired immunodeficiency syndrome (AIDS) patients are likely to be infected.460 

Some workers originally maintained that DC in vitro did not support productive human immunodeficiency virus (HIV) infection and were not a reservoir of infection.461,462 However, careful analyses showed that rapid replication of the virus occurred in a T-lymphocyte syncytium surrounding the DC.463 Others have produced data to suggest DC (identifying the cells becomes an issue) are infected directly.457,464 DC appear to host LTR retroviral transcriptional control elements310 capable of sustaining HIV replication. DC lack the Sp-1 transcription factor essential for HIV replication; however, this is provided in the DC-T lymphocyte syncytium.195 DC can bind HIV and infect CD4+ T lymphocytes while activating them,465 emphasizing that the interaction between these two cell types is critical.

The reduced APC function may result from reduced T-lymphocyte feedback (see earlier) as DC in HIV-infected patients present satisfactorily to normal syngeneic T lymphocytes.460 An influence of HIV on DC migration has also been postulated.460 DC can generate cytotoxic T-lymphocyte responses to HIV Ags123,297,298 and when infected produce IFN-α.132 

Other Clinical Viral Infections

There is good evidence that viral Ags for influenza276 and Herpes simplex286 are presented by DC to generate T-lymphocyte responses.

Mycobacterial Infection

Human DC present mycobacterial Ags for primary T-lymphocyte responses.77,292-295 Mycobacteria have been located within the cutaneous infiltrates of leprosy patients466 and DC constitute a minor population of the inflammatory infiltrate.

Other Bacteria, Protozoal, and Fungal Infections

Clearly DC present a range of Ags from different organisms, eg, Chlamydia296 Leishmania,48,317,318 Candida52 in vitro to T lymphocytes and for in vivo responses. Although major changes in DC numbers, localization, and function might be expected in these infections, the inability to monitor DC readily has prevented further clinical observations. However, a recent study467 showed that dermal DC but not LC engulf Borrelia burgdorferi (Bb), the organism associated with Lyme disease, and present Bb Ags.

There are no data as yet relating to DC numbers, viability, trafficking, etc in different clinical infections.

Immunodeficiency

There have been several reports of deficient APC function in primary immunodeficiencies, including hyper IgM syndrome.468 No DC specific defect has emerged as yet but deficiencies in MHC class II Ag expression and CD40L expression (hyper IgM syndrome) are likely to profoundly affect DC function.

Autoimmune Disease

It is no surprise that DC are also the most efficient APC for endogenous self Ags.338 Certainly, DC are capable of generating strong autologous responses in vitro — this is generally assumed to reflect exogenous Ags but may reflect activation of truly autoreactive clones — the specificities of those autoreactive clones have not been tested. Known autoantigens, eg, myelin basic protein or thyroid extract, may be incubated with DC and these will initiate autoimmune disease in mice.469 Experimental models of diabetes in mice suggest that APC may trigger autoimmunity by initiating a local or systemic response to an infective agent.470 Enhanced cluster formation by DC from nonobese diabetic (NOD) mice has been reported471 and DC incriminated in the islet pathology, which evolves in BB rats.472 In contrast, transfer of DC from pancreatic LN but not DC from other sources inhibited the development of autoimmune disease in NOD mice,473 perhaps by reinforcing tolerance.

Kabel et al474 have studied DC in autoimmune disease and demonstrated histological changes in the tissues, which suggest DC may contribute to the presentation of autoantigens. Some evidence that autoantigens on DC may switch off immune responses has come from human studies showing that thyroglobulin exposed DC reduce antithyroglobulin antibody production in vitro.475 

DC may also prolong or exacerbate local immune-based inflammatory reactions such as occur in arthritis. Increased numbers of DC-like cells have been described in chronic arthritic synovial fluid296,476-480 and in adjacent synovial membranes.481,482 Most work relates to rheumatoid arthritis but similar changes are also seen in chronic arthritis of other etiologies. Recent data suggest the DC in chronic arthritis do not have a fully activated phenotype and may express relatively low levels of the CD83/CMRF-44 DC differentiation/activation Ags (in preparation). Even more intriguing is the observation that they do not, despite originating from an inflammatory environment, have high levels of the CD80/86 costimulator molecules.479,480 This does not appear to be due to the IL-10 in synovial fluid and may be an indirect effect on DC as a result of T-lymphocyte feedback triggered by exogenous TGF-β.483 

The concept that vaccination with in vitro DC stimulated autoreactive clones may generate yet further “anti-idiotypic” control of an autoimmune response is a challenging prospect for therapy. So, too, are thoughts of fixing DC in a “tolerizing state” for reinforcing self tolerance.

Allergy and Hypersensitivity

LC are intimately involved in contact allergen sensitization261 and appear to have an affinity for certain metals and haptens. Some blood DC and LC express IgE receptors (see Cell Surface and Other DC-Associated Molecules) and may be capable of initiating the late phases of type I hypersensitivity. They may also be involved in the immunological hyper-reactivity of psoriasis364,484 and atopic dermatitis.466 DC appear in the early stages of the inflammatory infiltrate and appear to be located in dermal clusters in psoriasis but appear in diffuse infiltrates in atopic dermatitis.466 

The probability that DC within the respiratory tract contribute to inflammatory and immunological events, triggering airway hypersensitivity and allergen-induced asthma, is being investigated. Holt et al245,485 have shown that DC have a substantial turnover in the respiratory tract, influenced markedly by inflammatory mediators. Interaction with aerosolized Ag leads to DC migration, Ag priming of T-lymphocyte responses, and in turn a sensitized T-lymphocyte infiltrate.485 This generates further inflammatory mediators. An inhibitory effect of lung MØ (mediated by nitric oxide) on APC function251 is noted in these systems.

DC in Malignancy

The biology of DC in malignant disease is currently arousing great interest. Animal data from Muller et al486 suggests that skin carcinogenesis is associated with a paucity of DC. There is a large body of literature alluding to the effects of UV radiation on murine DC.487 This appears to reduce DC antigen-presenting function, probably via a direct effect on protein synthesis. DC are associated with tumors in mice; however, further mouse data have called into question whether DC recruited into skin tumors indicate a host immune response to tumor.488 

Human LC are likewise sensitive to UV effects.489 UV light decreases numbers in the skin490 and affects DC function,357 blocking upregulation of costimulator molecules. Various correlations have been proposed in human skin cancers. LC were reported to be increased around in situ and early invasive melanomas.491 With progression, LC numbers near deeply invasive melanomas decrease.491 In other skin tumors, increased LC infiltration appears to be associated with increasing tumor differentiation.492 Basal cell carcinoma associated DC do not show increased CD80/86 expression and in one case appeared to be poor APC compared with blood DC.493 Increased LC numbers in the cutaneous T-cell lymphomas suggested improved survival.494 Reduced LC numbers (perhaps via an indirect mechanism) were reported in the skin over breast cancer compared with distant sites.495 

There is a steadily increasing body of literature correlating DC or LC numbers (detected by S100 staining) in or adjacent to tumors with their prognosis. Increased numbers are associated with better outcomes in cervical carcinoma, colorectal adenocarcinoma,496 gastric carcinoma,497 nasopharyngeal carcinoma,498 lung adenocarcinoma,499 thyroid,500 and prostate cancer.501 

Whether the DC observed within malignant tissue take up tumor associated Ags (TAA) and migrate centrally to initiate effective T lymphocyte responses is unclear. It appears that splenic APC from tumor-bearing mice retain TAA and are capable of immunizing against fibrosarcomas in mice. However, they only acquire sufficient Ag or costimulator capacity late in the disease course when the host is already overwhelmed.502 Gabrilovich et al503 showed that spleen DC function in tumor-bearing mice was reduced. Defective DC function assessed by MLR and reduced ability to generate p53 specific CTL was noted. This was associated with reduced MHC Ag expression on DC. Tumor cell supernatants were shown to suppress the ability to generate functional DC from BM but not to affect the function of already mature control DC.427 Recent data suggest that vascular endothelial growth factor (VEGF ) produced by human carcinoma lines can block DC hematopoiesis in vitro.504 

Tumors might frustrate normal DC physiology as yet another mechanism of avoiding immunosurveillance. Studies show that few DC are recruited into the cellular infiltrate of renal cell carcinoma (RCC) and only a small proportion of these have the phenotype (CD83+ CMRF-44+) of mature activated DC. Those that do are located near lymphoid aggregates.190 Significantly, the few CMRF-44+ DC isolated from RCC had APC activity whilst the CMRF-44 DC did not stimulate in an allo-MLR. This, given the MLR-stimulating activity of similar isolated LC and dermal DC, argues in favor of RCC suppression of DC function. Additional data238 indicate that few DC are recruited into prostate carcinomas and these too lack an activated phenotype. Further data are accruing to suggest this phenomenon also occurs in breast and bowel carcinoma.

DC may sensitize for allogeneic responses either (1) directly, by presenting preformed MHC ± minor Ags, or (2) indirectly by processing and presenting MHC alloantigens presumably as peptides.14 UV irradiation of blood products abrogates MHC allosensitization, presumably through an effect on DC.505 Cyclosporin A may affect DC activity506 and unstimulated interstitial DC are steroid sensitive.14 

Solid-Organ Transplantation

Soon after interstitial DC were identified it was suggested that these might be the passenger leukocytes which contributed to the strong primary allograft reaction.6,9,11,12 Formal proof that DC in heart and kidney allografts contributed to the immunogenicity of a graft was then obtained.13,14 Treatment of mouse pancreatic islet507 and thyroid508 grafts with antibodies binding DC and C before transplantation prolonged their survival. Despite other expectations, allografted thymic DC appear to be immunogenic in mice.509 The situation with liver allografts in these models is less clear, because this organ appears to have special properties.173 Depletion of donor kidney DC in clinical practice appears to have some clinically beneficial effect510 but maximum efficacy will require very effective DC depletion.

Recipient T lymphocytes cluster with DC in rat allografts, implying direct sensitization of the host's immune response. Transplant studies in mice suggested that donor DC did not exit as expected into lymphatics but some at least migrated via the bloodstream to the spleen.262 This is more consistent with central sensitization rather than intragraft lymphocyte activation. Graft alloantigens are undoubtedly also processed by recipient DC and indirect responses contribute to ongoing anti-allograft responses. Recipient DC entering and trafficking through the donor organ may contribute to this process, as may antigenic material carried directly via blood and lymph to central LN.

The interstitial DC within solid organs are not fully activated/differentiated as regards APC function.24,173 Therefore, it is theoretically possible that these cells mediate a tolerogenic effect in certain circumstances. DC that are treated with UV or heat511 to destroy their Ag-presenting potential, yet retain alloantigenic activity, do have an immunosuppressive effect. Mouse BM-derived DC progenitors apparently prolong the survival of heart allografts.512 Recent data imply that donor leukocytes emigrate from long-term kidney and liver donors and provide low-level chimerism of the recipient.308 Whether these cells represent donor-organ DC or the end progeny of donor hematopoietic progenitors is unknown.49 Therefore, it is not possible to distinguish whether this phenomenon of low-level donor chimerism has an immunosuppressive effect akin to other chimeric/tolerance models229 or whether it is a unique effect attributable to the donor-derived DC. It is possible that in certain circumstances donor tissue–derived DC never achieve normal costimulatory activity, particularly if T-lymphocyte responsiveness is compromised by immunosuppression.

Blood and BM Transplantation

There is little direct experimental data. CD34+ cells stimulate an allo-MLR84 and human BM contains DC-like cells.44 Therefore, direct priming by graft APC of residual host T-lymphocyte immune response, which survives the myeloablative/immunosuppressive conditioning before stem cell transplantation, may contribute to stem cell graft failure/rejection. Fractionation of CD34+ cells to avoid allostimulatory activity, while retaining stem cell reconstituting activity (CD34+ HLA-DR), might be considered to avoid this.

Animal experiments suggest significant myeloablative treatment removes most, but not all, recipient DC.7 It is relevant that DC inhabit all the sites commonly associated with acute graft-versus-host disease (GVHD). These DC may contribute to direct donor T-lymphocyte allosensitization to minor Ags and prime for acute GVHD as predicted by recent in vitro studies.513 Their function514 in clinical BMT can now be examined to identify DC and define their activation/costimulation status. Clustering of donor T lymphocytes around recipient residual DC might be observed as part of the GVH reaction — activated DC-T lymphocyte association is rare in normal skin (Fig 1I).

The indirect presentation of recipient minor histocompatibility Ag via donor-derived APC may be the main route of GVHD sensitization. The cytokine storm after conditioning515 would be predicted to maximize DC migration and Ag-presenting activity. The reconstitution of DC after autologous stem cell grafting appears to be relatively prompt in animals,7 but it has yet to be studied in human autografts. The reconstitution of LC in humans after HLA-matched sibling BMT takes several weeks, but by 1 year the skin DC are all of donor type.514 Now that methods are available, we are studying DC reconstitution in the blood of HLA-matched sibling allografts.

Control of DC Ag presentation during allografting procedures has immense therapeutic potential as witnessed by the suppressive effects of CTLA-4-Ig and CD40-Ig on organ graft rejection and GVHD.516 

Clinical hematologists cannot fail to be intrigued by current investigations using DC as cellular therapy. As “nature's adjuvant,” DC may be the ideal vehicle for immunization against infectious agents, malignant tissue, and even autoreactive lymphocytes. Investigators are constructing cellular/molecular vaccines, expressing GM-CSF, costimulator molecules, and TAA for tumor therapy. Some of these approaches, eg, GM-CSF, may recruit DC into tumors517 and others, eg, naked DNA vaccination, may be ways of delivering Ag into host DC.518 Well-defined autologous, Ag-loaded, DC preparations may prove to be the fastest and safest means of vaccinating patients. Clearly there is much to be learned: the type of DC preparation (particularly as lymphoid DC may tolerize) and how to expose them to Ag in vitro (or in vivo) requires considerable experimentation. It is salutary to realize the dramatic differences that DC activation/differentiation state, Ag preparation (protein or peptide) Ag modification, dose of DC and Ag, route of administration, host milieu, host genotype, etc may have on outcome. Furthermore, the possibility that myeloid DC may in certain circumstances suppress475,512 519 rather than enhance immune responses should be remembered.

Infectious Agents

DC have been used in vitro to present bacterial and viral Ags to T lymphocytes. Herpes simplex peptide derivatized with a triacyl tail and delivered to DC by liposomes was more effective than virus alone.286 Strong influenza virus–specific cytotoxic responses can be induced in vitro by using virus-infected DC,520 but depletion of Mo was necessary to remove an inhibitory effect. Peptide rather than whole virus may be better for DC-generated influenza responses.290 DC-generated responses to HIV proteins have been obtained in vitro.123,297,298 521 

An impression has emerged that vaccination with Ag pulsed DC in animals is as efficacious, if not considerably more so, than conventional adjuvant vaccines.15 290 DC are likely to improve responses to bacterial or viral vaccines, purified proteins or peptides, but no formal clinical comparison has been undertaken yet. A large number of variables need to be addressed before an optimal DC-based vaccination program might be contemplated (eg, with HIV-derived peptides).

Malignant Disease

The immune response is capable of focusing on TAA which include (1) viral Ags in malignancies with a viral etiology; (2) tumor-specific Ags such as recombined or mutated oncoproteins and abnormally glycosylated molecules, or potentially (3) tumor-lineage–specific autoantigens. Responses may be directed against the malignant cells themselves or perhaps TAA presented by the stroma.

Viral Ags

Donor-derived cytotoxic T lymphocytes are active against post-BMT EBV+ lymphomas. Murine responses to HPPV have been produced in vitro and in vivo.

Tumor-Specific Ags

Some oncogenic proteins, which are unlikely to be subject to downregulation as a result of immuno-selective pressure, make ideal TAA. For example the bcr-abl fusion protein resulting from the 9; 22 translocation in CML and the PML-RARα fusion protein resulting from the 15; 17 translocation in promyelocytic leukemia, are potential leukemia-specific Ags. Likewise, mutations in oncogenes, eg, ras or tumor-suppressor genes, eg, p53 produce potential TAA, albeit with only 1 to 2 amino acid differences from the normal protein. The Ig idiotypes produced by non-Hodgkin lymphoma (NHL) cells are specific tumor markers, as are the Ig paraproteins produced in multiple myeloma. In some instances, the oncogenic changes may create a TAA on a malignant cell by indirect means. Thus, the mucin molecule, MUC-1, is abnormally glycosylated in breast, pancreatic, ovarian, and prostatic cancers as well as multiple myeloma, thereby exposing a repeated antigenic protein sequence.

Tissue-Specific Ags

Several Ags are sufficiently tissue restricted to allow an autoimmune response to be efficacious, eg, MAGE, tyrosinase in melanomas, prostate-specific Ag, or prostate-specific membrane Ag (PSMA) in prostate cancer. The CD33 Ag might also be a sufficiently highly restricted myeloid lineage autoantigen for limited (possibly pretransplant) autoreactive therapy. Finally, overexpression of p53 wild-type protein in tumors may be another target autoantigen.

DC Priming With TAA In Vitro

Studies showed that both mouse spleen DC and LC could present tumor Ag and generate specific T-lymphocyte responses.522 

It was by no means certain that the human TCR repertoire would contain receptors for TAA peptide-MHC complexes, but the early data are encouraging. Thus, bcr-abl specific T lymphocytes respond specifically to CML cells and DC incubated with bcr-abl peptides.341 DC also generate primary responses to bcr-abl peptides in vitro. Mo-DC obtained from CML patients generated CML-specific cytotoxic activity.446 Tyrosinase-transfected DC generated specific T-lymphocyte responses523 and DC present PMSA with some specificity in vitro.524 Finally, in vitro sensitization to the HER neu 2 breast cancer Ag has been reported using CD34+ generated DC.431 

DC Vaccination With TAA In Vivo

The T-lymphocyte responses described to date against known TAA and a host of as yet undefined TAA are often weak. Further, in any one individual with cancer, the host immune response to TAA has, by definition, failed. Therefore, reliable means of generating significant T-lymphocyte responses to TAA must be evolved, hence the interest in clinical DC immunotherapy. The data,190 suggesting that tumors circumvent DC in vivo, also argue in favor of this approach.

In vivo data to suggest DC are potent immunotherapeutic agents when pulsed with TAA are accumulating rapidly (Table 2). Early experiments suggested spleen APC (probably DC) from tumor-bearing mice and normal APC were capable of inducing immunity in normal mice to a subsequent tumor challenge.502 Skin-derived APC exposed to tumor lysate were also effective at vaccination against tumor challenge.525 Fascinating data obtained in a rat sarcoma model suggested mesenteric-derived lymphatic DC exposed in vivo or in vitro to tumor could protect against tumor inoculation.263 A novel strategy using DC fused to tumor cells appeared to protect against subsequent tumor inoculation in mice.526 

Table 2.

In Vivo Immune Responses to Tumors Induced by DC-Induced Anti-tumor Immune Responses In Vivo

SpeciesTumorDC PreparationTumor AGVaccinationOutcomeAuthor (reference)
(Pre/post)
Mouse MCA-induced sarcoma (McSa-1) Spleen DC Sarcoma cells 18 Days post IV Some dose-dependent prolonged survival Knight et al 1985519 
    0 Day IV All died 
Mouse Fibrosarcoma (CSAIM) T-depleted spleen cells Tumor membranes 5 Weekly pre SC Reduced tumor growth Shimizu et al 1991502 
Mouse MCA-induced fibrosarcoma (S1509a) Unseparated epithelial cells (%LC) Tumor lysate 7 Days pre SC Reduced tumor growth Grabbe et al 1991525 
Mouse Lymphoma (BCL1) Splenic low density (BSA) IgM idiotype 10 Days pre IV 23/28 Treated alive Flamand et al 1994535 
    5 Days Ig IV 2/20 Untreated alive 
Mouse C3 sarcoma HPV-16 Splenic HPV-16 E7 peptide Days pre IV Tumor protection Ossevoort 1995536 
Mouse Lewis lung-(3LL) carcinoma BM generated (GM-CSF/IL-4) MUT-1 peptide 0-7 Days post IV or SC No tumor growth Mayordomo et al 1995202 
 C3 sarcoma-HPV16 melanoma-OVAM-05  HPV16-E7 peptide OVA peptide 7, 15, 21 Days post Established tumor regression 
Mouse MCA-induced fibrosarcoma (MCA-205) mammary adenocarcinoma (TS/A) BM generated (GM-CSF/IL-4) Acid-eluted tumor peptides 4-7 Days post IV (multiple) Reduced tumor growth Zitvogel et al 1996426 
 C3 embryonal fibroblasts-HPV 16 
Mouse Melanoma (B16-OVA) BM generated (GM-CSF/IL-4) OVA peptide 0, 7 Days pre SC No tumor growth Celluzzi et al 1996203 
Mouse LBSN retroviral transformed fibroblast (F1-B gal) mastocytoma (P815-B gal) BM generated (GM-CSF) transduced DC line D2SC/1 B gal protein 10-12 & 5-7 Days pre IP No tumor growth Paglia et al 1996537 
Mouse Thymoma (EL-4-OVA) BM generated (GM-CSF) OVA peptide 8 Days pre IV/IP 5/6 relapse (27 tumor cells) Porgador et al 1996529 
     5/6 free (107 tumor cells)  
Mouse 3T3 Cells-human p53 mutant (D459) BM generated (GM-CSF/IL-4) p53 Mutant peptide 8 Days then every 4-5 days (post) Reduced tumor growth Gabrilovich et al 1996427 
Mouse MCA-induced fibrosarcoma (natural p53 mutant) BM generated (GM-CSF/IL-4 or TNF) p53 Mutant peptide 7 & 14 Days pre IV Reduced tumor growth Mayordomo et al 1996428 
    7 & 14 Days post IV Tumor regression or inhibition 
Mouse MC38 adenocarcinoma BM generated (GM-CSF)/MC38 fusion cells MC38 fusion cell 14 & 28 Days pre SC or IV No tumor growth Gong et al 1997526 
    4 & 18 Days post IV 24/25 No metastases at 28 days 
Rat Sarcoma (HSNic) Mesenteric lymphatic DC In vivo loading via PP 5 Weeks pre IP vaccination for each 9/16 Tumor free Gyure et al 1987263 
   In vitro loading  0/5 Tumor free (reduced tumor growth) 
Human Non-Hodgkin's lymphoma Blood DC Ig idiotype 0, 1, 2 L 5-6 mo (+14 day protein boost) post SLT-lymphocyte response. Possible tumor response Hsu et al 1996532 
Human Prostate cancer GM/IL-4 Mo-DC Prostate-specific membrane antigen 4-6 Weekly (≤4 IV infusions) Possible T-lymphocyte response Murphy et al 1996533 
SpeciesTumorDC PreparationTumor AGVaccinationOutcomeAuthor (reference)
(Pre/post)
Mouse MCA-induced sarcoma (McSa-1) Spleen DC Sarcoma cells 18 Days post IV Some dose-dependent prolonged survival Knight et al 1985519 
    0 Day IV All died 
Mouse Fibrosarcoma (CSAIM) T-depleted spleen cells Tumor membranes 5 Weekly pre SC Reduced tumor growth Shimizu et al 1991502 
Mouse MCA-induced fibrosarcoma (S1509a) Unseparated epithelial cells (%LC) Tumor lysate 7 Days pre SC Reduced tumor growth Grabbe et al 1991525 
Mouse Lymphoma (BCL1) Splenic low density (BSA) IgM idiotype 10 Days pre IV 23/28 Treated alive Flamand et al 1994535 
    5 Days Ig IV 2/20 Untreated alive 
Mouse C3 sarcoma HPV-16 Splenic HPV-16 E7 peptide Days pre IV Tumor protection Ossevoort 1995536 
Mouse Lewis lung-(3LL) carcinoma BM generated (GM-CSF/IL-4) MUT-1 peptide 0-7 Days post IV or SC No tumor growth Mayordomo et al 1995202 
 C3 sarcoma-HPV16 melanoma-OVAM-05  HPV16-E7 peptide OVA peptide 7, 15, 21 Days post Established tumor regression 
Mouse MCA-induced fibrosarcoma (MCA-205) mammary adenocarcinoma (TS/A) BM generated (GM-CSF/IL-4) Acid-eluted tumor peptides 4-7 Days post IV (multiple) Reduced tumor growth Zitvogel et al 1996426 
 C3 embryonal fibroblasts-HPV 16 
Mouse Melanoma (B16-OVA) BM generated (GM-CSF/IL-4) OVA peptide 0, 7 Days pre SC No tumor growth Celluzzi et al 1996203 
Mouse LBSN retroviral transformed fibroblast (F1-B gal) mastocytoma (P815-B gal) BM generated (GM-CSF) transduced DC line D2SC/1 B gal protein 10-12 & 5-7 Days pre IP No tumor growth Paglia et al 1996537 
Mouse Thymoma (EL-4-OVA) BM generated (GM-CSF) OVA peptide 8 Days pre IV/IP 5/6 relapse (27 tumor cells) Porgador et al 1996529 
     5/6 free (107 tumor cells)  
Mouse 3T3 Cells-human p53 mutant (D459) BM generated (GM-CSF/IL-4) p53 Mutant peptide 8 Days then every 4-5 days (post) Reduced tumor growth Gabrilovich et al 1996427 
Mouse MCA-induced fibrosarcoma (natural p53 mutant) BM generated (GM-CSF/IL-4 or TNF) p53 Mutant peptide 7 & 14 Days pre IV Reduced tumor growth Mayordomo et al 1996428 
    7 & 14 Days post IV Tumor regression or inhibition 
Mouse MC38 adenocarcinoma BM generated (GM-CSF)/MC38 fusion cells MC38 fusion cell 14 & 28 Days pre SC or IV No tumor growth Gong et al 1997526 
    4 & 18 Days post IV 24/25 No metastases at 28 days 
Rat Sarcoma (HSNic) Mesenteric lymphatic DC In vivo loading via PP 5 Weeks pre IP vaccination for each 9/16 Tumor free Gyure et al 1987263 
   In vitro loading  0/5 Tumor free (reduced tumor growth) 
Human Non-Hodgkin's lymphoma Blood DC Ig idiotype 0, 1, 2 L 5-6 mo (+14 day protein boost) post SLT-lymphocyte response. Possible tumor response Hsu et al 1996532 
Human Prostate cancer GM/IL-4 Mo-DC Prostate-specific membrane antigen 4-6 Weekly (≤4 IV infusions) Possible T-lymphocyte response Murphy et al 1996533 

A peptide specific for the rat Lewis lung cancer is a potent vaccine and causes regression of metastases in mice.527 In a further report, three different tumor models were investigated in mice using viral antigenic target (HPV-16 sarcoma), ovalbumin target (OVA-melanoma), and unmodified Lewis lung cancer.202 Vaccination with Ag pulsed, GM-CSF/IL-4–generated, BM-derived DC produced dramatic tumor responses. Subsequent reports203,426-428,528 529 have confirmed the efficacy of these BM generated “DC preparations” in a variety of mouse tumor models (see Table 2).

Acid-eluted tumor peptides also provide sufficient TAA to immunize against tumors426 and the resulting immune responses are highly specific,426,502 both of which augur well for clinical application. So too does the superiority of DC vaccination over Freunds incomplete adjuvant.428 However, many of the Ags used are transfected viral Ags or highly immunogenic proteins and the generality of the phenomenon with other tumor models needs to be confirmed. A model in which DC tumor vaccination in mice did not appear to be successful519 might be helpful.

The fact that patients with breast cancer respond to MUC-1 Ag530 and mice are protected against human breast cancer implants when vaccinated with oxidized mannan MUC-1531 suggests that this Ag, incubated with DC, may become a new treatment modality for this disease. Clinical studies are in progress using melanoma Ag. Perhaps bcr-abl peptide vaccination could be contemplated in appropriate HLA-typed CML patients either after IFN or autologous BMT has reduced the leukemic load or in patients with early relapse after allogeneic BMT. Percoll gradient, immunoselected DC have been used with tumor Ig idiotype in a phase 1 study to vaccinate patients with low-grade NHL.532 Significant T-lymphocyte and antibody responses to the Ig idiotype were detected, as were some clinical changes suggesting a tumor response. In another phase 1 study, autologous DC pulsed with PSMA-derived peptides were infused into patients with prostate cancer without ill effects and perhaps some effect on T-lymphocyte reactivity to PSMA.533 

Once variables regarding DC source, Ag preparation, Ag exposure, and route of administration are clarified, further phase I clinical trials will be appropriate. Incubation with antibody,149 incorporation with pH sensitive liposomes,286 triacylation of peptide,286 oxidized mannan conjugation,531 or fusion constructs with hsp70534 may all improve Ag/peptide loading into DC. The timing of Ag exposure to human blood DC preparations may be critical (submitted) and serum should be avoided because it contains peptidases.342 Once the technology is optimized, the use of DC to generate effective tumor-specific immune responses may become the third modality of anticancer treatment.

DC should now be recognized as unique subsets of leukocytes. Myeloid-derived DC have a unique ability to induce primary immune responses, but a lymphoid progenitor–derived DC with different properties has been described. It is unclear how much differentiation of Mo into DC occurs in vivo compared with the differentiation of Mo-DC induced by cytokines in vitro. DC show the essential properties required of APC: migration, Ag uptake, processing, and presentation of Ag and costimulation of lymphocytes. The MoAb, CD83, CMRF-44 and CMRF-56, are proving of value for studying human DC and a routine DC blood count based on CMRF-44 staining is now available (in preparation). The molecular events controlling DC function are not yet fully understood and their transcriptional control of critical genes will be an important new area. Surveillance DC respond to danger signals and initiate central migration and upregulation of Ag-presenting activity. The control of DC costimulation function is such that Ag-specific recognition and T-lymphocyte reciprocal signaling may render the DC Ag-presenting function effectively Ag specific. As such myeloid DC, as well as lymphoid DC, may contribute to central and peripheral tolerance. Better control of DC antigen-presenting function may improve BMT procedures and allow optimization of DC as “nature's adjuvant” for cancer immunotherapy.

Mouse BM-derived DC have been shown to phagocytose and process viable bacteria for presentation of both MHC class I and class II defined epitopes to T lymphocytes.538 

Novel CC chemokine receptors have been identified on human DC539,540 which bind the recently described CC chemokine MIP-3α. MIP-3α is chemotactic for CD34+-derived DC and T lymphocytes but is inactive on monocytes. CCR6 is expressed on CD34+-derived DC but not Mo-DC,540 further emphasizing the distinction necessary between these cell preparations. A new CC chemokine (DC-CK1) has been identified recently using an Mo-DC cDNA library.541 DC-CK1 which is detected in IDC by in situ hybridization has the most interesting property of preferentially attracting naive (CD45RA+) T lymphocytes. Another CC chemokine, macrophage-derived chemokine (MDC), is expressed by Mo-DC and induces their migration.542 

The presence of DC in human cord blood which was predicted on the basis of its allostimulatory potential has been confirmed (unpublished). Preliminary data suggest that cyclophosphamide/G-CSF–mobilized peripheral blood stem cells may contain relatively reduced numbers of blood DC (in preparation). This intriguing result may partially explain the lower-than-expected GVHD result that has been reported with allogeneic peripheral blood stem cell grafts.

Exciting data have been published linking the recruitment of DC into tumors with immune-mediated tumor regression in flt-3 ligand–treated mice.543 This was accompanied by a significant nonimmunologically mediated effect on tumor growth. First reports on phase I studies in humans suggest flt-3 ligand may be useful for mobilizing DC for cancer immunotherapy.544 Recent experiments describing a transient state of functional unresponsiveness after priming with DC pulsed with the P815AB TAA peptide, reversible by IL-12 administration,545 emphasize the need for careful development of clinical protocols.

I am grateful to colleagues who contributed unpublished data to assist with writing this review. Dr G.J. Clark had major input into the preparation of the text and Dr D.B. Fearnley provided invaluable assistance. The constructive criticism of Dr M.A. Baird, Dr A. Heiser, Dr M. Kato, A.B. McLellan, Dr R.V. Sorg, Dr W.N. Patton, Dr A.B. Troy, and Dr S. Vuckovic was invaluable and greatly appreciated. S. Banks provided the considerable secretarial support required to prepare this review.

The Haematology/Immunology/Transfusion Medicine Research Grants have been supported by the New Zealand Health Research Council, Canterbury Medical Research Foundation, Cancer Society of New Zealand, Lottery Health Research Council, McClelland Trust, Leukemia and Blood Foundation of New Zealand, Loch Maben Trust, University of Otago, and others.

Address reprint requests to Derek N.J. Hart, MD, D Phil, Haematology/Immunology/Transfusion Medicine Research Group, Canterbury Health Laboratories, PO Box 151, Christchurch, New Zealand.

1
Steinman
RM
Cohn
ZA
Identification of a novel cell type in peripheral lymphoid organs of mice. I. Morphology, quantitation, tissue distribution.
J Exp Med
137
1973
1142
2
Steinman
RM
Lustig
DS
Cohn
ZA
Identification of a novel cell type in peripheral lymphoid organs of mice. III. Functional properties in vivo.
J Exp Med
139
1974
1431
3
Steinman
RM
Witmer
MD
Lymphoid dendritic cells are potent stimulators of the primary mixed leucocyte reaction in mice.
Proc Natl Acad Sci USA
75
1978
5132
4
Witmer
MD
Steinman
RM
The anatomy of peripheral light-microscopic immunocytochemical studies of mouse spleen, lymph node and peyers patch.
Am J Anat
170
1984
465
5
Inaba
K
Witmer-Pack
MD
Steinman
RM
Clustering of dendritic cells, helper T lymphocytes and histocompatible B cells, during primary antibody responses in vitro.
J Exp Med
160
1984
858
6
Hart
DNJ
Fabre
JW
MHC antigens in rat kidney ureter and bladder: Localisation with monoclonal antibodies and demonstation of Ia positive dendritic cells.
Transplantation
31
1981
318
7
Hart
DNJ
Fabre
JW
Demonstration and characterization of Ia-positive dendritic cells in the interstitial connective tissues of rat heart and other tissues but not brain.
J Exp Med
153
1981
347
8
Hart
DNJ
Newton
MR
Reece-Smith
H
Fabre
JW
Morris
PJ
Major histocompatibility complex antigens in the rat pancreas, isolated pancreatic islets, thyroid and adrenal: Localization with monoclonal antibodies and demonstration of interstitial dendritic cells.
Transplantation
36
1983
431
9
Hart
DNJ
Fuggle
SV
Williams
KA
Fabre
JW
Ting
A
Morris
PJ
Localization of HLA-ABC and DR antigens in human kidney.
Transplantation
31
1981
428
10
Daar
AS
Fuggle
SV
Hart
DNJ
Dalchau
R
Abdulaziz
Z
Fabre
JW
Ting
A
Morris
PJ
Demonstration and phenotypic characterization of HLA-DR positive interstitial dendritic cells widely distributed in human connective tissues.
Trans Proc
1
1983
311
11
Hart
DNJ
Fabre
JW
Localization of MHC antigens in long surviving rat renal allografts: Probable implication of passenger leucocytes in graft adaption.
Trans Proc
13
1981
95
12
Hart
DNJ
Fabre
JW
The mechanism of induction of passive enhancement: Evidence for an interaction of enhancing antibody with donor interstitial dendritic cells.
Transplantation
33
1981
319
13
McKenzie
JL
Beard
MEJ
Hart
DNJ
Depletion of donor kidney dendritic cells prolongs graft survival.
Trans Proc
16
1984
948
14
McKenzie
JL
Beard
MEJ
Hart
DNJ
The effect of donor pretreatment on interstitial dendritic cell content and rat cardiac allograft survival.
Transplantation
38
1984
371
15
Steinman
RM
The dendritic cell system and its role in immunogenicity.
Annu Rev Immunol
9
1991
271
16
Hart DNJ, Calder VL: Human dendritic cells: Function and cytokine production. Immunopharmacology of macrophages and other antigen-presenting cells, in Bruijnzeel-Koomen CFAM, Hoefsmit ECM (eds): Handbook of Immunopharmacology. London, UK, Academic, 1994, p 63
17
Markowicz
S
Engleman
EG
GM-CSF promotes differentiation and survival of human peripheral blood dendritic cells.
J Clin Invest
85
1990
955
18
Reid
CD
Fryer
PR
Clifford
C
Kirk
A
Tikerpae
J
Knight
SC
Identification of haemopoietic progenitors of macrophages and dendritic Langerhans cells (DL-CFU) in human bone marrow and peripheral blood.
Blood
76
1990
1139
19
Inaba
K
Steinman
RM
Witmer-Pack
M
Aya
H
Inaba
M
Sudo
T
Wolpe
S
Schuler
G
Identification of proliferating dendritic cell precursors in mouse blood.
J Exp Med
175
1992
1157
20
Inaba
K
Inaba
M
Romani
N
Hideki
A
Deguchi
M
Ikehara
S
Muramatsu
S
Steinman
RM
Generation of large numbers of dendritic cells from mouse bone marrow cultures supplemented with GM-CSF.
J Exp Med
176
1992
1693
21
Scheicher
C
Mehlig
M
Zecher
R
Reske
K
Dendritic cells from mouse bone marrow: In vitro differentiation using low doses of recombinant granulocyte-macrophage colony-stimulating factor.
J Immunol Methods
154
1992
253
22
Hock
BD
Starling
GC
Daniel
PB
Hart
DNJ
Characterisation of CMRF-44, a novel monoclonal antibody to an activation antigen expressed by the allostimulatory cells within peripheral blood, including dendritic cells.
Immunology
83
1994
573
23
Zhou
L
Schwarting
R
Smith
HM
Tedder
TF
A novel cell-surface molecule expressed by human interdigitating reticulum cells, Langerhans cells, and activated lymphocytes is a new member of the Ig superfamily.
J Immunol
149
1992
735
24
Austyn
JA
Larsen
CP
Migration patterns of dendritic leukocytes.
Transplantation
49
1990
1
25
McLellan
A
Starling
GC
Williams
LA
Hock
B
Hart
DNJ
Activation of human peripheral blood dendritic cells induces the CD86 costimulatory molecule.
Eur J Immunol
25
1995
2064
26
McLellan
AD
Sorg
RV
Williams
LA
Hart
DNJ
Human dendritic cells activate T lymphocytes via a CD40:CD40 ligand-dependent pathway.
Eur J Immunol
26
1996
1204
27
Williams
L
Egner
W
Hart
DNJ
Isolation and function of human dendritic cells.
Int Review Cytol
153
1994
41
28
Schuler G, Romani N: Epidermal Langerhans cells as a model for studying the effects of GM-CSF and other cytokines on the biology of dendritic cells, in Van Furth R (ed): Haemopoietic Growth Factors and Mononuclear Phagocytes. Basel, Switzerland, Karger, 1993, p 197
29
Ardavin
C
Shortman
K
Thymic dendritic cells and T cells develop simultaneously in the thymus from a common precursor population.
Nature
362
1993
761
30
Fairchild
PJ
Austyn
JM
Thymic dendritic cells: Phenotype and function.
Int Rev Immunol
6
1990
187
31
Matzinger
P
Guerder
S
Does T-cell tolerance require a dedicated antigen-presenting cell?
Nature
338
1989
74
32
Kronin
V
Winkel
K
Suss
G
Kelso
A
Heath
W
Kirberg
J
von Boehmer
H
Shortman
K
A subclass of dendritic cells regulates the response of naive CD8 T cells by limiting their IL-2 production.
J Immunol
157
1996
3819
33
Romani
N
Schuler
G
The immunologic properties of epidermal Langerhans cells as a part of the dendritic cell system.
Springer Semin Immunopathol
13
1992
265
34
Kleijmeer
MJ
Ossevoort
MA
van Veen
CJH
van Hellemond
JJ
Neefjes
JJ
Kast
WM
Melief
CJM
Geuze
HJ
MHC class II compartments and the kinetics of antigen presentation in activated mouse spleen dendritic cells.
J Immunol
154
1995
5715
35
Nijman
HW
Kleijmeer
MJ
Ossevoort
MA
Oorschot
VMJ
Vierboom
MPM
van de Keur
M
Kenemans
P
Kast
WM
Geuze
HJ
Melief
CJM
Antigen capture and major histocompatibility class II compartments of freshly isolated and cultured human blood dendritic cells.
J Exp Med
182
1995
163
36
Inaba
K
Steinman
RM
Accessory cell-T lymphocyte interactions. Antigen dependent and independent clustering.
J Exp Med
48
1986
1039
37
Holt
PG
Schon-Hegrad
MA
Oliver
J
Localisation of T-cells, macrophages and dendritic cells in rat respiratory tract tissues: Implications for immune function studies.
Immunology
62
1987
349
38
Witmer-Pack
MD
Olivier
W
Valinsky
J
Schuler
G
Steinman
RM
Granulocyte/macrophage colony-stimulating factor is essential for the viability and function of cultured murine epidermal Langerhans cells.
J Exp Med
166
1987
1484
39
Romani
N
Lenz
A
Glassel
H
Stossel
H
Stanzl
U
Majdic
O
Fritsch
P
Schuler
G
Cultured human Langerhans cells resemble lymphoid dendritic cells in phenotype and function.
J Invest Dermatol
93
1989
600
40
Schuler
G
Steinman
RM
Murine epidermal Langerhans cells mature into potent allostimulatory dendritic cells in vitro.
J Exp Med
161
1985
526
41
Inaba
K
Schuler
G
Witmer
MD
Valinsky
J
Atassi
B
Steinman
RM
Immunologic properties of purified epidermal Langerhans cells.
J Exp Med
164
1986
605
42
Teunissen
MBM
Wormmeester
DJ
Kreig
SR
Peters
PJ
Vogels
IMC
Kapsenberg
ML
Bos
JD
Human epidermal Langerhans cells undergo profound morphologic and phenotypic changes during in vitro culture.
J Invest Dermatol
94
1990
166
43
Austyn
JA
Hankins
DF
Larsen
CP
Morris
PJ
Rao
AS
Roake
JA
Isolation and characterization of dendritic cells from mouse heart and kidney.
J Immunol
152
1994
2401
44
Egner
W
McKenzie
JL
Smith
SM
Beard
MEJ
Hart
DNJ
Identification of potent mixed leucocyte reaction-stimulatory cells in human bone marrow.
J Immunol
150
1993
3043
45
Drexhage
HA
Mullink
H
de Groot
J
Clarke
J
Balfour
BM
A study of cells present in peripheral lymph of pigs with special reference to a type of cell resembling the Langerhans cell.
Cell Tissue Res
202
1979
407
46
Hart
DNJ
McKenzie
JL
Isolation and characterization of human tonsil dendritic cells.
J Exp Med
168
1988
157
47
Agger
R
Witmer-Pack
M
Romani
N
Stossel
H
Swiggard
WJ
Metlay
JP
Storozynsky
E
Freimuth
P
Steinman
RM
Two populations of splenic dendritic cells detected with M342, a new monoclonal to an intracellular antigen of interdigitating dendritic cells and some B lymphocytes.
J Leuk Biol
52
1992
34
48
Will
A
Blank
C
Rollinghof
M
Moll
H
Murine epidermal Langerhans cells are potent stimulators of an antigen-specific T cell response to Leishmania major, the cause of cutaneous leishmaniasis.
Eur J Immunol
22
1992
1341
49
Matsuno
K
Ezaki
T
Kudo
S
Uehara
Y
A life stage of particle-laden rat dendritic cells in vivo: their terminal division, active phagocytosis and translocation from the liver to the draining lymph.
J Exp Med
183
1996
1865
50
Barfoot
R
Denham
S
Gyure
LA
Hall
JG
Hobbs
SM
Jackson
LE
Robertson
D
Some properties of dendritic macrophages from peripheral lymph.
Immunology
68
1989
233
51
Freudenthal
PS
Steinman
RM
The distinct surface of human blood dendritic cells, as observed after an improved isolation method.
Proc Natl Acad Sci USA
87
1990
7698
52
Van Voorhis
WC
Valinsky
J
Hoffmann
E
Luban
J
Hair
LS
Steinman
RM
Relative efficacy of human monocytes and dendritic cells as accessory cells for T cell replication.
J Exp Med
158
1983
174
53
Ashwell
JD
Chen
C
Schwartz
RH
High frequency and nonrandom distribution of alloreactivity in T cell clones selected for recognition of foreign antigen in association with self class II molecules.
J Immunol
136
1986
389
54
Sherman
LW
Chattopadhyay
S
The molecular basis of allorecognition.
Annu Rev Immunol
11
1993
385
55
Krieger
JI
Chesnut
RW
Grey
HM
Capacity of B cells to function as stimulators of a primary mixed leukocyte reaction.
J Immunol
137
1986
3117
56
Metlay
JP
Pure
E
Steinman
RM
Distinct features of dendritic cells and anti-Ig activated B cells as stimulators of the primary mixed leukocyte reaction.
J Exp Med
169
1989
239
57
Egner
W
Andreesen
R
Hart
DNJ
Allostimulatory cells in fresh human blood: Heterogeneity in antigen presenting cell populations.
Transplantation
56
1993
945
58
Thomas
R
Davis
LS
Lipskey
PE
Comparative accessory cell function of human peripheral blood dendritic cells and monocytes.
J Immunol
151
1993
6840
59
Inaba
K
Romani
N
Steinman
RM
An antigen-independent contact mechanism as an early step in T cell proliferative responses to dendritic cells.
J Exp Med
170
1989
527
60
Prickett
TCR
Hart
DNJ
Anti-leucocyte common (CD45) antibodies inhibit dendritic cell stimulation of CD4 and CD8 T lymphocyte proliferation.
Immunology
69
1990
250
61
Kuntz-Crow
M
Kunkel
HG
Human dendritic cells: Major stimulators of autologous and allogeneic MLR.
Clin Exp Immunol
49
1982
338
62
Pure
E
Inaba
K
Crowley
MT
Tardelli
L
Witmer-Pack
MD
Ruberti
G
Fathman
G
Steinman
RM
Antigen processing by epidermal Langerhans cells correlates with the level of biosynthesis of MHC class II molecules and expression of invariant chain.
J Exp Med
172
1990
1459
63
Inaba
K
Metlay
JP
Crowley
MT
Steinman
RM
Dendritic cells pulsed with protein antigens in vitro can prime antigen-specific, MHC-restricted T cells in situ.
J Exp Med
172
1990
631
64
Knight
SC
Krejci
J
Malkovsky
M
Colizzi
V
Gautum
A
Asherson
GL
The role of dendritic cells in the initiation of immune responses to contact sensitizers.
Cell Immunol
94
1985
427
65
Macatonia
SE
Knight
SC
Edwards
AJ
Griffiths
S
Fryer
P
Localization of antigen on lymph node dendritic cells after exposure to the contact sensitizer fluorescein isothiocyanate.
J Exp Med
166
1987
1654
66
Grainger
R
Hart
DNJ
Watson
JD
Baird
MA
Antigen-pulsed, IL-4 treated B cells activate primed T cells in vitro but not naive T cells in vivo.
Scand J Immunol
42
1995
517
67
Guery
J
Ria
F
Adorini
L
Dendritic cells but not B cells present antigenic complexes to class II-restricted T cells after administration of protein in adjuvant.
J Exp Med
183
1996
751
68
McLellan
AD
Starling
GC
Hart
DNJ
Isolation of human blood dendritic cells by Nycodenz discontinuous gradient centrifugation.
J Immunol Methods
184
1995
81
69
Young
JW
Steinman
RM
Accessory cell requirements for the MLR and polyclonal mitogens, as studied with a new technique for enriching blood dendritic cells.
Cell Immunol
111
1988
167
70
Xu
H
Friedrichs
U
Gieseler
RKH
Ruppert
J
Ocklind
G
Peters
JH
Human blood dendritic cells exhibit a distinct T-cell stimulating mechanism and differentiation pattern.
Scand J Immunol
36
1992
689
71
Van Voorhis
WC
Hair
LS
Steinman
RM
Kaplan
G
Human dendritic cells. Enrichment and characterisation from peripheral blood.
J Exp Med
155
1982
1172
72
Brooks
CF
Moore
M
Differential MHC class II expression on human peripheral blood monocytes and dendritic cells.
Immunology
63
1988
303
73
Knight
SC
Farrant
J
Bryant
A
Edwards
AJ
Burman
S
Lever
A
Clarke
J
Webster
ADB
Non-adherent, low density cells from human peripheral blood contain dendritic cells and monocytes, both with veiled morphology.
Immunology
57
1986
595
74
McLellan AD, Sorg RV, Fearnley DB, Hart DNJ: Fresh human Langerhans cells express the CD83 and CMRF-44 differentiation/activation antigens and costimulate T lymphocytes via CD86. (submitted) 1996
75
Kabel
PJ
de Haan-Meulman
M
Voorbij
AM
Kleingeld
M
Knol
EF
Drexhage
HA
Accessory cells with a morphology and marker pattern of dendritic cells can be obtained from elutriator-purified blood monocyte fractions. An enhancing effect of metrizamide in this differentiation.
Immunobiology
179
1989
395
76
Schumann
RR
van der Bosch
J
Ruller
S
Ernst
M
Schlaak
M
Monocyte long-term cultivation on microvascular endothelial cell monolayers: Morphologic and phenotypic characterization and comparison with monocytes cultured on tissue culture plastic.
Blood
73
1989
818
77
Fearnley
DB
McLellan
AD
Mannering
SI
Hock
BD
Hart
DNJ
Isolation of human blood dendritic cells using the CMRF-44 monoclonal antibody: Implications for studies on antigen presenting cell function and immunotherapy.
Blood
89
1997
3708
78
Zhou
L
Tedder
TF
Human blood dendritic cells selectively express CD83, a member of the immunoglobulin superfamily.
J Immunol
154
1995
3821
79
Buckley
PJ
Smith
MR
Braverman
MF
Dickson
SA
Human spleen contains phenotypic subsets of macrophages and dendritic cells that occupy discrete microanatomic locations.
Am J Pathol
128
1987
505
80
Thomas
R
Davis
LS
Lipsky
PE
Isolation and characterization of human peripheral blood dendritic cells.
J Immunol
150
1993
821
81
MacPherson
G
Properties of lymph-borne (veiled) cells in culture 1. Modulation of phenotype, survival and function: Partial dependence on GM-CSF.
Immunology
68
1989
102
82
Inaba
K
Inaba
M
Deguchi
M
Hagi
K
Yasumizu
R
Ikehara
S
Muramatsu
S
Steinman
RM
Granulocytes, macrophages and dendritic cells arise from a common MHC class II-negative progenitor in mouse bone marrow.
Proc Natl Acad Sci USA
90
1993
3038
83
Bowers
WE
Berkowitz
MR
Differentiation of dendritic cells in cultures of rat bone marrow cells.
J Exp Med
163
1986
872
84
Egner
W
Hart
DNJ
The phenotype of freshly isolated and cultured human bone marrow allostimulatory cells: Heterogeneity in dendritic cell populations.
Immunology
85
1995
611
85
Thomas
R
Lipsky
PE
Human peripheral blood dendritic cell subsets: Isolation and characterization of precursor and mature antigen presenting cells.
J Immunol
153
1994
4016
86
Metlay
JP
Witmer-Pack
MD
Agger
R
Crowley
MT
Lawless
D
Steinman
RM
The distinct leucocyte integrins of mouse spleen DC as identified with new hamster mAb.
J Exp Med
171
1990
1753
87
Calabi
F
Bradbury
A
The CD1 system.
Tissue Antigens
37
1991
1
88
Darden
AG
Clarke
Forbes RD
Darden
PM
Guttman
RD
The effect of genetics and age on expression of MHC class II and CD4 antigens on rat cardiac interstitial dendritic cells.
Cell Immunol
126
1990
322
89
Aidong
Y
Baird
MA
Changes in the density and function of rat spleen dendritic cells occur with age.
Immunol Infect Dis
5
1994
121
90
Liu
Y
Barthelemy
C
Bouteiller
O
Arpin
C
Durand
I
Banchereau
J
Memory B cells from human tonsils colonize mucosal epithelium and directly present antigen to T cells by rapid upregulation of B7.1 and B7.2.
Immunity
2
1995
239
91
Hock
BD
Hart
DNJ
Cellular protein profiles of human dendritic cells; a comparative study.
Transplant Proc
24
1992
2323
92
Nussenzweig
MC
Steinman
RM
Witmer
MD
Gutchinov
B
A monoclonal antibody specific for mouse dendritic cells.
Proc Natl Acad Sci USA
79
1982
161
93
Kraal
G
Breel
M
Janse
M
Bruin
G
Langerhans cells, veiled cells and interdigitating cells in the mouse recognised by a monoclonal antibody.
J Exp Med
163
1986
981
94
Jiang
W
Swiggard
WJ
Heufler
C
Peng
M
Mirza
A
Steinman
RM
Nussenzweig
MC
The receptor DEC-205 expressed by dendritic cells and thymic epithelial cells is involved in antigen processing.
Nature
375
1995
151
95
Stahl
PD
The mannose receptor and other macrophage lectins.
Current Opin Immunol
4
1992
49
96
Brenan
M
Puklavec
M
The MRC OX-62 antigen: A useful marker in the purification of rat veiled cells with the biochemical properties of an integrin.
J Exp Med
175
1992
1457
97
Davis
AL
McKenzie
JL
Hart
DNJ
HLA-DR positive leukocyte subpopulations in human skin include dendritic cells, macrophages and CD7-negative T cells.
Immunology
65
1988
573
98
Richters
CD
Reits
EAJ
Vanpelt
AM
Hoekstra
MJ
Vanbaare
J
Dupont
JS
Kamperdijk
EWA
Effect of low dose UVB irradiation on the migratory properties and functional capacities of human skin dendritic cells.
Clin Exp Immunol
104
1996
191
99
Cattoretti G, Berti E, Mancuso A, D'Amato L, Schiro R, Soligo D, Delia D: CD1, a MHC class I related family of antigens with widespread distribution on resting and activated cells, in McMichael AJ (ed): Leucocyte Typing III. Oxford, UK, Oxford University Press, 1987, p 89
100
Kasinrerk
W
Baumruker
T
Majdic
O
Knapp
W
Stockinger
H
CD1 molecule expression on human monocytes induced by GM-CSF.
J Immunol
150
1993
579
101
Caux
C
Dezutter-Dambuyant
C
Schmitt
D
Banchereau
J
GM-CSF and TNFα cooperate in the generation of dendritic Langerhans cells.
Nature
360
1992
258
102
Goordyal
P
Issacson
PG
Immunocytochemical characterisation of monocyte colonies of human bone marrow: A clue to the origin of Langerhans cells and interdigitating reticulum cells.
J Pathol
146
1985
189
103
Hogg
N
Human mononuclear phagocyte molecules and the use of monoclonal antibodies in their detection.
Clin Exp Immunol
69
1987
687
104
Weber-Matthiesen
K
Sterry
W
Organisation of the monocyte/macrophage system of normal human skin.
J Invest Dermatol
95
1990
83
105
O'Doherty
U
Steinman
RM
Peng
M
Cameron
PU
Gezelter
S
Dendritic cells freshly isolated from human blood express CD4 and mature into typical immunostimulatory dendritic cells after culture in monocyte-conditioned medium.
J Exp Med
178
1993
1067
106
Wuerzner
R
Xu
H
Franzke
A
Schulze
M
Peters
JH
Goetze
O
Blood dendritic cells carry terminal complement complexes on their cell surface as detected by newly developed neoepitope-specific monoclonal antibodies.
Immunology
74
1991
132
107
Poulter
LW
Campbell
DA
Munroe
C
Janossy
G
Discrimination of human macrophages and dendritic cells by means of monoclonal antibodies.
Scand J Immunol
24
1986
51
108
Kashihara
M
Ueda
M
Horiguchi
Y
Furukawa
F
Hanaoka
M
Imamura
S
A monoclonal antibody specifically reactive to human Langerhans cells.
J Invest Dermatol
87
1986
602
109
Hsu
PL
Hsu
SM
Identification of an Mr 70,000 antigen associated with Reed-Sternberg cells and interdigitating reticulum cells.
Cancer Res
50
1990
350
110
Cerio
R
Griffiths
CEM
Cooper
KD
Nickoloff
BJ
Headington
JT
Characterization of factor XIIIa positive dermal dendritic cells in normal and inflamed skin.
Br J Dermatol
121
1989
421
111
Twist CJ, Disteche C, Beier D, Tedder TF: Structure of the mouse CD83 gene. Blood 86:323a, 1995 (abstr, suppl 1)
112
Wood
GS
Freudenthal
PS
Edinger
A
Steinman
RM
Warnke
RA
CD45 epitope mapping of human CD1+ dendritic cells and peripheral blood dendritic cells.
Am J Pathol
138
1991
1451
113
Hart DNJ, Fearnley D: Dendritic Cell Studies Report, in Kishimoto T, Miyasaka M, Mason D, Sugamura K, Springer T, Goyert S, Moretta L, Zola H, Kr vd Borne AEG, Okumura K: (eds): Leucocyte Typing VI. New York, NY, Garland, 1997
114
Egner
W
Hock
BD
Hart
DNJ
Dendritic cells have reduced cell surface membrane glycoproteins including CD43 determinants.
Adv Exp Med Biol
329
1993
71
115
Roake
JA
Rao
AS
Morris
PJ
Larsen
CP
Hankins
DF
Austyn
JM
Systemic lipopolysaccharide recruits dendritic cell progenitors to nonlymphoid tissues.
Transplantation
59
1995
1319
116
Lenz
A
Heine
M
Schuler
G
Romani
N
Human and murine dermis contain dendritic cells: Isolation by means of a novel method and phenotypical and functional characterization.
J Clin Invest
92
1993
2587
117
Nestle
FO
Zheng
X
Thompson
CB
Turka
LA
Nickoloff
BJ
Characterization of dermal dendritic cells obtained from normal human skin reveals phenotypic and functionally distinctive subsets.
J Immunol
151
1993
6535
118
Grouard
G
Rissoan
M
Filgueira
L
Durand
I
Banchereau
J
Liu
Y
The enigmatic plasmacytoid T cells develop into dendritic cells with interleukin (IL)-3 and CD40-ligand.
J Exp Med
185
1997
1101
119
Freeman
SD
Kelm
S
Barber
EK
Crocker
PR
Characterization of CD33 as a new member of the sialoadhesion family of cellular interaction molecules.
Blood
85
1995
2005
120
Wood
GS
Freudenthal
RS
CD5 monoclonal antibody reacts with human peripheral blood dendritic cells.
Am J Pathol
141
1992
789
121
Wright
SD
Ramos
RA
Tobias
PS
Ulevitch
RJ
Mathison
JC
CD14, a receptor for complexes of lipopolysaccharide (LPS) and LPS binding protein.
Science
249
1990
1431
122
Pugin
J
Heumann
D
Tomasz
A
Kravchenko
VV
Akamatsu
R
Nishijima
M
Glauser
MP
Tobias
PS
Ulevitch
RJ
CD14 is a pattern recognition receptor.
Immunity
1
1996
509
123
Mehta-Damani
A
Markowicz
S
Engelman
EG
Generation of antigen-specific CD8+ CTLs from naive precursors.
J Immunol
153
1994
996
124
Summers
KL
O'Donnell
JL
Sew
Hoy M
Peart
M
Dekker
J
Rothwell
A
Hart
DNJ
Monocyte-macrophage antigen expression on chondrocytes.
J Rheum
22
1995
1326
125
Lauener
RP
Goyert
SM
Geha
RS
Vercelli
D
Interleukin 4 down-regulates the expression of CD14 in normal human monocytes.
Eur J Immunol
20
1990
2375
126
Landmann
R
Wesp
M
Ludwig
R
Knusli
C
Obrecht
JP
Recombinant interferon gamma up-regulates in vivo and down-regulates in vitro monocyte CD14 antigen expression in cancer patients.
Cancer Immunol Immunother
31
1990
292
127
Szabolcs
P
Avigan
D
Gezelter
S
Ciocon
DH
Moore
MAS
Steinman
RM
Young
JW
Dendritic cells and macrophages can mature independently from a human bone marrow-derived, post-colony-forming unit intermediate.
Blood
87
1996
4520
128
Shan
HH
Talle
MA
Goldstein
G
Chess
L
Functional subsets of human monocytes defined by monoclonal antibodies: A distinct subset of monocytes contains the cells capable of inducing the autologous mixed lymphocyte culture.
J Immunol
130
1983
698
129
Zeng
L
Takeya
M
Takahashi
K
AM-3K, a novel monoclonal antibody specific for tissue macrophages and its application to pathological investigation.
J Pathol
178
1996
207
130
Takamizawa
M
Rivas
A
Fagnoni
F
Benike
C
Kosek
J
Hyakawa
H
Engelman
EG
Dendritic cells that process and present nominal antigens to naive T lymphocytes are derived from CD2+ precursors.
J Immunol
158
1997
2134
131
Hancock
WW
Atkins
RC
Immunohistologic analysis of the cell surface antigens of human dendritic cells using monoclonal antibodies.
Trans Proc
16
1984
963
132
Ferbas
JJ
Toso
JF
Logar
AJ
Navratil
JS
Rinaldo
CR
CD4+ blood dendritic cells are potent producers of IFN-α in response to in vitro HIV-1 infection.
J Immunol
152
1994
4649
133
Vremec
D
Zorbas
M
Scollay
R
Saunders
DJ
Ardavin
CF
The surface phenotype of dendritic cells purified from mouse thymus and spleen: Investigation of the CD8 expression by a subpopulation of dendritic cells.
J Exp Med
176
1992
47
134
Prickett
TCR
McKenzie
JL
Hart
DNJ
Adhesion molecules on human tonsil dendritic cells.
Transplantation
53
1992
483
135
Starling
GC
Egner
W
McLellan
AD
Fawcett
J
Simmons
DL
Hart
DNJ
Intercellular adhesion molecule-3 is a costimulatory ligand for LFA-1 expressed on human blood dendritic cells.
Eur J Immunol
25
1995
2528
136
Hart
DNJ
Prickett
TCR
Intercellular adhesion molecule-2 (ICAM-2) expression on human dendritic cells.
Cell Immunol
148
1993
447
137
Rambukkana
A
Bos
JD
Irik
D
Menko
WJ
Kapsenberg
ML
Das
PK
In situ behaviour of human Langerhans cells in skin organ culture.
Lab Invest
73
1995
521
138
Vedel
J
Vincendeau
P
Bezian
JH
Taieb
A
Flow cytometry analysis of adhesion molecules on human Langerhans cells.
Clin Exp Dermatol
17
1992
240
139
Zambruno
G
Cossarizza
A
Zacchi
V
Ottani
D
Luppi
AM
Gianetti
A
Girolomoni
G
Functional intercellular adhesion molecule-3 is expressed by frehsly isolated epidermal Langerhans cells and is not regulated during culture.
J Invest Dermatol
105
1995
215
140
Lee
MG
Borkowski
TA
Udey
MC
Regulation of expression of B7 by murine Langerhans cells: A direct relationship between B7 mRNA levels and the level of surface expression of B7 by Langerhans cells.
J Invest Dermatol
101
1993
883
141
Norton
J
Sloane
JP
Al-Saffar
N
Haskard
DO
Expression of adhesion molecules in human intestinal graft-versus-host disease.
Clin Exp Immunol
87
1992
231
142
Le Varlet
B
Dezutter-Dambuyant
C
Staquet
MJ
Delorme
P
Schmitt
D
Human epidermal Langerhans Cells express integrins of the β 1 subfamily.
J Invest Dermatol
96
1991
518
143
Inaba
K
Steinman
RM
Monoclonal antibodies to LFA-1 and to CD4 inhibit the mixed leucocyte reaction after the antigen dependent clustering of dendritic cells and T lymphocytes.
J Exp Med
165
1987
1403
144
Shibaki
A
Meunier
L
Ra
C
Shimada
S
Ohkawara
A
Cooper
KD
Differential responsiveness of Langerhans cell subsets of varying phenotypic states in normal human epidermis.
J Invest Dermatol
104
1995
42
145
Borkowski
TA
Van Dyke
BJ
Schwarzenberger
K
McFarland
VW
Farr
AG
Udey
MC
Expression of E-cadherin by murine dendritic cells: E-cadherin as a dendritic cell differentiation antigen characteristic of epidermal Langerhans cells and related cells.
Eur J Immunol
24
1994
2767
146
Blauvelt
A
Katz
SI
Udey
MC
Human Langerhans cells express E-Cadherin.
J Invest Dermatol
104
1995
293
147
Rosen
SD
Bertozzi
CR
The selectins and their ligands.
Curr Opin Cell Biol
6
1994
663
148
Koszik
F
Strunk
D
Simonitsch
I
Picker
LJ
Stingl
G
Payer
E
Expressions of monoclonal antibody HECA-452-defined E-selectin ligands on Langerhans cells in normal and diseased skin.
J Invest Dermatol
102
1994
773
149
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 tumour necrosis factorα.
J Exp Med
179
1994
1109
150
Pearson
AM
Scavenger receptor in innate immunity.
Curr Opin Immunol
8
1996
20
151
Zhang
H
Shaw
ARE
Mak
A
Letarte
M
Endoglin is a component of the transforming growth factor (TGF )-β receptor complex of human pre-B leukemic cells.
J Immunol
156
1996
565
152
Greenwalt
DE
Lipsky
RH
Ochenhouse
CF
Ikeda
H
Membrane glycoprotein CD36: A review of its roles in adherence, signal transduction, and transfusion medicine.
Blood
80
1992
1105
153
Stingl
G
Wolff-Schreiner
EC
Pichler
WJ
Gschnait
F
Knapp
W
Wolff
K
Epidermal Langerhans cells bear Fc and C3 receptors.
Nature
268
1977
245
154
Morelli
A
Larregina
A
Chuluyan
E
Kolkowski
E
Fainboim
L
Expression and modulation of C5a receptor (CD88) on skin dendritic cells. Chemotactic effect of C5a on skin migratory dendritic cells.
Immunology
89
1996
126
155
Fanger
NA
Wardwell
K
Shen
L
Tedder
TF
Guyre
PM
Type I (CD64) and Type II (CD32) Fc-γ receptor-mediated phagocytosis by human blood dendritic cells.
J Immunol
157
1996
541
156
Reiger
A
Wang
B
Kilgus
O
Ochiai
K
Maurer
D
Fodinger
D
Kinet
J
Stingl
G
FcεRI mediates IgE binding to human epidermal Langerhans cells.
J Invest Dermatol
99
1992
30s
157
Bieber
T
Reiger
A
Neuchrist
C
Prinz
JC
Reiber
EP
Boltz-Nitulescu
G
Scheiner
O
Kraft
D
Ring
J
Stingl
G
Induction of FcεR2/CD23 on human epidermal Langerhans cells by human recombinant interleukin 4 and γ interferon.
J Exp Med
170
1989
309
158
Maurer
D
Fiebiger
E
Ebner
C
Reininger
B
Fischer
GF
Wichlas
S
Jouvin
M
Schmitt-Egenolf
M
Kraft
D
Kinet
J
Stingl
G
Peripheral blood dendritic cells express FcεRI as a complex composed of FcεRIα- and FcεRI-γ-chains and can use this receptor for IgE-mediated allergen presentation.
J Immunol
157
1996
607
159
Jackson
DG
Hart
DNJ
Starling
GC
Bell
JI
Molecular cloning of a novel member of the immunoglobulin gene superfamily homologous to the polymeric immunoglobulin receptor.
Eur J Immunol
22
1992
1157
160
Ezekowitz
RAB
Sastryh
K
Bailly
P
Warner
A
Molecular characterization of the human macrophage mannose receptor: Demonstration of multiple carbohydrate recognition-like domains and phagocytosis of yeasts in Cos-1 cells.
J Exp Med
172
1990
1785
161
Wu
K
Yuan
J
Lasky
LA
Characterization of a novel member of the macrophage mannose receptor type C lectin family.
J Biol Chem
271
1996
21323
162
Sallusto
F
Cella
M
Danieli
C
Lanzavecchia
A
Dendritic cells use macropinocytosis and the mannose receptor to concentrate macromolecules in the major histocompatibility complex class II compartment: Downregulation by cytokines and bacterial products.
J Exp Med
182
1995
389
163
Caux
C
Vanbervliet
B
Massacrier
C
Durand
I
Banchereau
J
Interleukin-3 cooperates with tumor necrosis factor α for the development of human dendritic/Langerhans cells from cord blood CD34+ hematopoietic progenitor cells.
Blood
87
1996
2376
164
Pugh
CW
McPherson
GG
Steer
HW
Characterization of nonlymphoid cells derived from rat peripheral lymph.
J Exp Med
172
1983
1459
165
Ingalls
RR
Golenbock
DT
CD11c/CD18, A transmembrane signalling receptor for lipopolysaccharide.
J Exp Med
181
1995
1473
166
Hughes
DA
Fraser
IP
Gordon
S
Murine macrophage scavenger receptor: In vivo expression and function as a receptor for macrophage adhesion in lymphoid and non-lymphoid organs.
Eur J Immunol
25
1995
466
167
Elomaa
O
Kangas
M
Sahlberg
C
Tuukkanen
J
Sormunen
R
Liakka
A
Thesleff
I
Kraal
G
Tryggvason
K
Cloning of a novel bacteria-binding receptor structurally related to scavenger receptors and expressed in a subset of macrophages.
Cell
80
1995
603
168
Epstein
J
Eichbaum
Q
Sheriff
S
Ezekowitz
RAB
The collectins in innate immunity.
Curr Opin Immunol
8
1996
29
169
Hart
DNJ
Starling
GC
Calder
VL
Fernando
NJ
B7/BB-1 is a leucocyte differentiation antigen on human dendritic cells induced by activation.
Immunology
79
1993
616
170
Symington
FW
Brady
W
Linsley
PS
Expression and function of B7 on human epidermal Langerhans cells.
J Immunol
150
1993
1286
171
Halliday
GM
Lucas
AD
Barnetson
RSC
Control of Langerhans cell density by a skin tumour-derived cytokine.
Immunology
77
1992
13
172
Enk
AH
Katz
SI
Heat stable antigen is an important costimulatory molecule on epidermal langerhans cells.
J Immunol
152
1994
3260
173
Lu
L
Woo
J
Li
Y
Watkins
SC
Qian
S
Starzl
TE
Demetris
AJ
Thomson
AW
Propagation of dendritic cell progenitors from normal mouse liver using granulocyte macrophage colony-stimulating factor and then maturational development in the presence of type-1 collagen.
J Exp Med
179
1994
1823
174
Crowley MT, Inaba K, Witmer-Pack M, Steinman RM: The cell surface of mouse dendritic cells: FACS analyses of dendritic cells from different tissues including thymus. Cell Immunol 108, 1989
175
Kay
R
Rosten
PM
Humphries
RK
CD24, a signal transducer molecule modulating B cell activation responses, is a very short peptide with a glycosyl phosphatidylinositol membrane anchor.
J Immunol
147
1991
1412
176
Williams
LA
Hock
BD
Hart
DNJ
Human T lymphocytes and haemopoietic cell lines express CD24 associated carbohydrate epitopes in the absence of CD24 mRNA or protein.
Blood
88
1996
3048
177
Williams
LA
McLellan
AD
Summers
KL
Sorg
RV
Fearnley
DB
Hart
DNJ
Identification of a novel dendritic cell surface antigen defined by carbohydrate specific CD24 antibody cross reactivity.
Immunology
89
1996
120
178
Daish
A
Hock
BD
Hart
DNJ
Characterization of a novel leucocyte activation antigen recognised by the antibody CMRF-37.
Immunol Cell Biol
72
1994
13
179
Gruss
H
DaSilva
N
Hu
Z
Uphoff
CC
Goodwin
RG
Drexler
HG
Expression and regulation of CD30 ligand and CD30 in human leukemia-lymphoma cell lines.
Leukemia
8
1994
2083
180
Ludewig
B
Graf
D
Gelderblom
HR
Becker
Y
Kroczek
RA
Pauli
G
Spontaneous apoptosis of dendritic cells is efficiently inhibited by TRAP (CD40-ligand) and TNF-α, but strongly enhanced by interleukin-10.
Eur J Immunol
25
1995
1943
181
Kampgen E, Gold R, Eggert A, Keikavoussi P, Grauer O, Toyka K, Broecker E: Steroids and UV-B are potent inducers of programmed cell death in immature and mature dendritic cells. J Cell Biochem 18B:12, 1994 (abstr, suppl)
182
Bender
A
Sapp
M
Schuler
G
Steinman
RM
Bhardwaj
N
Improved methods for the generation of dendritic cells from nonproliferating progenitors in human blood.
J Immunol Methods
196
1996
121
183
Suss
G
Shortman
K
A subclass of dendritic cells kills CD4 T cells via Fas/fas-ligand-induced apoptosis.
J Exp Med
183
1996
1789
184
Witmer-Pack
MD
Crowley
MT
Inaba
K
Macrophages, but not dendritic cells, accumulate colloidal carbon following administration in situ.
J Cell Sci
105
1993
965
185
Prickett
TCR
McKenzie
JL
Hart
DNJ
Characterization of human interstitial dendritic cells in liver.
Transplantation
46
1988
754
186
Betjes
MGH
Haks
MC
Tuk
CW
Beelen
HJ
Monoclonal antibody EBM11 (anti-CD68) discriminates between dendritic cells and macrophages after short-term culture.
Immunobiology
183
1991
79
187
Galkowska
H
Olsewski
WL
Immune events in skin. 1. Spontaneous cluster formation of dendritic (veiled) cells and lymphocytes from skin lymph.
Scand J Immunol
35
1992
727
188
Arkema
JMS
Schadee-Eestermans
IL
Beelen
RHJ
Hoefsmith
ECM
A combined method for both endogenous myeloperoxidase and acid phosphatase cytochemistry as well as immunoperoxidase surface labelling discriminating human peripheral blood-derived dendritic cells and monocytes.
Histochemistry
95
1991
573
189
Takahashi
Y
Yamaguchi
H
Ishizeki
J
Nakajima
T
Nakazota
Y
Immunohistochemical and immunoelectron microscopic localization of S-100 protein in the interdigitating reticulum cells of the human lymph node.
Virchows Arch (Cell Pathol)
37
1982
125
190
Troy AJ, Summers KL, Davidson PJT, Atkinson CA, Hart DNJ: The dendritic cell infiltrate in human renal cell carcinoma: A probable mechanism for tumor escape from immune surveillance. 1997 (submitted)
191
Mosialos
G
Birkenbach
M
Ayehunie
S
Matsumura
F
Pinkus
GS
Kieff
E
Langhoff
E
Circulating human dendritic cells differentially express high levels of a 55-kd actin-bundling protein.
Am J Pathol
148
1996
593
192
Bilbe
G
Delabie
J
Bruggen
J
Richener
H
Asselbergs
FA
Cerletti
N
Sorg
C
Odink
K
Tarcsay
L
Wiesendanger
W
de Wolf-Peeters
C
Shipman
R
A novel intermediate filament-associated protein highly expressed in the Reed-Sternberg cells of Hodgkin's disease.
EMBO J
11
1992
2103
193
Fridman
R
Lider
O
Naparstek
K
Fuks
Z
Vlodavsky
I
Cohen
IR
Soluble antigen induces T lymphocytes to secrete an endoglycosidase that degrades the heparan sulfate moiety of subendothelial extracellular matrix.
J Cell Physiol
130
1987
85
194
Carrasco
D
Ryseck
R
Bravo
R
Expression of rel B transcripts during lymphoid organ development: Specific expression in dendritic antigen-presenting cells.
Development
118
1993
1221
195
Granelli-Piperno
A
Pope
M
Inaba
K
Steinman
RM
Coexpression of NF-kB/Rel and Sp 1 transcription factors in human immunodeficiency virus 1-induced, dendritic cell-T-cell syncytia.
Proc Natl Acad Sci USA
92
1995
10944
196
Zhang
DE
Hetherington
CJ
Tan
S
Dziennis
SE
Gonzalez
DA
Chen
HM
Tenen
DG
Sp-1 is a critical factor for the monocytic specific expression of CD14.
J Biol Chem
269
1994
11425
197
Akagawa
KS
Takasuka
N
Nozaki
Y
Komuro
I
Azuma
M
Ueda
M
Naito
M
Takahasi
K
Generation of CD1+ RelB+ dendritic cells and tartrate-resistant acid phsophatase-positive osteoclast-like multinucleated giant cells from human monocytes.
Blood
88
1996
4029
198
Feuillard
J
Korner
M
Israel
A
Vassy
J
Raphael
M
Differential nuclear localization of p50, p52 and RelB proteins in human accessory cells of the immune response in situ.
Eur J Immunol
26
1996
2547
199
Boehmelt
G
Madruga
J
Dorfler
P
Briegel
K
Schwarz
H
Enrietto
PJ
Zenke
M
Dendritic cell progenitor is transformed by a conditional v-rel estrogen reception fusion protein v-rel ER.
Cell
80
1995
341
200
Weih
F
Carrasco
D
Durham
SK
Barton
DS
Rizzso
CA
Ryseck
R
Lira
SA
Bravo
R
Multiorgan inflammation and hematopoietic abnormalities in mice with a targeted disruption of RelB, a member of the NF-kB/Rel family.
Cell
80
1995
331
201
Burkly
L
Hession
C
Ogata
L
Reilly
C
Marconi
LA
Olson
D
Tizard
R
Cate
R
Lo
D
Expression of relB is required for the development of thymic medulla and dendritic cells.
Nature
373
1995
531
202
Mayordomo
JI
Zorina
T
Storkus
WJ
Zitvogel
L
Celluzzi
C
Falo
LD
Melief
CJ
Ildstad
ST
Kast
WM
DeLeo
AB
Lotze
MT
Bone marrow-derived dendritic cells pulsed with synthetic tumour peptides elicit protective and therapeutic antitumour immunity.
Nature Med
1
1995
1297
203
Celluzzi
CM
Mayordomo
JI
Storkus
WJ
Lotze
MT
Falo
LD
Peptide-pulsed dendritic cells induce antigen-specific, CLT-mediated protected tumor immunity.
J Exp Med
183
1996
283
204
Wiktor-Jedrzejczak
W
Ratajczak
MZ
Ptasznik
A
Sell
KW
Ahmed-Ansari
A
Ostertag
W
CSF-1 deficiency in the op/op mouse has differential effects on macrophage populations and differentiation stages.
Exp Hematol
20
1992
1004
205
Witmer-Pack
MD
Hughes
DA
Schuler
G
Lawson
L
McWilliam
A
Inaba
K
Steinman
RM
Gordon
S
Identification of macrophages and dendritic cells in the osteopetrotic (op/op) mouse.
J Cell Science
104
1993
1021
206
Takahashi
K
Naito
M
Schultz
LD
Hayashi
S
Nishikawa
S
Differentiation of dendritic cell populations in macrophage colony-stimulating factor-deficient mice homozygous for the osteopetrosis (op) mutations.
J Leuk Biol
53
1993
19
207
Gieseler
RKH
Rober
R
Kuhn
R
Weber
K
Osborn
M
Peters
JH
Dendritic accessory cells derived from rat bone marrow precursors under chemically defined conditions in vitro belong to the myeloid lineage.
J Cell Biol
54
1991
171
208
de Frassinette
A
Schmitt
D
Dezutter-Dambuyant
C
Guyotat
D
Zabot
MT
Thivolet
J
Culture of putative Langerhans cell bone marrow precursors: Characterisation of their phenotype.
Exp Haematol
16
1988
764
209
Reid
CDL
Stackpoole
Meager A
Tikerpae
J
Interactions of tumour necrosis factor with granulocyte-macrophage 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
149
1992
2681
210
Strunk
D
Egger
C
Leitner
G
Hanau
D
Stingl
G
A skin homing molecule defines the Langerhans cell progenitor in human peripheral blood.
J Exp Med
185
1996
1151
211
Saraya
K
Reid
CDL
Stem cell factor and the regulation of dendritic cell production from CD34(+) progenitors in bone marrow and cord blood.
Br J Haematol
93
1996
258
212
Young
JW
Szabolcs
P
Moore
MAS
Identification of dendritic cell colony-forming units among normal human CD34+ bone marrow progenitors that are expanded by c-kit ligand and yield pure dendritic cell colonies in the presence of granulocyte/macrophage colony-stimulating factor and tumor necrosis factorα.
J Exp Med
182
1995
1120
213
Szabolcs
P
Moore
MAS
Young
JW
Expansion of immunostimulatory dendritic cells among the myeloid progeny of human CD34+ bone marrow precursors cultured with c-kit ligand, granulocyte-macrophage colony-stimulating factor, and TNF-α.
J Immunol
154
1995
5851
214
Galy
A
Travis
M
Cen
D
Chen
B
Human T, B, natural killer and dendritic cells arise from a common bone marrow progenitor cell subset.
Immunity
3
1995
459
215
Res
P
Martinezcaceres
E
Jaleco
AC
Staal
F
Noteboom
E
Weijer
K
Spits
H
CD34(+) CD38 (dim) cells in the human thymus can differentiate into T, natural killer, and dendritic cells but are distinct from pluripotent stem cells.
Blood
87
1996
5196
216
Santiagoschwarz
F
Tucci
J
Carsons
SE
Endogenously produced interleukin 6 is an accessory cytokine for dendritic cell hematopoiesis.
Stem Cells
14
1996
225
217
Maraskovsky
E
Brasel
K
Teepe
M
Roux
ER
Lyman
SD
Shortman
K
McKenna
HJ
Dramatic increase in the numbers of functionally mature dendritic cells in Flt3 ligand-treated mice: Multiple dendritic cell subpopulations identified.
J Exp Med
184
1996
1953
218
Siena
S
Di Nicola
M
Bregni
M
Mortarini
R
Anichini
A
Lombardi
L
Ravagnani
F
Parmiani
G
Gianni
AM
Massive ex vivo generation of functional dendritic cells from mobilized CD34+ blood progenitors for anticancer therapy.
Exp Hematol
23
1995
1463
219
Vremec
D
Lieschke
GJ
Dunn
AR
Robb
L
Metcalf
D
Shortman
K
The influence of granulocyte/macrophage colony-stimulating factor on dendritic cell levels in mouse lymphoid organs.
Eur J Immunol
27
1997
40
220
Jansen
JH
Wientjens
GHM
Fibbe
WE
Willemze
R
Kluin-Nelemans
HC
Inhibition of human macrophage colony formation by interleukin 4.
J Exp Med
170
1989
577
221
Piemonti
L
Bernasconi
S
Luini
W
Trobonjaca
Z
Minty
A
Allavena
P
Mantovani
A
IL-13 supports differentiation of dendritic cells from circulating precursors in concert with GM-CSF.
Eur Cytokine Network
6
1995
245
222
Ziegler-Heitbrock
HWL
Fingerle
G
Strobel
M
Schraut
W
Stelter
F
Schutt
C
Passlick
B
Pforte
A
The novel subset of CD14+/CD16+ blood monocytes exhibits features of tissue macrophages.
Eur J Immunol
23
1993
2053
223
Peters
JH
Ruppert
J
Gieseler
RK
Najar
HM
Xu
H
Differentiation of human monocytes into CD14 negative accessory cells: Do dendritic cells derive from the monocyte lineage.
Pathobiology
59
1991
122
224
Andreesen
R
Brugger
W
Scheibenbogen
C
Kreutz
M
Leser
H
Rehm
A
Lohr
GW
Surface phenotype analysis of human monocyte to macrophage maturation.
J Leuk Biol
47
1990
490
225
Zhou
L
Tedder
TF
CD14+ blood monocytes can differentiate into functionally mature CD83+ dendritic cells.
Proc Natl Acad Sci USA
93
1996
2588
226
Tuft
SJ
McKenzie
JL
Hawkins
P
Hart
DNJ
The localization of HLA antigens in human cornea.
Transactions NZ Opthmalmology Soc
36
1984
36
227
Kaplan
G
Walsh
G
Guido
LS
Meyn
P
Burkhardt
RA
Abalos
RM
Barker
J
Frindt
PA
Fajardo
TT
Celona
R
Cohn
ZA
Novel responses of human skin to intradermal recombinant granulocyte/macrophage colony-stimulating factor: Langerhans cell recruitment, keratinocyte growth and enhanced wound healing.
J Exp Med
175
1992
1717
228
Sozzani
S
Sallusto
F
Luini
W
Zhou
D
Piemonti
L
Allavena
P
Van Damme
J
Valitutti
S
Lanzavecchia
A
Mantovani
A
Migration of dendritic cells in response to formyl peptides, C5a, and a distinct set of chemokines.
J Immunol
155
1995
3292
229
Matzinger
P
Tolerance, danger and the extended family.
Annu Rev Immunol
12
1994
991
230
Roake
JA
Rao
AS
Morris
PJ
Larsen
CP
Hankins
DF
Austyn
JM
Dendritic cell loss from nonlymphoid tissues after systemic administration of lipopolysaccharide, tumor necrosis factor and interleukin 1.
J Exp Med
181
1995
2237
231
Larsen
CP
Steinman
RM
Witmer-Pack
M
Hankins
DF
Morris
PJ
Austyn
JM
Migration and maturation of Langerhans cells in skin transplants and explants.
J Exp Med
172
1990
1483
232
Hussain
LA
Lehner
T
Comparative investigations of Langerhans' cells and potential receptors for HIV in oral, genitourinary and rectal epithelia.
Immunology
85
1995
475
233
Czernieleswski
J
Vaigot
P
Prunieras
M
Epidermal Langerhans cells — A cycling cell population.
J Invest Derm
84
1985
424
234
Cohen
PJ
Katz
SI
Cultured human Langerhans cells process and present intact protein antigens.
J Invest Derm
99
1992
331
235
Morris
J
Alaibac
M
Jia
M
Chu
T
Purification of functional active epidermal Langerhans cells: A simple and efficient new technique.
J Invest Derm
99
1992
237
236
Hanau
D
Schmitt
DA
Fabre
MF
Cazenave
JP
A method for the rapid isolation of human epidermal Langerhans cells using immunomagnetic microspheres.
J Invest Derm
91
1988
274
237
Morhenn VB, Wood GS, Engleman EG, Oseroff AR: Selective enrichment of human epidermal cell subpopulations using monoclonal antibodies. J Invest Derm 81:127s, 1983 (suppl)
238
Troy A, Davidson P, Atkinson C, Hart D: Phenotypic characterization of the dendritic cell infiltrate in prostate cancer. 1997 (submitted)
239
Spencer
SC
Fabre
JW
Characterization of the tissue macrophage and the interstitial dendritic cell as distinct leukocytes normally resident in the connective tissue of rat heart.
J Exp Med
171
1990
1841
240
Williams
KA
Coster
DJ
The role of the limbus in corneal allograft rejection.
Eye
3
1989
158
241
Steptoe
RJ
Holt
PG
McMenamin
PG
Functional studies of major histocompatibility class II-positive dendritic cells and resident tissue macrophages isolated from the rat iris.
Immunology
85
1995
630
242
Sato
T
Kaneko
M
Hama
A
Kusakari
T
Fujieda
H
Expression of class II MHC molecules in the rat pineal gland during development and effects of treatment with carbon tetrachloride.
Cell Tissue Res
284
1996
65
243
Barrett
AW
Ross
DA
Goodacre
JA
Purified human oral mucosal Langerhans cells function as accessory cells in vitro.
Clin Exp Immunol
92
1993
158
244
Pavli
P
Hume
DA
Van De Pol
E
Doe
WF
Dendritic cells, the major antigen-presenting cells of the human colonic lamina propia.
Immunology
78
1993
132
245
Holt
PG
Schon-Hegrad
MA
Oliver
J
MHC class II antigen-bearing dendritic cells in pulmonary tissues of the rat.
J Exp Med
167
1988
262
246
Schon-Hegrad
MA
Oliver
J
McMenamin
PG
Holt
PG
Studies on the density, distribution and surface phenotype of intraepithelial class II major histocompatibility complex antigen (Ia)-bearing dendritic cells (DC) in the conducting airways.
J Exp Med
173
1991
1345
247
Nelson
DJ
Holt
PG
Defective regional immunity in the respiratory tract of neonates is attributable to hyporesponsiveness of local dendritic cells to activation signals.
J Immunol
155
1995
3517
248
McWilliam
AS
Nelson
D
Thomas
JA
Holt
PG
Rapid dendritic cell recruitment is a hallmark of the acute inflammatory response at mucosal surfaces.
J Exp Med
179
1994
1331
249
Pavli
P
Woodhams
CE
Doe
WF
Hume
DA
Isolation and characterization of antigen-presenting dendritic cells from the moust intestinal lamina propria.
Immunology
70
1990
40
250
Nicod
LP
Lipscomb
MF
Weissler
JC
Lyons
R
Albertson
J
Toews
GB
Mononuclear cells in human lung parenchyma.
Am Rev Respir Dis
136
1987
818
251
Holt
PG
Degebrodt
A
O'Leary
C
Krska
K
Plozza
T
T cell activation by antigen-presenting cells from lung tissue digests: Suppression by endogenous macrophages.
Clin Exp Immunol
62
1985
586
252
Bujdoso
R
Hopkins
J
Dutia
BM
Young
P
McConnell
I
Characterization of sheep afferent lymph dendritic cells and their role in antigen carriage.
J Exp Med
170
1989
1285
253
Kripke
ML
Munn
CG
Jeevan
A
Tang
J
Bucana
C
Evidence that cutaneous APC migrate to regional lymph nodes during contact sensitisation.
J Immunol
145
1990
2833
254
Pistoor
FHM
Rambukkana
A
Kroezen
M
Lepoittevin
JP
Bos
JD
Kapsenberg
ML
Das
PK
Novel predictive assay for contact allergens using human skin explant cultures.
Am J Pathol
149
1996
337
255
Cumberbatch
M
Kimber
I
Dermal TNFα induces dendritic cell migration to draining lymph nodes, and possibly provides one stimulus for Langerhans cell migration.
Immunology
75
1992
257
256
Spry
CJF
Pflug
AJ
Janossy
G
Humphrey
JH
Large mononuclear (veiled) cells with “Ia-like” membrane antigens in human afferent lymph.
Clin Exp Immunol
39
1980
750
257
Fossum
S
Lymph-borne dendritic leucocytes do not recirculate but enter the lymph node paracortex to becoming interdigitating cells.
Scand J Immunol
27
1988
97
258
Larsen
CP
Barker
H
Morris
PJ
Austyn
JM
Failure of mature dendritic cells of the host to migrate from the blood into cardiac or skin allografts.
Transplantation
50
1990
294
259
Hopkins
J
Dutia
B
Bujdoso
R
McConnell
I
In vivo modulation of CD1 and MHC class II expression by sheep afferent lymph dendritic cells.
J Exp Med
170
1989
1303
260
Sokolowski
J
Jakobsen
E
Johanneseen
JV
Cells in peripheral leg lymph of normal men.
Lymphology
11
1978
202
261
Macatonia
SE
Edwards
AJ
Knight
SC
Dendritic cells and the inhibition of contact sensitivity to fluorescein isothiocyanate.
Immunology
59
1986
509
262
Larsen
CP
Morris
PJ
Austyn
JM
Migration of dendritic leukocytes from cardiac allografts into host spleens.
J Exp Med
171
1990
307
263
Gyure
LA
Barfoot
R
Denham
S
Hall
JG
Immunity to a syngeneic sarcoma induced in rats by dendritic lymph cells exposed to the tumour either in vivo or in vitro.
Br J Cancer
55
1987
17
264
MacPherson
GG
Jenkins
CD
Stein
MJ
Edwards
C
Endotoxin-mediated dendritic cell release from the intestine: characterisation of released dendritic cells and TNF dependence.
J Immunol
159
1995
1317
265
Rhodes
JM
Balfour
BM
Blom
J
Agger
R
Comparison of antigen uptake by peritoneal macrophages and veiled cells from thoracic duct using isotope-, FITC-, or gold labelled antigen.
Immunology
68
1989
403
266
Veerman
AJP
Van Ewijk
W
White pulp compartments in the spleen of rats and mice. A light and EM study of lymphoid and non-lymphoid cells in T- and B-areas.
Cell Tissue Res
156
1975
417
267
Wood
GS
Turner
RR
Shiurba
RA
Eng
L
Warnke
RA
Human dendritic cells and macrophages. In-situ immunophenotypic definition of subsets that exhibit specific morphologic and microenvironmental characteristics.
Am J Pathol
119
1985
73
268
McKenzie
JL
Prickett
TCR
Hart
DNJ
Human dendritic cells stimulate allogeneic T cells in the absence of IL-1.
Immunology
67
1989
290
269
King
PD
Katz
DR
Human tonsillar dendritic cell-induced T cell responses: analysis of molecular mechanisms using monoclonal antibodies.
Eur J Immunol
19
1989
581
270
Vythdreese
FA
Dellemijn
TAM
Majoor
D
Dejong
D
Localization in situ of the co-stimulatory molecules B7.1, B7.2, CD40 and their ligands in normal human lymphoid tissue.
Eur J Immunol
25
1995
3023
271
Grouard
G
Durand
I
Filgueira
L
Banchereau
J
Liu
Y
Dendritic cells capable of stimulating T cells in germinal centres.
Nature
384
1996
364
272
De Smedt
T
Pajak
B
Muraille
E
Lespagnard
L
Heinen
E
De Baetselier
P
Urbain
J
Leo
O
Moser
M
Regulation of dendritic cell numbers and maturation by lipopolysaccharide in vivo.
J Exp Med
184
1996
1413
273
Lu
L
Hsieh
M
Oriss
TB
Morel
PA
Starzl
TE
Rao
AS
Thomson
AW
Generation of DC from mouse spleen cell cultures in response to GM-CSF: immunophenotypic and functional analyses.
Immunology
84
1995
127
274
Hosmalin A, McIlroy D, Cheynier R, Clauvel JP, Oksenhendler E, Wain-Hobson S, Debre P, Autran B: Splenic interdigitating dendritic cells in humans: characterization and HIV infection frequency in vivo. 2:439, 1995 (abstr)
275
Flechner
ER
Freudenthal
PS
Kaplan
G
Steinman
RM
Antigen specific T lymphocytes efficiently cluster with dendritic cells in the human primary mixed leucocyte reaction.
Cell Immunol
111
1988
183
276
Bender
A
Bui
LK
Feldman
MAV
Larsson
M
Bhardwaj
N
Inactivated influenza virus, when presented on dendritic cells, elicits human CD8+ cytolytic T cell responses.
J Exp Med
182
1995
1663
277
Inaba
K
Young
JW
Steinman
RM
Direct activation of CD8+ cytotoxic T lymphocytes by dendritic cells.
J Exp Med
166
1987
182
278
Boog
CJP
Boes
J
Melief
CJM
Stimulation with dendritic cells decreases or obviates the CD4+ helper cell requirement with cytotoxic T lymphocyte responses.
Eur J Immunol
18
1988
219
279
Young
JW
Steinman
RM
Dendritic cells stimulate primary human cytolytic lymphocytes in the absence of CD4+ helper T cells.
J Exp Med
171
1990
1315
280
Shah
PD
Gilbertson
SM
Rowley
DA
Dendritic cells that have interacted with antigen are targets for natural killer cells.
J Exp Med
162
1985
625
281
Faassen
AE
Pierce
SK
Cross-linking cell surface class II molecules stimulates Ig-mediated B cell antigen processing.
J Immunol
155
1995
1737
282
Botham
PA
Rattray
NJ
Walsh
ST
Riley
EJ
Control of the immune response to contact sensitizing chemicals by cutaneous antigen-presenting cells.
Br J Dermatol
117
1987
1
283
Hauser
C
Cultured epidermal Langerhans cells activate effector cells for contact sensitivity.
J Invest Dermatol
95
1990
436
284
Inaba
K
Inaba
M
Naito
M
Steinman
RM
Dendritic cell progenitors phagocytose particulates, including bacillus Calmette-Geurin organisms and sensitize mice to Mycobacterial antigens in vivo.
J Exp Med
178
1993
479
285
Baird
MA
Hart
DNJ
Abernethy
N
Watson
J
Dendritic cell presentation of PPD and 19kDa protein of Mycobacterium tuberculosis and emergent T helper cell phenotype.
Immunol Cell Biol
73
1995
537
286
Nair
S
Babu
JS
Dunham
RG
Kanda
P
Burke
RL
Rouse
BT
Induction of primary, antiviral cytotoxic and proliferative responses with antigens administered via dendritic cells.
J Virol
67
1993
4062
287
Macatonia
SE
Doherty
TM
Knight
SC
O'Garra
A
Differential effect of IL-10 on dendritic cell-induced T cell proliferation and IFN-γ production.
J Immunol
150
1993
3755
288
Hengel
H
Lindner
M
Wagner
H
Heeg
K
Frequency of herpes simplex virus-specific murine cytotoxic T lymphocyte precursors in mitogen- and antigen-driven primary in vitro T cell responses.
J Immunol
139
1987
4196
289
Kast
WM
Boog
CJP
Roep
BO
Voordouw
AC
Melief
CJM
Failure or success in the restoration of virus-specific cytotoxic T lymphocyte response defects by dendritic cells.
J Immunol
140
1988
3186
290
Nonacs
R
Humborg
C
Tam
JP
Steinman
RM
Mechanisms of mouse spleen dendritic cell function in the generation of influenza-specific, cytolytic T lymphocytes.
J Exp Med
176
1992
519
291
Sornasse
T
Flamand
V
De Becker
G
Bazin
H
Tielemans
K
Urbain
J
Leo
O
Moser
M
Antigen-pulsed dendritic cells can efficiently induce an antibody response in-vivo.
J Exp Med
175
1992
15
292
Fagnoni
FF
Takamizawa
WR
Godfrey
WR
Rivas
A
Azuma
M
Okumura
K
Engelman
EG
Role of B70/B7-2 in CD4+ T-cell immune responses induced by dendritic cells.
Immunology
85
1995
467
293
Kaye
PM
Chain
BM
Feldmann
M
Non-phagocytic dendritic cells are effective accessory cells for anti-mycobacterial responses in vitro.
J Immunol
134
1985
1930
294
McKenzie JL: Dendritic cells: Phenotype and function. 1989. PhD thesis, University of Otago
295
Pancholi
P
Steinman
RM
Bhardwaj
N
Dendritic cells efficiently immunoselect mycobacterial-reactive T cells in human blood, including clonable antigen-reactive precursors.
Immunology
76
1992
217
296
Waalen
K
Forre
O
Teigland
J
Natvig
B
Human rheumatoid synovial and normal blood dendritic cells as antigen presenting cells — comparison with autologous monocytes.
Clin Exp Immunol
70
1987
1
297
Macatonia
SE
Patterson
S
Knight
SC
Primary proliferative and cytotoxic T-cell responses to HIV induced in vitro by human dendritic cells.
Immunology
74
1991
399
298
Takahashi
H
Nagagawa
Y
Yokomuro
K
Berzofsky
JA
Induction of CD8+ cytotoxic T lymphocytes by immunization with syngeneic irradiated HIV-1 envelope derived peptide-pulsed dendritic cells.
Int Immunol
5
1993
849
299
Bhardwaj
N
Friedman
SM
Cole
BC
Nisanian
AJ
Dendritic cells are potent antigen-presenting cells for microbial superantigens.
J Exp Med
175
1992
267
300
Pelletier
N
Tautu
C
Landry
D
Montplaisir
S
Chertrand
C
Perreault
C
Characterisation of human thymic dendritic cells in culture.
Immunology
58
1986
263
301
Barclay
AN
Mayrhofer
CJ
Bone marrow origin of Ia positive cells in the medulla of rat thymus.
J Exp Med
153
1981
1666
302
Carlow
DA
van Oers
NSC
Teh
S
Teh
H
Deletion of antigen specific immature thymocytes by dendritic cells requires LFA-1/ICAM interactions.
J Immunol
148
1992
1595
303
Jenkinson
EJ
Anderson
G
Owen
JJT
Studies on T cell maturation on defined thymic stromal cell populations in vitro.
J Exp Med
176
1992
845
304
Saunders
D
Lucas
K
Ismaili
J
Wu
L
Maraskovsky
E
Dunn
A
Shortman
K
Dendritic cell development in culture from thymic precursor cells in the absence of granulocyte/macrophage colony-stimulating factor.
J Exp Med
184
1996
2185
305
Landry
D
La Fontaine
M
Cossette
M
Barthelemy
H
Chartrand
C
Human thymic dendritic cells: characterization, isolation and functional assays.
Immunology
65
1988
135
306
La Fontaine
M
Landry
D
Montplaisir
S
The human thymic dendritic cell phenotype and its modification in culture.
Cell Immunol
142
1992
238
307
Finkelman
FD
Lees
A
Birnbaum
R
Gause
WC
Morris
SC
Dendritic cells can present antigen in vivo in a tolerogenic or immunogenic fashion.
J Immunol
157
1996
1406
308
Thomson
AW
Lu
L
Subbotin
VM
Li
S
Qian
S
Rao
AS
Fung
JJ
Starzl
TE
In vitro propagation and homing of liver-derived dendritic cell progenitors to lymphoid tissues of allogeneic recipients.
Transplantation
59
1995
544
309
Lu
L
McCaslin
D
Starzl
TE
Mouse bone marrow-derived dendritic cell progenitors (NLDC145+, MHC class II+,B7-1dim,B7-2-) induce alloantigen-specific hyporesponsiveness in murine T lymphocytes.
Transplantation
60
1995
1539
310
Salomon
B
Lores
P
Pioche
C
Racz
P
Jami
J
Klatzmann
D
Conditional ablation of dendritic cells in transgenic mice.
J Immunol
152
1994
537
311
Gerondakis
S
Strasser
A
Metcalf
D
Grigoriadis
G
Scheerlinck
JY
Grumont
RJ
Rel-deficient T cells exhibit defects in production of interleukin 3 and granulocyte-macrophage colony-stimulating factor.
Proc Natl Acad Sci USA
93
1996
3405
312
Crowley
M
Inaba
K
Steinman
RM
Dendritic cells are the principal cells in mouse spleen bearing immunogenic fragments of foreign proteins.
J Exp Med
172
1990
383
313
Fossum
S
Rolstad
B
The roles of interdigitating cells and natural killer cells in the rapid rejection of allogeneic lymphocytes.
Eur J Immunol
16
1986
440
314
Crowley
MT
Inaba
K
Witmer-Pack
MD
Gezelter
S
Steinman
RM
Use of the fluorescence-activated cell sorter to enrich dendritic cells from mouse spleen.
J Immunol Methods
133
1990
55
315
Scheicher
C
Mehlig
M
Dienes
H
Reske
K
Uptake of microparticle-adsorbed protein antigen by bone marrow-derived dendritic cells results in up-regulation of interleukin-1α and interleukin-12 p40/p35 and triggers prolonged, efficient antigen presentation.
Eur J Immunol
25
1995
1566
316
Hart
DNJ
McKenzie
JL
Interstitial dendritic cells.
Int Rev Immunol
6
1990
127
317
Reis e Sousa
C
Stahl
PD
Austyn
JM
Phagocytosis of antigens by Langerhans cells in vitro.
J Exp Med
178
1993
509
318
Moll
H
Fuchs
H
Blank
C
Rollinghoff
M
Langerhans cells transport of Leishmania major from the infected skin to the draining lymph node for presentation to antigen-specific T cells.
Eur J Immunol
23
1993
1595
319
Pancholi
P
Mirza
A
Schauf
V
Steinman
RM
Bhardwaj
N
Presentation of mycobacterial antigens by human dendritic cells: Lack of transfer from infected macrophages.
Infect Immun
61
1993
5326
320
Takahashi
S
Hashimoto
K
Derivation of Langerhans cell granules from cytomembrane.
J Invest Dermatol
84
1985
469
321
Martinez-Pomares
L
Kosco-Vilbois
M
Darley
E
Tree
P
Herren
S
Bonnefoy
J
Gordon
S
Fc chimeric protein containing the cysteine-rich domain of the murine mannose receptor binds to macrophages from splenic marginal zone and lymph node subcapsular sinus and to germinal centers.
J Exp Med
184
1996
1927
322
Harkiss
GD
Hopkins
J
McConnell
I
Uptake of antigen by afferent lymph dendritic cells mediated by antibody.
Eur J Immunol
20
1990
2367
323
Hanau
D
Fabre
M
Schmitt
DM
Garaud
JC
Pauly
G
Tongio
MM
Mayer
S
Casenave
JP
Human epidermal Langerhans cells cointernalize by receptor mediated endocytosis “nonclassical” MHC class I molecules (T6 antigen) and class II (HLA-DR antigens).
Proc Natl Acad Sci USA
84
1987
2901
324
Romani
N
Koide
S
Crowley
M
Witmer-Pack
M
Livingstone
AM
Fantham
CG
Inaba
K
Steinman
RM
Presentation of exogenous protein antigens by dendritic cells to T cell clones. Intact protein is presented best by immature, epidermal Langerhans cells.
J Exp Med
169
1989
1169
325
Roche
A
Midoux
P
Bouchard
P
Monsigny
M
Membrane lectins on human monocytes. Maturation-dependent modulation of 6-phosphomannose and mannose receptors.
FEBS Lett
193
1985
63
326
Resnick
D
Pearson
A
Krieger
M
The SRCR superfamily: A family reminiscent of the Ig superfamily.
Trends Biochem Sci
19
1994
5
327
Sugita
M
Jackman
RM
van Donselaar
E
Behar
SM
Rogers
RA
Peters
PJ
Brenner
MB
Porcelli
SA
Cytoplasmic tail-dependent localization of CD1b antigen-presenting molecules to MIICs.
Science
273
1996
349
328
Beckman
EM
Porcelli
SA
Morita
CT
Behar
SM
Furlong
ST
Brenner
MB
Recognition of a lipid antigen by CD1-restricted αβ+ T cells.
Nature
372
1994
691
329
Stossel
H
Koch
F
Kampgen
E
Stogen
P
Lenz
A
Heufler
C
Romani
N
Schuler
G
Disappearance of certain organelles (endosome and Langerhans cell granules) accompanies loss of antigen processing capacity upon culture of epidermal Langerhans cells.
J Exp Med
172
1989
1471
330
Girolomoni
G
Cruz
PD
Bergstresser
PR
Internalization and acidification of surface HLA-DR molecules by epidermal Langerhans cells: A paradigm for antigen processing.
J Invest Dermatol
94
1990
753
331
Peguet-Navarro J, Dalbiez-Gauthier C, Schmidt D: Accessory function of human Langerhans cells in the primary allogeneic T-cell response. J Invest Dermatol 99:87s, 1992 (suppl)
332
Lanzavecchia
A
Mechanisms of antigen uptake for presentation.
Curr Opin Immunol
8
1996
348
333
Norbury
CC
Chambers
BJ
Prescott
AR
Ljunggren
H
Watts
C
Constitutive macropinocytosis allows TAP-dependent major histocompatibility complex class I presentation of exogenous soluble antigen by bone marrow-derived dendritic cells.
Eur J Immunol
27
1997
280
334
Levine
TP
Chain
BM
Endocytosis by antigen presenting cells: Dendritic cells are as endocytically active as other antigen presenting cells.
Proc Natl Acad Sci USA
89
1992
8342
335
Chain
BM
Kay
PM
Feldmann
M
The cellular pathway of antigen presentation: Biochemical and functional analysis of antigen processing in dendritic cells and macrophages.
Immunology
58
1986
271
336
Aiba
S
Katz
SI
The ability of cultured Langerhans cells to process and present protein antigens is MHC-dependent.
J Immunol
46
1991
2479
337
deBruijn
LH
Nieland
JD
Harding
JV
Melief
CJM
Processing and presentation of intact hen egg-white lysozyme by dendritic cells.
Eur J Immunol
22
1992
2347
338
Guery
J
Adorini
L
Dendritic cells are the most efficient in presenting endogenous naturally processed self-epitopes to class II-restricted T cells.
J Immunol
154
1995
536
339
Macatonia
SE
Hosken
NA
Litton
M
Vieira
P
Hsieh
C
Culpepper
JA
Wysocka
M
Trinchieri
G
Murphy
KM
O'Garra
A
Dendritic cells produce IL-12 and direct the development of Th1 cells from naive CD4+ T cells.
J Immunol
154
1995
5071
340
van Elsas
A
van der Burg
SH
van der Minne
CE
Borghi
M
Mourer
JS
Melief
CJM
Schrier
PI
Peptide-pulsed dendritic cells induce tumoricidal cytotoxic T lymphocytes from healthy donors against stably HLA-A*0201-binding peptides from the Melan-A/MART-1 self antigen.
Eur J Immunol
26
1996
1683
341
Mannering SI, McKenzie JL, Fearnley DB, Hart DNJ: HLA-DR beta*0101 presents bcr-abl (b3a2) peptide to CD4+ T lymphocytes: Optimal presentation by blood dendritic cells. Blood 1997 (in press)
342
Falo
LD
Colarusso
LJ
Benacerraf
B
Rock
KL
Serum proteases alter the antigenicity of peptides presented by class I major histocompatibility complex molecules.
Proc Natl Acad Sci USA
89
1992
8347
343
Girolomoni
G
Simon
JC
Bergstresser
PR
Cruz
PD
Freshly isolated spleen dendritic cells and epidermal Langerhans cells undergo similar phenotypic and functional changes during short term culture.
J Immunol
145
1990
2820
344
Simon JC, Thiele DL, Schopf E, Sontheimer RD: Effects of the immunosupressive dipeptide L-leucyl-L-leucine O-methyl ester on epidermal Langerhans cells. J Invest Dermatol 99:80s, 1992 (suppl)
345
Harding
CV
Roof
RW
Unanue
ER
Turnover of Ia-peptide complexes is facilitated in viable antigen-presenting cells: Biosynthetic turnover of Ia vs peptide exchange.
Proc Natl Acad Sci USA
86
1990
4230
346
Boog
CJP
Neefjes
JJ
Boes
J
Ploegh
HL
Melief
CJM
Specific immune responses restored by alteration in carbohydrate chains of surface molecules on antigen-presenting cells.
Eur J Immunol
19
1989
537
347
Porcelli
S
Morita
CT
Brenner
MB
CD1b restricts the response of human CD4−8− T lymphocytes to a microbial antigen.
Nature
360
1992
593
348
Porcelli
S
Brenner
MB
Greenstein
JL
Balk
SP
Terhorst
C
Bleicher
PA
Recognition of cluster of differentiation 1 antigens by human CD4− CD8+ cytolytic T lymphocytes.
Nature
341
1989
447
349
Faure
F
Jitsukawa
S
Miossec
C
Hercend
T
CD1c as a target recognition structure for human T lymphocytes: Analysis with peripheral blood γ/δ cells.
Eur J Immunol
20
1990
703
350
Scheeren
RA
Koopman
G
Van der Baan
S
Meijer
CJLM
Pals
ST
Adhesion receptors involved in clustering of blood dendritic cells and T lymphocytes.
Eur J Immunol
21
1991
1101
351
Schwartz
RH
A cell culture model for T lymphocyte clonal anergy.
Science
248
1990
1349
352
Jenkins
MK
The ups and downs of T cell costimulation.
Immunity
1
1994
443
353
Tan
P
Anasetti
C
Hansen
JA
Melrose
J
Brunvand
M
Bradshaw
J
Ledbetter
JA
Linsley
PS
Induction of alloantigen-specific hyporesponsiveness in human T lymphocytes by blocking interaction of CD28 with its natural ligand B7/BB1.
J Exp Med
177
1993
165
354
Guinan
EC
Gribben
JG
Boussiotis
VA
Freeman
GJ
Nadler
LM
Pivotal role of the B7:CD28 pathway in transplantation tolerance and tumor immunity.
Blood
84
1994
3261
355
Gribben
JG
Freeman
CJ
Boussiotis
VA
Rennert
P
Jellis
CL
Greenfield
E
Barber
M
Restivo
VAJ
Ke
X
Gray
GS
CTLA-4 mediates antigen-specific apoptosis of human T cells.
Proc Natl Acad Sci USA
92
1995
811
356
Kuchroo
VK
Das
MP
Brown
JA
Ranger
AM
Zamvii
SS
Sobel
RA
Weiner
HL
Nabavi
N
Glimcher
LH
B7.1 and B7.2 costimulatory molecules activate differentially the Th1/Th2 developmental pathways: Application to autoimmune disease therapy.
Cell
80
1995
707
357
Weiss
JM
Renkl
AC
Denfeld
RA
de Roche
R
Spitzlei
M
Schopf
E
Simon
JC
Low-dose UVB radiation perturbs the functional expression of B7.1 and B7.2 co-stimulatory molecules on human Langerhans cells.
Eur J Immunol
25
1995
2858
358
Larsen
CP
Ritchie
SC
Hendrix
R
Linsley
PS
Hathcock
KS
Hodes
RJ
Lowry
RP
Pearson
TC
Regulation of immunostimulatory function and costimulatory molecule (B7-1 and B7-2) expression on murine dendritic cells.
J Immunol
152
1994
5208
359
Young
JW
Koulova
L
Soergel
SA
Clarke
EA
Steinman
RM
Dupont
B
The B7/BB-1 antigen provides one of several co-stimulatory signals for the activation of CD4+ T lymphocytes by human blood DC in-vitro.
J Clin Invest
90
1992
229
360
Caux
C
Vanbervliet
B
Massacrier
C
Azuma
M
Okumura
K
Lanier
LL
Banchereau
J
B70/7.2 is identical to CD86 and is the major functional ligand for CD28 expressed on human dendritic cells.
J Exp Med
180
1994
1841
361
Girolomoni
G
Zambruno
G
Manfredini
R
Zacchia
V
Ferrari
S
Cossarizza
A
Giannetti
A
Expression of B7 costimulatory molecule in cultured human epidermal Langerhans cells is regulated at the mRNA level.
J Invest Dermatol
103
1994
54
362
Inaba
K
Witmer-Pack
M
Inaba
M
Hathcock
KS
Sakuta
H
Azuma
M
Yagita
H
Okumura
K
Linsley
PS
Ikehara
S
Muramatsu
S
Hodes
RJ
Steinman
RM
The tissue distribution of the B7-2 costimulator in mice: Abundant expression on dendritic cells in situ and during maturation in vitro.
J Exp Med
180
1994
1849
363
Kawamura
T
Furue
M
Comparative analysis of B7.1 and B7.2 expression in Langerhans cells: Differential regulation by T helper type 1 and T helper type 2 cytokines.
Eur J Immunol
25
1995
1913
364
Mitra
RS
Judge
TA
Nestle
FO
Turka
LA
Nickoloff
BJ
Psoriatic skin-derived dendritic cell function is inhibited by exogenous IL-10.
J Immunol
154
1995
2668
365
Fanslow
WC
Srinivasan
S
Paxton
R
Gibson
MG
Spriggs
MK
Armitage
RJ
Structural characteristics of CD40 ligand that determine biological function.
Semin Immunol
6
1994
267
366
Caux
C
Massacier
C
Vanbervliet
B
Dubois
B
Van Cooten
C
Durand
I
Banchereau
J
Activation of human dendritic cells through CD40 cross-linking.
J Exp Med
180
1994
1263
367
Lui
Y
Jones
B
Aruffo
A
Sullivan
KM
Linsley
PS
Janeway
CAJ
Heat stable antigen is a costimulatory molecule for T cell growth.
J Exp Med
175
1992
437
368
Lui
Y
Jones
B
Brady
W
Janeway
CAJ
Linsley
PS
Costimulation of murine CD4 T cell growth: Co-operation between B7 and heat-stable antigen.
Eur J Immunol
22
1992
2855
369
Williams LA: Molecular analysis of CD24 epitope-bearing antigens in human peripheral blood leukocyte populations, including dendritic cells. 1996. PhD thesis, University of Otago
370
Latza
U
Durkop
H
Schnittger
S
Ringeling
J
Eitelbach
F
Hummel
M
Fonatsch
C
Stein
H
The human OX40 homolog: cDNA structure, expression and chromosomal assignment of the ACT35 antigen.
Eur J Immunol
24
1994
677
371
Calderhead
DM
Buhlmann
JE
van den Eertwegh
AJ
Claasen
E
Noelle
RJ
Fell
HP
Cloning of mouse OX40: A T cell activation marker that may mediate T-B cell interactions.
J Immunol
151
1993
5261
372
Godfrey
WR
Fagnoni
FF
Harara
MA
Buck
D
Engleman
EG
Identification of a human OX-40 ligand, a costimulator of CD4+ T cells with homology to tumor necrosis factor.
J Exp Med
180
1994
757
373
Stuber
E
Neurath
M
Calderhead
D
Fell
HP
Strober
W
Cross-linking of OX40 ligand, a member of the TNF/NGF cytokine family, induces proliferation and differentiation in murine spleen cells.
Immunity
2
1996
507
374
Schwarz
H
Valbracht
J
Tuckwell
J
von Kempis
J
Lotz
M
ILA, the human 4-1BB homologue, is inducible in lymphoid and other cell lineages.
Blood
85
1995
1043
375
Pollok
KE
Kim
YI
Hurtado
J
Zhou
Z
Kim
KK
Kwon
BS
4-1BB T-cell antigen binds to mature B cells and macrophages, and costimulates anti-mu-primed B cells.
Eur J Immunol
24
1994
367
376
Pollok
KE
Kim
YJ
Zhou
Z
Hurtado
J
Kim
KK
Pickard
RT
Kwon
BS
Inducible T cell antigen 4-1BB. Analysis of expression and function.
J Immunol
150
1993
771
377
Garni-Wagner
B
Lee
ZH
Kim
Y
Wilde
C
Kang
C
Kwon
BS
4-1BB is expressed on CD45RAhiROhi transitional T cell in humans.
Cell Immunol
169
1996
91
378
Goodwin
RG
Din
WS
Davis-Smith
T
Anderson
DM
Gimpel
SD
Sato
TA
Maliszewski
CR
Brannan
CI
Copeland
NG
Jenkins
NA
Molecular cloning of a ligand for the inducible T cell gene 4-1BB: A member of the emerging family of cytokines with homology to tumor necrosis factor.
Eur J Immunol
23
1993
2631
379
Alderson
MA
Smith
CA
Tough
TW
Davis-Smith
T
Armitage
RJ
Falk
B
Roux
E
Baker
E
Sutherland
GR
Molecular and biological characterization of human 4-1BB and its ligand.
Eur J Immunol
24
1994
2219
380
DeBenedette
MA
Chu
NR
Pollok
KE
Hurtado
J
Wade
WF
Kwon
BS
Watts
TH
Role of 4-1BB ligand in costimulation of T lymphocyte growth and its upregulation on M12 B lymphoma by cAMP.
J Exp Med
181
1995
985
381
Morimoto
C
Rudd
CE
Letvin
NL
Hagan
M
Schlossman
SF
2H4: A novel antigen involved in T lymphocyte triggering.
J Immunol
140
1988
2165
382
Bowen
MA
Patel
DD
Li
X
Modrell
B
Malacko
AR
Wang
W
Harquardt
H
Neubauer
M
Pesando
JM
Francke
U
Haynes
BF
Aruffo
A
Cloning, mapping and characterization of activated leukocyte-cell adhesion molecule (ALCAM), a CD6 ligand.
J Exp Med
181
1995
2213
383
Gagemi
RM
Swack
JA
Gaviria
DM
Romain
PL
Anti-T12, an anti-CD6 monoclonal anitbody, can activate human T lymphocytes.
J Immunol
143
1989
2439
384
Bott
CM
Doshi
JB
Morimoto
C
Romain
PL
Fox
DA
Activation of human T cells through CD6: Functional effects of a novel anti-CD6 monoclonal antibody and definition of four peptides of the CD6 glycoprotein.
Int Immunol
7
1993
783
385
Cocks
BG
Chang
CJ
Carballido
JM
Yssel
H
de Vries
JE
Aversa
G
A novel receptor involved in T-cell activation.
Nature
376
1995
260
386
Navabi
N
Freeman
GJ
Gault
A
Godfrey
D
Nadler
LM
Glimcher
LH
Signalling through the MHC class II cytoplasmic domain is required for antigen presentation and induced B7 expression.
Nature
360
1992
266
387
Koide
SL
Inaba
K
Steinman
RM
Interleukin 1 enhances T-dependent immune responses by amplifying the function of dendritic cells.
J Exp Med
165
1987
515
388
Vakkila
J
Sihvola
M
Hurme
M
Human peripheral blood-derived dendritic cells do not produce IL-1α , IL-1β or IL-6.
Scand J Immunol
31
1990
345
389
Zhou
L
Tedder
TF
A distinct pattern of cytokine gene expression by human CD83+ blood dendritic cells.
Blood
86
1995
3295
390
Heufler
C
Topar
G
Koch
F
Trockenbacher
B
Kampgen
E
Romani
N
Schuler
G
Cytokine gene expression in murine epidermal cell suspensions: interleukin 1β and macrophage inflammatory protein 1α are selective expressed in Langerhans cells but are differentially regulated in culture.
J Exp Med
176
1992
1221
391
Matsue
H
Cruz
PD
Bergstresser
PR
Takashima
A
Cytokine expression by epidermal cell subpopulations.
J Invest Dermatol
42
1992
42
392
McKenzie
JL
Calder
VC
Starling
GC
Hart
DNJ
Role of tumor necrosis factor α in dendritic cell mediated primary mixed leukocyte reactions.
Bone Marrow Transplant
15
1995
163
393
Shalaby
MR
Espevik
T
Rice
GC
Ammann
AS
Figari
IS
Ranges
GE
Palladino
MA
The involvement of TNFα and TNFβ in the MLR.
J Immunol
141
1988
499
394
Cumberbatch
M
Dearman
RJ
Kimber
I
Constitutive and inducible expression of interleukin-6 by Langerhans cells and lymph node dendritic cells.
Immunology
87
1996
513
395
Ghanekar
S
Zheng
L
Logar
A
Navratil
J
Browski
L
Gupta
P
Rinaldo
C
Cytokine expression by human peripheral blood dendritic cells stimulated in vitro with HIV-1 and herpes simplex virus.
J Immunol
157
1996
4028
396
Ming
JE
Steinman
RM
Granelli-Piperno
A
IL-6 enhances the generation of cytotoxic T lymphocytes in the allogeneic MLR.
Clin Exp Immunol
89
1992
148
397
Goodwin
RG
Lupton
S
Schmierer
A
Hjerrild
KJ
Jerzy
R
Clevenger
W
Gillis
S
Cosman
D
Namen
AE
Human interleukin 7: Molecular cloning and growth factor activity on human and murine B-lineage cells.
Proc Natl Acad Sci USA
86
1989
302
398
Murray
R
Suda
T
Wrighton
N
Lee
F
Zlotnik
A
IL-7 is a growth and maintenance factor for mature and immature thymocyte subsets.
Int Immunol
1
1989
526
399
Suda
T
Zlotnik
A
IL-7 maintains the T cell precursor potential of CD3−CD4−CD8− thymocytes.
J Immunol
146
1991
3068
400
Armitage
RJ
Namen
AE
Sassenfeld
HM
Grabstein
KH
Regulation of human T cell proliferation by IL-7.
J Immunol
144
1990
938
401
Grabstein
GH
Namen
AE
Shanebeck
K
Voice
RF
Reed
SG
Widmer
MB
Regulation of T cell proliferation by IL-7.
J Immunol
144
1990
3015
402
Mehrotra
PT
Grant
AJ
Siegel
JP
Synergistic effects of IL-7 and IL-12 on human T cell activation.
J Immunol
154
1995
5093
403
Sorg RV, McLellan AD, Hock BD, Fearnley DB, Hart DNJ: Expression and functional relevance of interleukin-7 for human peripheral blood dendritic cell T lymphocyte stimulation. Immunobiology 1997 (in press)
404
Heufler
C
Koch
F
Stanzl
U
Topar
G
Wysocka
M
Trinchiera
G
Enk
A
Steinman
RM
Romani
N
Schuler
G
Interleukin-12 is produced by dendritic cells and mediates T helper 1 development as well as interferon-γ production by T helper 1 cells.
Eur J Immunol
26
1996
649
405
Yawalkar
N
Brand
CU
Braathen
LR
IL-12 gene expression in human skin-derived CD1a(+) dendritic lymph cells.
Arch Dermatol
288
1996
79
406
Vezzio
N
Sarfati
M
Yang
LP
Demeure
CE
Delespesse
G
Human Th2-like cell clones induce IL-12 production by dendritic cells and may express several cytokine profiles.
Int Immunol
8
1996
1963
407
Grabstein
KH
Eisenman
J
Shanebeck
K
Rauch
C
Srinivasan
S
Fung
V
Beers
C
Richardson
J
Schoenborn
MA
Ahdieh
Me
Cloning of a T cell growth factor that interacts with the β chain of the interleukin-2 receptor.
Science
264
1994
965
408
Doherty
TM
Seder
RA
Sher
A
Induction and regulation of IL-15 expression in murine macrophages.
J Immunol
156
1996
735
409
Jonuleit
H
Wiedemann
K
Muller
G
Degwert
J
Hoppe
U
Knop
J
Enk
AH
Induction of IL-15 messenger RNA and protein in human blood-derived dendritic cells.
J Immunol
158
1997
2610
410
Kampgen
E
Koch
F
Heufler
C
Eggert
A
Gill
L
Gillis
S
Dower
S
Romani
N
Schuler
G
Understanding the dendritic cell lineage through a study of cytokine receptors.
J Exp Med
179
1994
1776
411
Larregina
A
Morelli
A
Kolkowski
E
Fainboim
L
Flow cytometric analysis of cytokine receptors on human Langerhans cells. Changes observed after short-term culture.
Immunology
87
1996
317
412
Ryffel
B
Brockhaus
M
Greiner
B
Mihatsch
MJ
Gudat
F
TNF receptor distribution in human lymphoid tissue.
Immunology
74
1991
446
413
Koch
F
Huefler
G
Kampgen
E
Schneeweiss
D
Boch
G
Schuler
G
Tumor necrosis factor α maintains the viability of murine epidermal Langerhans cells in culture but in contrast to granulocyte/macrophage colony-stimulatory factor, does not induce their functional maturation.
J Exp Med
171
1990
159
414
Steiner
G
Tschachler
E
Tani
M
Malek
TR
Schevach
EM
Holter
W
Knapp
W
Wolff
K
Stingl
G
Interleukin-2 receptors on cultured murine epidermal Langerhans cells.
J Immunol
137
1986
155
415
Xu
S
Ariizumi
K
Edelbaum
D
Bergstresser
PR
Takashima
A
Cytokine-dependent regulation of growth and maturation in murine epidermal dendritic cell lines.
Eur J Immunol
25
1995
1018
416
Chang
C
Furue
M
Tamaki
K
Selective regulation of ICAM-1 and major histocompatibility complex class I and II molecule expression on epidermal Langerhans cells by some of the cytokines released by keratinocytes and T cells.
Eur J Immunol
24
1994
2889
417
Gajewski
TF
Pinnas
M
Wong
T
Fitch
FW
Murine Th1 and Th2 clones proliferate optimally in response to distinct antigen-presenting cell populations.
J Immunol
146
1991
1750
418
Freeman
GJ
Boussiotis
VA
Anumanthan
A
Bernstein
GM
Ke
X
Rennert
PD
Gray
GS
Gribben
JG
Nadler
LM
B7-1 and B7-2 do not deliver identical costimulatory signals, since B7-2 but not B7-1 preferentially costimulates the initial production of IL-4.
Immunity
2
1995
523
419
Ohshima
Y
Delepesse
G
T cell-derived IL-4 and dendritic cell-derived IL-12 regulate the lymphokine-producing phenotype of alloantigen-primed naive human CD4 T cells.
J Immunol
158
1997
629
420
Roth
R
Spiegelberg
HL
Activation of cloned human CD4(+) TH-1 and TH-2 cells by blood dendritic cells.
Scand J Immunol
43
1996
646
421
Morikawa
Y
Furotani
M
Kuribayashi
K
Matsura
N
Kakudo
K
The role of APC in the regulation of delayed-type hypersensitivity. 1 Spleen dendritic cells.
Immunology
77
1992
81
422
Morikawa
Y
Tohya
K
Ishida
H
Matsuura
N
Kakudo
K
Different migration patterns of antigen-presenting cells correlate with Th1/Th2-type responses in mice.
Immunology
85
1995
575
423
Ronchese
F
Hausmann
B
LeGros
G
Interferon-γ and interleukin-4-producing cells can be primed on dendritic cells in vivo and do not require the presence of B cells.
Eur J Immunol
24
1994
1148
424
Everson
MP
McDuffie
DS
Lemak
DG
Koopman
WJ
McGhee
JR
Beagley
KW
Dendritic cells from different tissues induce production of different T cell cytokine profiles.
J Leukoc Biol
59
1996
494
425
Dillon
SM
Hart
DNJ
Abernethy
N
Watson
JD
Baird
MA
Murine antigen presenting cells generated in vitro with GM-CSF or GM-CSF and IL-4 prime naive T lymphocytes to myobacterial antigens poorly in vivo.
Scand J Immunol
46
1997
1
426
Zitvogel
L
Mayordomo
JI
Tjandrawan
T
DeLeo
AB
Therapy of murine tumors with tumor peptide-pulsed dendritic cells: dependence on T cells, B7 costimulation and T helper cell 1-associated cytokines.
J Exp Med
183
1996
87
427
Gabrilovich
DI
Nadaf
S
Corak
J
Berzofsky
JA
Carbone
DP
Dendritic cells in antitumor immune responses.
Cell Immunol
170
1996
111
428
Mayordomo
JI
Loftus
DJ
Sakamoto
H
De Cesare
CM
Appasamy
PM
Lotze
MT
Storkus
WJ
Appella
E
De Leo
AB
Therapy of murine tumors with p53 wide-type and mutant sequence peptide-based vaccines.
J Exp Med
183
1996
1357
429
Santiago-Schwarz
F
Belilos
E
Diamond
B
Carsons
SE
TNFα in combination with GM-CSF enhances the differentiation of neonatal cord blood stem cells into dendritic cells and macrophages.
J Leukoc Biol
52
1992
274
430
Caux
C
Massacrier
C
Duzetter-Dambuyant
C
Vanbervliet
B
Jacquet
C
Schmitt
D
Banchereau
J
Human dendritic Langerhans cells generated in vitro from CD34+ progenitors can prime naive CD4+ T cells and process soluble antigen.
J Immunol
155
1995
5427
431
Bernhard
H
Disis
ML
Heimfeld
S
Hand
S
Gralow
JR
Cheever
MA
Generation of immunostimulatory dendritic cells from human CD34+ hematopoietic progenitor cells in the bone marrow and peripheral blood.
Cancer Res
55
1995
1099
432
Mackensen
A
Herbst
B
Kohlter
G
Wolff-Vorbeck
G
Rosenthal
F
Veelken
H
Kulmburg
P
Schaefer
HE
Mertelsmann
R
Lindemann
A
Delineation of the dendritic cell lineage by generating large numbers of birbeck granule-positive Langerhans cells from human peripheral blood progenitor cells in vitro.
Blood
86
1995
2699
433
Rosenzwajg
M
Canque
B
Gluckman
JC
Human dendritic cell differentiation pathway from CD34+ hematopoietic precursor cells.
Blood
87
1996
535
434
Strobl
H
Riedl
E
Scheinecker
C
Bello-Fernandez
C
Pickl
WF
Rappersberger
K
Majdic
O
Knapp
W
TGF-β1 promotes in vitro development of dendritic cells from CD34+ hemopoietic progenitors.
J Immunol
157
1996
1499
435
Romani
N
Gruner
S
Brang
D
Kampgen
E
Lenz
A
Trockenbacher
B
Konwalinka
G
Fritsch
PO
Steinman
RM
Schluer
G
Proliferating dendritic cell progenitors in human blood.
J Exp Med
180
1994
83
436
Peters
JH
Ruhl
S
Friedrichs
D
Veiled accessory cells deduced from monocytes.
Immunobiology
176
1987
154
437
Rossi
G
Heveker
N
Thiele
B
Gelderblom
H
Steinbach
F
Development of a Langerhans cell phenotype from peripheral blood monocytes.
Immunol Lett
31
1992
189
438
Morel
A
Quaratino
S
Douek
DC
Londei
M
Split activity of interleukin-10 on antigen capture and antigen presentation by human dendritic cells: definition of a maturative step.
Eur J Immunol
27
1997
26
439
Cohen
PL
Kaplan
AM
Adherent Ia+ murine cell lines with characteristics of dendritic cells.
Cell Immunol
80
1983
349
440
Clayberger
C
DeKruyff
RH
Fay
R
Pavlakis
M
Cantor
H
Immunoregulation of T-dependent responses by a cloned dendritic cell.
J Immunol
133
1984
1174
441
O'Neill
HC
Ni
K
Characterization of unique lymphoid cells derived from murine spleen which constitutively produce interleukin-6.
Immunology
79
1993
220
442
Lutz
MB
Aβmann CU
Girolomoni
G
Ricciardi-Castagnoli
P
Different cytokines regulate antigen uptake and presentation of a precursor dendritic cell line.
Eur J Immunol
26
1996
586
443
Shen
Z
Reznikoff
G
Dranoff
G
Rock
KL
Cloned dendritic cells can present exogenous antigens on both MHC class I and class II molecules.
J Immunol
158
1997
2723
444
Santiago-Schwartz
F
Coppock
DL
Hindenburg
AA
Kern
J
Identification of a malignant counterpart of the monocyte-dendritic cell progenitor in an acute myeloid leukemia.
Blood
84
1994
3054
445
Misery
L
Campos
L
Dezutter-Dambuyant
C
Guyotat
D
Treille
D
Schmitt
D
Thivolet
J
CD1-reactive leukemic cells in bone marrow: Presence of Langerhans cell marker on leukaemic monocytic cells.
Eur J Haematol
48
1992
27
446
Choudhury
A
Gajewski
JL
Liang
JC
Popat
U
Claxton
DF
Kliche
K
Andreeff
M
Champlin
RE
Use of leukemic dendritic cells for the generation of antileukemic cellular cytotoxicity against Philadelphia chromosome-positive chronic myelogenous leukemia.
Blood
89
1997
1133
447
Chu
T
Histiocytosis syndromes in children.
Lancet
1
1987
208
448
Egeler
RM
Neglia
JP
Puccetti
DM
Brennan
CA
Nesbit
ME
Association of Langerhans cell histiocytosis with malignant neoplasms.
Cancer
71
1993
865
449
Berti
E
Gianotti
R
Alessi
E
Unusual cutaneous histiocytosis expressing an intermediate immunophenotype between Langerhans' cells and dermal macrophages.
Arch Dermatol
124
1988
1250
450
Vasef
MA
Zaatari
GS
Chan
WC
Sun
NCJ
Weiss
LM
Brynes
RK
Dendritic cell tumors associated with low-grade B-cell malignancies.
Am J Clin Pathol
104
1995
696
451
Weiss
LM
Berry
GJ
Dorfman
RF
Spindle cell neoplasms of lymph nodes of probable reticulum cell lineage: True reticulum cell sarcoma?
Am J Surg Pathol
145
1990
405
452
Ellis
PA
Hart
DNJ
Colls
BM
Nimmo
JC
MacDonald
JE
Angus
HB
Hodgkins cells express a novel pattern of adhesion molecules.
Clin Exp Immunol
90
1992
117
453
Kennedy
ICS
Hart
DNJ
Colls
BM
Nimmo
JC
Willis
DA
Angus
HB
Nodular sclerosing, mixed cellularity and lymphocyte-depleted variants of Hodgkin's disease are probable dendritic cell malignancies.
Clin Exp Immunol
76
1989
324
454
Sorg
UR
Morse
TM
Patton
WN
Hock
BD
Angus
HB
Robinson
BA
Colls
BM
Hart
DNJ
Hodgkin's cells express CD83, a dendritic cell lineage associated marker.
Pathology
29
1997
294
455
McKenzie
JL
Egner
W
Calder
VL
Hart
DNJ
Hodgkin's disease cell lines: A model for IL-1 independent accessory cell function.
Immunology
77
1992
345
456
Hock
BD
Hart
DNJ
Cellular protein profiles of the Hodgkin's disease cell lines L428, KM-H2 and HDLM-2: A comparative study.
Leuk Res
16
1992
253
457
Macatonia
SE
Lau
R
Patterson
S
Pinching
AJ
Knight
SC
Dendritic cell infection, depletion and dysfunction in HIV-infected individuals.
Immunology
71
1990
38
458
Macatonia
SE
Gompels
M
Pinching
AJ
Patterson
S
Knight
SC
Antigen presentation by macrophages but not by dendritic cells in HIV infection.
Immunology
75
1992
576
459
Soto-Ramirez
LE
Renjifo
B
McLane
MF
Marlink
R
O'Hara
C
Sutthent
R
Wasi
C
Vithayasai
P
Vithayasai
V
Apichartpiyakul
C
Auewarakul
P
Cruz
VC
Chui
D
Osathanondh
R
Mayer
K
Lee
T
Essex
M
HIV-1 Langerhans' cell tropism associated with heterosexual transmission of HIV.
Science
271
1996
1291
460
Blauvelt
A
Clerici
M
Lucey
DR
Steinberg
SM
Yarchoan
R
Walker
R
Shearer
GM
Katz
SI
Functional studies of epidermal Langerhans cells and blood monocytes in HIV-infected persons.
J Immunol
154
1995
3506
461
Cameron
PU
Freudenthal
PS
Barker
JM
Gezelter
S
Inaba
K
Steinman
RM
Dendritic cells exposed to HIV-1 transmit a vigorous cytopathic infection to CD4+ T cells.
Science
257
1992
383
462
Cameron
PU
Forsum
U
Teppler
H
Granelli-Piperno
A
Steinman
RM
During HIV-1 infection most blood dendritic cells are not productively infected and can induce allogeneic CT4+ T cells clonal expansion.
Clin Exp Immunol
88
1992
226
463
Pope
M
Betjes
MGH
Romani
N
Hirmand
H
Cameron
PU
Hoffman
L
Gezelter
S
Schuler
G
Steinman
RM
Conjugates of dendritic cells and memory T lymphocytes from skin facilitate productive infection with HIV-1.
Cell
78
1994
389
464
Langhoff
E
Terwilliger
EF
Bos
HJ
Kalland
KH
Poznansky
MC
Bacon
OML
Hazeltine
WA
Replication of HIV-1 in primary dendritic cell cultures.
Proc Natl Acad Sci USA
88
1988
7998
465
Weiss
D
Li
YX
Orenstein
JM
Fauci
AS
Both a precursor and a mature population of dendritic cells can bind HIV- however, only the mature population that expresses CD80 can pass infection to unstimulated CD4(+) T cells.
J Immunol
155
1995
4111
466
Poulter
LW
Janossy
G
The involvement of dendritic cells in chronic inflammatory disease.
Scand J Immunol
21
1985
401
467
Filgueria
L
Nestle
FO
Rittig
M
Joller
HI
Groscurth
P
Human dendritic cells phagocytose and process Borrelia burgdorferi.
J Immunol
157
1996
2998
468
Roszynska
KE
Spickett
GP
Millrain
M
Edwards
A
Bryant
A
Webster
ADB
Farrant
J
Accessory and T cell defects in acquired and inherited hypogammaglobulinaemia.
Clin Exp Immunol
78
1989
1
469
Knight
SC
Farrant
J
Chan
J
Bryant
A
Bedford
PA
Bateman
C
Induction of autoimmunity with dendritic cells: Studies on thyroiditis in mice.
Clin Immunol Immunopathol
48
1988
277
470
Ohashi
PS
Oehen
S
Buerki
K
Pircher
H
Ohashi
CT
Odermatt
B
Malissen
B
Zinkernagel
KM
Hengartner
H
Ablation of “tolerance” and induction of diabetes by virus infection in viral antigen transgenic mice.
Cell
65
1991
305
471
Clare-Salzler
M
Mullen
Y
Marked dendritic cell-T cell cluster formation in the pancreatic lymph node of the non-obese diabetic mouse.
Immunology
76
1992
478
472
Voorbij
HAM
Jeucken
PHM
Kabel
PJ
De Haan
M
Drexhage
HA
Dendritic cells and scavenger macrophages in pancreatic islets of prediabetic BB rats.
Diabetes
38
1989
1623
473
Clare-Salzier
MJ
Brooks
J
Chai
A
Van Herle
K
Anderson
C
Prevention of diabetes in nonobese diabetic mice by dendritic cell transfer.
J Clin Invest
90
1992
741
474
Kabel
PJ
Voorbij
HAM
DeHaan
M
van der Gaag
RD
Drexhage
HA
Intrathyroidal dendritic cells.
J Clin Endocrinol Metab
65
1988
199
475
Farrant
J
Bryant
AE
Chan
J
Hinsworth
RL
Thyroglobulin treated blood dendritic cells reduce IgG anti-thyroglobulin antibody in vitro in Hashimoto's thyroiditis.
Immunol Immunopathol
41
1986
433
476
Zvaifler
NJ
Steinman
RM
Kaplan
G
Lau
LL
Rivelis
M
Identification of immunostimulatory dendritic cells in the synovial effusions of patients with rheumatoid arthritis.
J Clin Invest
76
1985
789
477
Knight
SC
Fryer
P
Griffiths
S
Harding
B
Dixey
J
Mansell
B
Class II antigens on dendritic cells from synovial fluids of patients with inflammatory arthritis.
Clin Exp Immunol
78
1989
19
478
Stagg
AJ
Harding
B
Hughes
RA
Keat
A
Knight
SC
The distribution and functional properties of dendritic cells in patients with seronegative arthritis.
Clin Exp Immunol
84
1991
66
479
Summers
KL
Daniel
PB
O'Donnell
J
Hart
DNJ
Dendritic cells in synovial fluid chronic inflammatory arthritis lack CD80 surface expression.
Clin Exp Immunol
100
1995
81
480
Summers
K
O'Donnell
J
Williams
LA
Hart
DNJ
Expression and function of CD80 and CD86 costimulator molecules on synovial dendritic cells in chronic arthritic disease.
Arth Rheum
39
1996
1287
481
Thomas
R
Quinn
C
Functional differentiation of dendritic cells in rheumatoid arthritis.
J Immunol
156
1996
3074
482
Thomas
R
Davis
LS
Lipsky
PE
Rheumatoid synovium is enriched in mature antigen-presenting dendritic cells.
J Immunol
153
1994
2613
483
Summers KL, O'Donnell JL, Highton J, Hart DNJ: Rheumatoid synovial fluid TGF-β inhibits dendritic cell-T lymphocyte interactions. 1997 (submitted)
484
Demidem
A
Taylor
JR
Grammer
SF
Streilein
JW
T-lymphocyte-activating properties of epidermal antigen-presenting cells from normal and psoriatic skin: Evidence that psoriatic epidermal antigen-presenting cells resemble cultured normal Langerhans cells.
J Invest Dermatol
97
1991
454
485
Holt
PG
Oliver
J
McMenamin
C
Schon-Hegrad
MA
Studies on the surface phenotype and functions of dendritic cells in parenchymal lung tissue of the rat.
Immunology
75
1992
582
486
Muller
HK
Bucana
C
Kripke
ML
Antigen-presentation in the skin: Modulation by UV radiation and chemical carcinogens.
Semin Immunol
4
1992
205
487
Kin
T
Kripke
ML
Ullrich
SE
Immunosuppression by factors released from UV-irradiated epidermal cells: Selective effects on the generation of contract and delayed hypersensitivity after exposure to UVA or UVB radiation.
J Invest Dermatol
94
1990
26
488
Halliday
GM
Reeve
VE
Barnetson
RS
Langerhans cell migration into ultraviolet light-induced squamous skin tumors is unrelated to anti-tumor immunity.
J Invest Dermatol
97
1991
830
489
Ulrich
SE
Modulation of immunity by ultraviolet radiation: Key effects on antigen presentation.
J Invest Dermatol
105
1995
30S
490
Yoshikawa
T
Rae
V
Bruins-Slot
W
Van den Berg
J
Taylor
JR
Streinlein
JW
Susceptibility to effects of UVB radiation on induction of contact hypersensitivity as a risk factor for skin cancer in human.
J Invest Dermatol
95
1996
530
491
Stene
MA
Babajanians
M
Bhuta
S
Cochran
AJ
Quantitative alterations in cutaneous Langerhans cells during evolution of malignant melanoma of the skin.
J Invest Dermatol
91
1988
125
492
Smolle
J
Soyer
H
Ehall
R
Barstenstein
S
Kerl
H
Langerhans cell density in epithelial skin tumors correlates with epithelial differentiation but not with the pretumoral infiltrate.
J Invest Dermatol
87
1986
477
493
Nestle
FO
Burg
G
Fah
J
Wrone-Smith
T
Nickoloff
BJ
Human sunlight-induced basal-cell-carcinoma-associated dendritic cells are deficient in T cell co-stimulatory molecules and are impaired as antigen-presenting cells.
Am J Pathol
150
1997
641
494
Meissner
K
Michaelis
K
Rehpenning
W
Loning
T
Epidermal Langerhans cell densities influence survival in mycosis fungoides and Sezary syndrome.
Cancer
65
1990
2069
495
Wilson
AJ
Maddox
PH
Jenkins
D
CD1a and S100 antigen expression in skin Langerhans cells in patients with breast cancer.
J Pathol
163
1991
25
496
Ambe
K
Mori
M
Enjoji
M
S-100 protein-positive dendritic cells in colorectal adenocarcinomas. Distribution and relation to the clinical prognosis.
Cancer
63
1989
496
497
Tsujitani
S
Kakeji
Y
Watanabe
A
Kohnoe
S
Maehara
Y
Sugimachi
K
Infiltration of dendritic cells in relation to tumor invasion and lymph node metastasis in human gastric cancer.
Cancer
66
1990
2012
498
Nomori
H
Watanabe
S
Nakajima
T
Shiosata
Y
Kameya
T
Histiocytes in nasopharyngeal carcinoma in relation to prognosis.
Cancer
57
1986
100
499
Furukawa
T
Watanabe
S
Kodama
T
Sato
Y
Shimosato
Y
Suemasu
K
T-zone histiocytes in adenocarcinoma of the lung in relation to postoperative prognosis.
Cancer
56
1985
2651
500
Schroder
S
Schwarz
W
Rehpenning
W
Loning
T
Bocker
W
Dendritic/Langerhans cells and prognosis in patients with papillary carcinoma. Immunohistochemical study of 106 thyroid neoplasms correlated to follow up data.
Am J Clin Pathol
99
1988
295
501
Bigotti
G
Coli
A
Castagnola
D
Distribution of Langerhans cells and HLA class II molecules in prostatic carcinomas of different histopathological grade.
Prostate
19
1991
73
502
Shimizu
J
Zou
J
Ikegame
K
Katagiri
T
Fujiware
H
Hamaoka
T
Evidence for the functional binding in vivo of tumor rejection antigens to antigen-presenting cells in tumor-bearing hosts.
J Immunol
146
1991
1708
503
Gabrilovich
DI
Ciernik
F
Carbone
DP
Dendritic cells in antitumor immune responses. I. Defective antigen presentation in tumor-bearing hosts.
Cell Immunol
170
1996
101
504
Gabrilovich
DI
Chen
HL
Girgis
KR
Cunningham
HT
Meny
GM
Nadaf
S
Kavanaugh
D
Carbone
DP
Production of vascular endothelial growth factor by human tumors inhibits the functional maturation of dendritic cells.
Nature Med
2
1996
1096
505
Pamplilon
DH
Alnagdy
AA
Wallington
TB
Immunomodulation by UV light: Clinical studies and biological effects.
Immunol Today
12
1991
119
506
Roberts
MS
Knight
SC
Low-dose immunosuppression by cyclosporine operating via antigen-presenting dendritic cells.
Transplantation
50
1990
91
507
Faustman
D
Steinman
RM
Gebel
H
Hauptfeld
V
Davis
J
Lacy
P
Prevention of rejection of murine islet allografts by pretreatment with anti-dendritic cell antibody.
Proc Natl Acad Sci USA
81
1984
3864
508
Iwai
H
Kuma
S
Inaba
MM
Acceptance of murine throid allografts by pretreatment of anti-Ia antibody or anti-dendritic cell antibody.
Transplantation
47
1989
45
509
Jenkinson
EJ
Benson
MT
Buckley
G
Owen
JJT
Survival of deoxyguanosine-treated fetal thymus allografts is prevented by priming with dendritic cells.
Immunology
60
1987
593
510
Brewer
Y
Bewick
M
Palmer
A
Severn
A
Welsh
K
Taube
D
Prevention of renal allograft rejection by perfusion with anti-leukocyte monoclonal antibody is dependent on good uptake of antibody by interstitial DC.
Trans Proc
21
1989
1772
511
Freeman
AJ
Baird
MA
Immunosuppression mediated by heat-treated cells: suppression or anergy?
Transplantation
55
1992
1439
512
Fu
F
Li
Y
Qian
S
Lu
L
Chambers
F
Starzl
TE
Fung
JJ
Thomson
AW
Costimulatory molecule-deficient dendritic cell progenitors (MHC class II+, CD80dim, CD86−) prolong cardiac allograft survival in nonimmunosuppressed recipients.
Transplantation
62
1996
659
513
van Lochem
E
van der Keur
M
Mommaas
AM
de Gast
GC
Goulmy
E
Functional expression of minor histocompatibility antigens on human peripheral blood dendritic cells and epidermal Langerhans cells.
Trans Immunol
4
1996
151
514
Perreault
C
Pelletier
M
Belanger
R
Boileau
J
Bonny
Y
David
M
Gyger
M
Landry
D
Montplaisir
S
Persistence of host Langerhans cells following allogeneic bone marrow transplantation: Possible relationship with acute graft-versus-host disease.
Br J Haematol
60
1985
253
515
Ferrara
JLM
Paradigm shift for graft-versus-host disease.
Bone Marrow Transplant
14
1994
183
516
Gribben
JG
Guinan
EC
Boussiotis
VA
Ke
X
Linsley
L
Sieff
C
Gray
GS
Freeman
GJ
Nadler
LM
Complete blockade of B7 family-mediated costimulation is necessary to induce human alloantigen-specific anergy: A method to ameliorate graft-versus-host disease and extend the donor pool.
Blood
87
1996
4887
517
Si
Z
Hersey
P
Coates
AS
Clinical responses and lymphoid infiltrates in metastatic melanoma following treatment with intralesional GM-CSF.
Melanoma Res
6
1996
247
518
Corr
M
Lee
DJ
Carson
DA
Tighe
H
Gene vaccination with naked plasmid DNA: Mechanism of CTL priming.
J Exp Med
184
1996
1555
519
Knight
SC
Hunt
R
Dore
C
Medawar
PB
Influence of dendritic cells in tumor growth.
Proc Natl Acad Sci USA
82
1985
4495
520
Bhardwaj
N
Bender
A
Gonzalez
N
Kim
Bui L
Garrett
MC
Steinman
RM
Influenza virus-infected dendritic cells stimulate strong proliferative and cytolytic responses from human CD8+ T cells.
J Clin Invest
94
1994
797
521
Manca
F
Giuseppina
LP
Fenoglio
D
Fang
SP
Habeshaw
A
Knight
SC
Dalgleish
AG
Dendritic cells are potent antigen-presenting cells for in vitro induction of primary human CD4+ T-cell lines specific for HIV gp120.
J Acq Immunodef Syndromes
7
1994
15
522
Cohen
PJ
Cohen
PA
Rosenberg
SA
Katz
SI
Mule
JJ
Murine epidermal Langerhans cells and splenic dendritic cells present tumor-associated antigens to primed T cells.
Eur J Immunol
24
1994
315
523
Alijagic
SA
Moller
P
Artuc
M
Jurgovsky
K
Czarnetzki
BM
Schadendorf
D
Dendritic cells generated from peripheral blood transfected with human tyrosinase induce specific T cell activation.
Eur J Immunol
25
1995
3100
524
Tjoa
B
Boynton
A
Kenny
G
Ragde
H
Misrock
SL
Murphy
G
Presentation of prostate tumor antigens by dendritic cells stimulates T-cell proliferation and cytotoxicity.
Prostate
28
1996
65
525
Grabbe
S
Bruvers
S
Gallo
RL
Knisely
TL
Nazareno
R
Granstein
RD
Tumor antigen presentation by murine epidermal cells.
J Immunol
146
1991
3656
526
Gong
J
Chen
D
Kashiwaba
M
Kufe
D
Induction of antitumour activity by immunization with fusions of dendritic and carcinoma cells.
Nature Med
3
1997
558
527
Mandelboim
O
Vadai
E
Fridkin
M
Katz-Itillel
A
Feldman
M
Berke
G
Eisenbach
L
Regression of established murine carcinoma metastases following vaccination with tumor associated antigen peptides.
Nature Med
1
1995
1179
528
Paglia
P
Chiodoni
C
Rodolfo
M
Colombo
MP
Murine dendritic cells loaded in vitro with soluble protein prime cytotoxic T lymphocytes against tumor antigen in vivo.
J Exp Med
183
1996
317
529
Porgador
A
Snyder
D
Gilboa
E
Induction of antitumor immunity using bone marrow-generated dendritic cells.
J Immunol
156
1996
2918
530
Apostolopoulos
V
McKenzie
IFC
Cellular mucins: Targets for immunotherapy.
Crit Rev Immunol
14
1994
293
531
Apostolopoulos V, Pietersz GA, Loveland BE, Sandrin MS, McKenzie IFC: Oxidative/reductive conjugation of mannan to antigen selects for T1 or T2 immune responses. Proc Natl Acad Sci USA 92:1995
532
Hsu
FJ
Benike
C
Fagnoni
F
Liles
TM
Czerwinski
D
Taidi
B
Engelman
EG
Levy
R
Vaccination of patients with B-cell lymphoma using autologous antigen-pulsed dendritic cells.
Nature Med
2
1996
52
533
Murphy
G
Tjoa
B
Ragde
H
Kenny
G
Boynton
A
Phase I clinical trial: T-cell therapy for prostate cancer using autologous dendritic cells pulsed with HLA-A0201-specific peptides from prostate-specific membrane antigen.
Prostate
29
1996
371
534
Suzue
K
Young
RA
Adjuvant-free hsp70 fusion protein system elicits humoral and cellular immune responses to HIV-1 p24.
J Immunol
156
1996
873
535
Flamand
V
Sornasse
T
Thielemans
K
Demanet
C
Bakkus
M
Bazin
H
Tielemans
F
Leo
O
Urbain
J
Moser
M
Murine dendritic cells pulsed in vitro with tumor antigen induce tumor resistance in vivo.
Eur J Immunol
24
1994
605
536
Ossevoort
MA
Feltkamp
MCW
van Veen
KJH
Melief
CJM
Kast
WM
Dendritic cells as carriers for a cytotoxic T-lymphocyte epitope-based peptide vaccine in protection against a human papilloma virus type 16-induced tumor.
J Immunother
18
1995
86
537
Paglia
P
Chiodoni
C
Rodolfo
R
Colombo
MP
Murine dendritic cells loaded in vitro with soluble protein prime cytotoxic T lymphocytes against tumor antigen.
J Exp Med
183
1996
317
538
Svensson
M
Stockinger
B
Wick
MJ
Bone marrow-derived dendritic cells can process bacteria for MHC-I and MHC-II presentation to T cells.
J Immunol
158
1997
4229
539
Power
CA
Church
DJ
Meyer
A
Alouani
S
Proudfoot
AEI
Clark-Lewis
I
Sozzani
S
Mantovani
A
Wells
TNC
Cloning and characterization of a specific receptor for the novel CC chemokine MIP-3α from lung dendritic cells. J.
Exp Med
186
1997
825
540
Greaves
DR
Wang
W
Daifaghi
DJ
Dieu
MC
de Saint-Vis
B
Franz-Bacon
K
Rossi
D
Caux
C
McClanahan
T
Gordon
S
Slotnik
A
Schall
TJ
CCR6, a CC chemokine receptor that interacts with macrophage inflammatory protein 3α and is highly expressed in human dendritic cells.
J Exp Med
186
1997
837
541
Adema
GJ
Hartgers
F
Verstraten
R
de Vreis
E
Marland
G
Menon
S
Foster
J
Xu
Y
Nooyen
P
McClanahan
T
Bacon
KB
Figdor
CG
A dendritic-cell derived C-C chemokine that preferentially attracts naive T cells.
Nature
387
1997
713
542
Godiska
R
Chantry
D
Raport
CJ
Sozzani
S
Allavena
P
Leviten
D
Mantovani
A
Gray
PW
Human macrophage-derived chemokine (MDC), a novel chemoattractant for monocytes, monocyte-derived dendritic cells, and natural killer cells.
J Exp Med
185
1997
1595
543
Lynch
DH
Andreasen
A
Maraskovsky
E
Whitmore
J
Miller
RE
Schuh
JCL
Flt3 ligand induces tumor regression and antitumor immune responses in vivo.
Nature Med
3
1997
625
544
Lyman SD, Brasel K, Maraskovsky E, Miller RE, Schuh J, Maliszewski C, Pulandran B, Stocking K, Peschon J, Lynch D, McKenna H: Biological activities of flt-3 ligand. Exp Hematol 25:729, 1997 (abstr)
545
Grohmann
U
Nianchi
R
Ayroldi
E
Belladonna
ML
Surace
D
Fioretti
MC
Puccetti
P
A tumor-associated and self antigen peptide presented by dendritic cells may induce T cell anergy in vivo, but IL-12 can prevent or revert the anergic state.
J Immunol
158
1997
3593
Sign in via your Institution